Open Access
How to translate text using browser tools
18 February 2011 Glutathione
Graham Noctor, Guillaume Queval, Amna Mhamdi, Sejir Chaouch, Christine H. Foyer
Author Affiliations +
Abstract

Glutathione is a simple sulfur compound composed of three amino acids and the major non-protein thiol in many organisms, including plants. The functions of glutathione are manifold but notably include redox-homeostatic buffering. Glutathione status is modulated by oxidants as well as by nutritional and other factors, and can influence protein structure and activity through changes in thiol-disulfide balance. For these reasons, glutathione is a transducer that integrates environmental information into the cellular network. While the mechanistic details of this function remain to be fully elucidated, accumulating evidence points to important roles for glutathione and glutathione-dependent proteins in phytohormone signaling and in defense against biotic stress. Work in Arabidopsis is beginning to identify the processes that govern glutathione status and that link it to signaling pathways. As well as providing an overview of the components that regulate glutathione homeostasis (synthesis, degradation, transport, and redox turnover), the present discussion considers the roles of this metabolite in physiological processes such as light signaling, cell death, and defense against microbial pathogen and herbivores.

1. HISTORICAL OVERVIEW

In 1888 de Rey-Pailhade discovered that yeast cells contain a substance that he named “philothion”, which reacts spontaneously with elemental sulfur to give hydrogen sulfide (reviewed by Meister, 1988a). His subsequent studies etablished the widespread presence of philothion in animal and plant tissues, and led to the conclusion that the compound contained cysteine that can undergo reversible oxidation to a disulfide form in the presence of oxygen (de Rey-Pailhade, 1928). A compound (subsequently accepted to be philothion) was extracted from muscle tissue in 1921 by Hopkins, who reported that it was autooxidizable and contained glutamic acid as well as cysteine. In the absence of a definitive composition, he suggested that “provisionally, for easy reference, the name glutathione will perhaps be admissible” (Hopkins, 1921). Hopkins' view that glutathione was a dipeptide of glutamic acid and cysteine was challenged by a report that the compound was possibly a tripeptide (Hunter and Eagles, 1927). Two years later, the development of a new procedure for preparing the compound in crystalline form confirmed that glutathione was indeed a tripeptide of glutamic acid, cysteine, and glycine (Hopkins, 1929; Figure 1). In the same year, Hopkins was awarded the Nobel Prize for this and other work on vitamins and related nutritional factors.

The path to the characterization of glutathione functions in plants was opened up by another Nobel Prize laureate working on vitamins. Szent-Gyorgyi (1931) observed that cabbage leaf tissue was able to reduce the oxidized form of ascorbic acid, a process that involved the simultaneous oxidation of glutathione. Indeed, the early recognition that ascorbic acid is nearly always found in the reduced state in plants, despite the presence of terminal oxidases and oxidizing catalysts, led to the conclusion that a reducing mechanism must be present. In 1936, Hopkins and Morgan concluded that the reducing agent was glutathione (Hopkins and Morgan, 1936).

The growing recognition of the biological importance of reactive oxygen species (ROS) in the 1960's, including the discovery of superoxide, promoted renewed interest in glutathione in plants. This included placing ascorbate-glutathione interactions within a physiological context. Thus, the ascorbate-glutathione pathway was assigned a key function in the metabolism of hydrogen peroxide in the chloroplasts (Foyer and Halliwell, 1976) while glutathione became implicated in stress resistance in general (Esterbauer and Grill, 1978; reviewed by Tausz et al., 2004). At the same time, the roles of chloroplast thiols in enzyme regulation were being characterized (Wolosiuk and Buchanan, 1977). Growing attention was subsequently paid to the details of glutathione synthesis, degradation, transport and compartmentation (Rennenberg, 1982; Alscher, 1989).

With the advent of cloning and transgenic technology, the recognition of the antioxidant role of glutathione saw the production of plants with enhanced glutathione contents or reduction capacity (Foyer et al., 1991, 1995; Aono et al., 1993; Broadbent et al., 1995; Strohm et al., 1995; Noctor et al., 1996, 1998a,b; Creissen et al., 1999) as well as the identification of genes encoding the enzymes of glutathione synthesis and reduction in plants (May and Leaver, 1994; Creissen et al., 1995; Rawlins et al., 1995; Ullman et al., 1996). The notion that glutathione status could be involved in transmitting oxidative stress signals began to develop (Foyer et al., 1997; May et al., 1998a). Since then, analysis of Arabidopsis mutants has provided convincing demonstrations that glutathione is required for plant development (Vernoux et al., 2000; Cairns et al., 2006). Today, there is unequivocal evidence that glutathione is a multifunctional metabolite in plants that is important in redox homeostasis and signaling, as well as in defense reactions. Some of the most important of these insights have been made possible by the availability of Arabidopsis mutants and the ever-growing body of information generated by genomic and post-genomic approaches. Together, this information places this small thiol compound at the heart of the regulation of plant development and responses to the environment.

Figure 1.

Glutathione: the basics.

(A) Glutathione identity card. (B) Simple overview of the key groups of glutathione-associated genes. The schematic composition of gene types shown in part (B) is based on the latest annotation on the TAIR website ( www.arabidopsis.org). Note that the distribution is not intended to be exhaustive (for instance, potential or identified transporters are not included and other glutathione-dependent proteins no doubt await discovery) while annotated glutathione peroxidases are included in redox turnover, but may preferentially use thioredoxins as reductants. For a list of useful gene identities and associated information, refer to Supplementary Data 1. Information on GSTs can be found in Table 2 of Dixon and Edwards (2010).

2. BIOCHEMICAL AND REDOX PROPERTIES OF GLUTATHIONE

2.1. Structure and Biological Distribution

The thiol tripeptide, glutathione (GSH; γ-L-glutamyl-L-cysteinylglycine), is found in most organisms. Linkage of the γ-carboxyl group of glutamate to the amino group of cysteine distinguishes this bond from peptide bonds in proteins. This is thought to confer stability on the molecule because it allows degradation through specific amino acid transferases (see section 5). In some organisms, such as halobacteria, glutathione can be replaced by other sulfur compounds like thiosulfate (Newton and Javor, 1985). Another example is trypanothione (N1,N8-bis(glutathionyl) spermidine), a compound that performs similar redox functions to glutathione in some parasitic protozoa (Fairlamb et al., 1985). In plants, glutathione typically accumulates in cells to millimolar concentrations, with tissue contents exceeding free cysteine by 10-to 50-fold.

In some plant taxa, homologous glutathione forms are found, as well as GSH. Instead of glycine, these compounds contain other C-terminal amino acids such as serine, β-alanine, or glutamate (Rennenberg, 1982; Klapheck, 1988; Klapheck et al., 1992; Meuwly et al., 1993). In legumes, in which homoglutathione (γGlu-Cys-β-Ala) can be found alongside GSH, the two homologs are synthesized by distinct enzymes encoded by different genes (MacNicol, 1987; Frendo et al., 2001). The path to hydroxymethylGSH (γ-Glu-Cys-Ser), which is found in cereals such as wheat (Klapheck et al., 1992), is less clear. A purified wheat glutathione synthetase (GSH-S) showed no activity against L-serine or D-serine though low activity was reported against certain other L- and D-amino acids as well as glycine (Skipsey et al., 2005a). HydroxymethylGSH could result from modifications of GSH catalyzed by enzymes with transpeptidase activity such as carboxypeptidase Y (Okumura et al., 2003). The disulfide forms of homoglutathione and γ-Glu-Cys-Ser are reducible by glutathione reductase (GR; Klapheck, 1988; Klapheck et al., 1992; Oven et al., 2001). To our knowledge, no GSH homologs have been reported in significant amounts in Arabidopsis, though analysis of purified GSH-S suggest that novel homologs may await discovery in some species (Skipsey et al., 2005a). In poplar overexpressing bacterial GSHS, two novel peaks appeared on HPLC traces and these accumulated most strongly in plants kept in the dark, when glycine contents are low (G. Noctor, A.C.M. Arisi, L. Jouanin, C.H. Foyer, unpublished data). The chemical identity of these peaks remains unknown.

2.2. Redox Reactions

Many GSH functions are linked to reversible redox reactions of the cysteine sulfur group. Like other thiols, glutathione can undergo numerous redox reactions. Oxidized forms notably include disulfides, either with another glutathione cysteine residue to produce glutathione disulfide (GSSG) or with a different thiol to form ‘mixed disulfides’, as well as more oxidized forms in which the thiol group is converted to sulfenic, sulfinic or sulfonic acids (Foyer and Noctor, 2005). However, glutathione biochemistry is not restricted to these compounds. The potential for formation of an array of glutathione conjugates with endogenous and xenobiotic electrophilic species is vast (Wang and Ballatori, 1998; Dixon and Edwards, 2010). In particular, GSNO is receiving increasing attention for its physiological significance as a potential signaling molecule or as a reservoir of NO (Lindermayr et al., 2005).

Compounds able to oxidize GSH at high rates include ROS such as superoxide or the hydroxyl radical. Due to its relatively high cellular concentration, therefore, glutathione serves as a scavenger or sacrificial nucleophile. Thus, glutathione is a highly reducing chemical barrier that prevents excessive oxidation of sensitive cellular components. Many primary and secondary metabolites can act similarly as sacrificial antioxidants. However, unlike glutathione, the oxidized forms of these compounds are either not stable and/or not recycled to the reduced form at high rates. By contrast, glutathione is maintained in a predominantly reduced state by specific GRs found in the cytosol, plastids, mitochondria, and peroxisomes that have high affinity for GSSG and, particularly, NADPH (Halliwell and Foyer, 1978; Smith et al., 1989; Edwards et al., 1990; Jiménez et al., 1997; Chew et al., 2003; Kataya and Reumann, 2010). Further, glutathione oxidation occurs not only chemically but is also catalyzed by enzymes able to use GSH to reduce H2O2 or other peroxides to water or the corresponding alcohol (see section 7.1). Finally, the enzyme dehydroascorbate reductase (DHAR) links glutathione oxidation to ascorbate regeneration, though this reaction can also occur chemically at significant but lower rates (Foyer and Mullineaux, 1998). GSH-dependent reduction of DHA allows NADPH oxidation to be coupled to ROS removal via ascorbate and glutathione pools (Foyer and Halliwell, 1976). While plant antioxidative systems have been shown to be complex (Mittler et al., 2004) and the functions of many components remain to be defined, current information confirms that ascorbate and glutathione act together, alongside catalases, in high capacity redox-homeostatic H2O2-processing pathways.

2.3. How Glutathione is Measured in Plants

A range of useful methods is now available for the determination of glutathione and associated redox reactions in cells. Sensitive quantitative techniques for measuring glutathione in vitro are available by both HPLC and spectrophotometric enzymatic assay, while other methods have been developed to provide information on glutathione in situ (Figure 2). None of these measurements provide exhaustive information. In each case the data generated must be viewed within the context of the limitations imposed by the procedures employed to extract and estimate GSH:GSSG ratios in vitro or to probe glutathione pools in situ.

2.3.1. Enzymatic quantification

The most common spectrophotometric glutathione assay uses 5,5′-dithiobis(2-nitrobenzoic acid), otherwise known as DTNB or Ellman's reagent. Because this compound binds non-specifically to thiol groups, it can be used to measure protein and non-protein thiols alike. In acid extracts, from which proteins have been precipitated, this method measures total non-protein thiols (Figure 2A, bottom). The assay can provide a useful gross estimate of GSH content as this compound is usually the major non-protein thiol in most tissues. However, it should be noted that the method does not specifically measure GSH but all DTNB-reactive thiols present in the extract. It also does not detect GSSG. If NADPH and GR are include in the reaction mix (Figure 2A, middle), DNTB is continuously reduced by electrons provided by NADPH, and glutathione cycles between GSSG and GSH in an intermediate, catalytic role. This technique provides much greater sensitivity and specificity for glutathione than the end-point DTNB assay (Tietze, 1969). Without pre-treatment of extracts, the method measures both GSH and GSSG, ie, total free glutathione. The most common approach to distinguish between the two forms is to treat aliquots with a thiol reagent to remove GSH, then to assay the residual GSSG in the same way as total glutathione. Typical thiol reagents are N-ethylmaleimide and 2-vinylpyridine (Griffith, 1980). These methods can be readily adapted to multi-well assay in plate-readers (Queval and Noctor, 2007).

Figure 2.

Some commonly used methods for measuring glutathione.

(A) Standard in vitro techniques for assaying glutathione and related thiols in protein-free extracts.

(B) Three techniques recently used to measure glutathione in situ. The image in the right middle panel was kindly provided by Bernd Zechmann (University of Graz, Austria). It shows a section through an Arabidopsis wild-type leaf mesophyll cell labelled with an anti-glutathione antibody. The black particles correspond to glutathione immunoprecipitates visible in mitochondria (M), nucleus (N), and plastid (P), and which are much less abundant in the vacuole (V). The scale bar indicates 1 µm. DTNB, 5,5′-dithiobis(2-nitrobenzoic acid). F, fluorescence. GRX, glutaredoxin. GST, glutathione S-transferase. MCB, monochlorobimane.

2.3.2. Liquid chromatography-based assays

The most commonly used HPLC technique is based on fluorescent labelling. Developed during the late 1970's and early 1980's (Fahey et al., 1981), this technique was first used to analyze plant extracts by the Rennenberg group (Schupp and Rennenberg, 1988; Rauser et al., 1991). It involves quantification of thiol-bimane adducts formed by treating extracts with reagents such as monobromobimane (MBB). The advantages of this sensitive technique are that the bimane derivatives have a high fluorescence yield and are quite stable (up to several days in solutions kept in the dark), and that other thiols present in the extract, such as cysteine, can be separated and measured in the same run with glutathione (Figure 2A, top). In the most common application of the technique, aliquots are treated, prior to bimane derivatization, with a reductant such as dithiothreitol (DTT). This pre-treatment converts any GSSG present in the aliquot to GSH, and so only the total glutathione pool is determined, not the redox state. As described above for the spectrophotometric assay, GSSG can be measured by pre-treating extract aliquots with thiol reagents to remove GSH. However, this procedure is somewhat more onerous for the HPLC method because, after GSH is fixed in a complex, GSSG must subsequently be reduced to GSH prior to incubation with MBB. Approximate glutathione redox states can be estimated by simply comparing GS-bimane peak yields in aliquots from the same extract pre-treated or not with reductants like DTT, though the alkaline pH required for reaction with MBB probably causes some artefactual oxidation of extracted GSH if DTT is not present during the derivatization. This method can therefore give somewhat lower reduction states than the spectrophotometric assay (eg, Queval et al., 2011).

Other sensitive HPLC-fluorescence methods can be used based on derivatizing the primary amine group of glutathione. One example is O-phthaldialdehyde (OPA), a reagent that entered widespread use for HPLC analysis of amino acids in the 1980's (Lindroth and Mopper, 1979). Using OPA allows glutathione to be measured along with amino acids (Noctor and Foyer, 1998), though one drawback compared to bimane derivatization is that cysteine is not readily detectable. Other techniques applied to measuring glutathione in plants include LC-MS (Muckenschnabel et al., 2001) and capillary electrophoresis (Mendoza et al., 2008). Compared to targeted thiol analyses, however, these techniques require more specialized equipment and may also be less reproducible and quantitative. Techniques such as LCMS are, however, among the methods of choice for identifying or analyzing GS-conjugates (eg, Brazier-Hicks et al., 2008; Mueller et al., 2008).

2.3.3 Data output

Regardless of the analytical method used to quantify glutathione in extracts, a key factor is sample preparation. The extraction procedure is crucial, and great care has to be taken to ensure that no loss of glutathione or change in redox state occurs. Best results are obtained if samples are quenched and ground in liquid nitrogen and then extracted directly into acid as the material thaws, taking care to homogenize regularly during this procedure. While some neutralization of protein-free acid supernatants is required for most of the techniques, maintaining a slightly acidic pH throughout the sample preparation minimizes GSH oxidation. Once the pH has been adjusted, extracts should be assayed as soon as possible, though from our experience GSH and GSSG are usually quite stable for a few hours in extracts kept on ice. It is not advisable to refreeze extracts, even if kept at acidic pH. For some extracts, this can lead to artefactual oxidation or degradation. As for any metabolite assay, it is advisable to check extraction efficiency, for example, with recovery experiments using known amounts of GSH or GSSG added at the stage of extraction in the presence or absence of the tissue under examination.

Measuring glutathione by enzyme assay or by HPLC-fluorescence with MBB or OPA produces comparable data for tissue contents (Noctor and Foyer, 1998; Queval and Noctor, 2007). Proper application of these methods to analysis of plants grown in controlled conditions gives typical values for leaf glutathione of the order of 200–600 nmol.g1 fresh weight with the exact value depending on developmental stage and nutritional/light regime (eg, Cobbett et al., 1998; Veljovic-Jovanovic et al., 2001; Hartmann et al., 2003; Queval and Noctor, 2007). Higher values may be obtained for plants grown in less controlled conditions, eg, in crop plants growing in the field.

The overall redox state of glutathione [GSH/(GSH + 2GSSG)] in leaf extracts is around 0.9–0.95. Increases in GSSG relative to GSH are a useful indicator of oxidative stress or ‘disulfide stress’ (Åslund and Beckwith, 1999), and in plants can be observed in response to conditions such as ozone, pathogens, and cold, as well as deficiency in important H2O2-metabolizing enzymes such as catalase and ascorbate peroxidase (APX; Smith et al., 1984; Sen Gupta et al., 1991; Edwards et al., 1991; May and Leaver, 1993; May et al., 1996a; Willekens et al., 1997; Vanacker et al., 2000; Kocsy et al., 2000; Riszhky et al., 2002; Gomez et al., 2004a; Queval et al., 2007). A key property of glutathione homeostasis function is that increased oxidation, unless accompanied by appreciable consumption of GSH, is often followed or accompanied by increases in the total pool.

2.3.4. Analyzing glutathione in situ

Several methods have been developed to monitor glutathione in situ. It is important to note that these techniques measure different features of glutathione status, which may be of greater or lesser relevance depending on the information that is sought. Like in vitro analysis of glutathione, all have potential drawbacks and limitations in terms of the information they provide. In some cases, the necessary preparation imposes a stress on the sample, which could modify the factor being measured. Another caveat relates to the use of in vivo techniques under conditions different from those in which the plants are growing (eg, different actinic irradiance in studies where the rate of photosynthesis is important). Nevertheless, the availability of in situ techniques is undoubtably an important development that is adding to the repertoire of tools to analyze the complexities of glutathione homeostasis and related issues of cellular redox metabolism.

Monochlorobimane (MCB) is a less reactive analog of MBB that was used to specifically label the free GSH pool in mammalian cells in situ (Rice et al., 1986). While MBB reacts rapidly and non-specifically with thiol groups in alkaline pH conditions, the chemical reaction of MCB with glutathione is slower and accelerated by endogenous GSTs (Rice et al., 1986). In plants, in situ derivatization of the glutathione pool with MCB was used for fluorescence imaging of epidermal cells (Müller et al., 1999a,b). Subsequent studies in Arabidopsis cells and poplar leaves adapted this technique to quantify extracted GS-bimane in vitro by HPLC (Meyer et al., 2001; Hartmann et al., 2003). This approach has been used to investigate the regulation of glutathione synthesis and breakdown, as well as the compartmentation of glutathione per se (Meyer and Fricker, 2002; Hartman et al., 2003; Grzam et al., 2007; Ohkamu-Ohtsu et al., 2007b; Blum et al., 2010; Queval et al., 2011). Because GSTs are particularly abundant in the cytosol (Dixon et al., 2009), the requirement for these enzymes to obtain rapid labelling may be the most important factor contributing to the potential compartmental specificity of the technique (Hartmann et al., 2003). Based on work in poplar, it was concluded that MCB labels only the cytosolic pool, and that this pool represents about half of total leaf glutathione (Hartmann et al., 2003). In wildtype Arabidopsis leaves, however, the entire leaf glutathione pool appeared to be accessible to MCB labelling within 2h incubation (Queval et al., 2011). Among the many factors that might explain this difference are inter-species specificity in GST compartmentalization or chloroplast glutathione transporters, variation in leaf developmental stage, or the inclusion (or not) of inhibitors during the incubation.

In situ immunolocalization has been used to examine glutathione distribution at the subcellular level by quantitative immunohistochemistry coupled to electron microscopy (Figure 2B). This approach relies on specific antibodies generated against glutaraldehyde-fixed glutathione and on the quantitative evaluation of the obtained signal by stereological methods. Antibodies that detect both GSH and GSSG have been used to label glutathione in several subcellular compartments (Zechmann et al., 2007, 2008; Kolb et al., 2010; Zechmann and Müller, 2010). Interestingly, in both Arabidopsis and tobacco, it was reported that the highest glutathione concentrations were not found in plastids but in mitochondria (Zechmann et al., 2008; Hölleret al., 2010). In wild-type plants, glutathione was found to be negligible or undetectable in the vacuole, in agreement with earlier concepts that glutathione concentrations are usually low in this compartment (Rennenberg, 1982). A leaf glutathione content of 200 nmol.g-1FW converts to a global concentration of about 0.2 mM. Assuming low concentrations in the vacuole and apoplast, this would give a concentration of around 2 mM in other compartments, in reasonable agreement with reported values in spinach chloroplasts and the cytosol/cytoplasm of Arabidopsis cells (Foyer and Halliwell, 1976; Meyer et al., 2001). Calculations from immunolocalization studies performed in parallel with whole tissue quantification yielded estimated glutathione concentrations of 1 to 6 mM in the plastids, cytosol, and mitochondria of Arabidopsis Col-0 leaf mesophyll cells, with concentrations in the vacuole below 50 µM (Queval et al., 2011). Using non-aqueous fractionation of Arabidopsis leaves, cytosolic, plastid and vacuolar glutathione concentrations were estimated to be 3.5, 3.2 and 0.7 mM, respectively (Krueger et al., 2009). Thus, both these techniques give concentrations of the same order, regardless of potential problems such as cross-contamination between compartments or fractions during sample preparation. Differences in concentrations obtained in independent studies may also reflect differences in plant nutrition (eg, sulfur supply), which can affect the distribution of glutathione (Höller et al., 2010). Oxidative stress is among other factors that affect glutathione concentrations in a compartment-specific manner (Queval et al., 2011).

While the above two methods measure glutathione contents (and, potentially, concentrations) in vivo probes for the redox potential of glutathione have also been developed, based on fluorescent proteins containing oxidizable thiol groups. These include glutathione-specific redox-sensitive protein variants of the yellow (rxYFP) or green (roGFP) fluorescent protein (Hanson et al., 2004; Maulucci et al., 2009). These probes have been adapted for use in plants, particularly to monitor glutathione status in the cytosol (Jiang et al., 2006a; Meyer et al., 2007; Maughan et al., 2010). An important advantage of this method is that the introduced fluorescent proteins are designed to ‘sense’ the glutathione redox potential by equilibration of their thiol-disulfide status with that of glutathione (Figure 2B). Because the glutathione redox potential depends on GSH:GSSG and overall concentration (see section 8.1.2), these probes do not simply measure the glutathione pool size or even the oxidation state. However, as discussed further below, redox potential may be a key factor in glutathione-dependent signaling, and so these tools are providing useful information in studies of mutants or responses to stress (Meyer et al., 2007; Schwarzlander et al., 2008; Marty et al., 2009; Jubany-Mari et al., 2010). Potential constraints of these techniques are related to the limitations of optical microscopy. Most of the information generated to date has been obtained from studies of epidermal cells, and signal intensity may decrease for internal cell layers. However, such problems are less severe for the dual wavelength fluorescence ratio imaging roGFP variants (Schwarzlander et al., 2010). Because these proteins are dependent on glutaredoxin (GRX) activity, it is likely that glutathione redox potential is the major factor contributing to the signal (Meyer et al., 2007). However, recent studies have highlighted interactions between glutathione, GRX, and thioredoxins (TRX; Reichheld et al., 2007; Marty et al., 2009; Gao et al., 2010). It therefore remains to be definitively established that glutathione redox potential is the only component that is sensed by the thiol/disulfide-dependent fluorescent proteins.

3. MULTIPLE FUNCTIONS OF GLUTATHIONE

While this chapter is focused largely on glutathione homeostasis and related signaling functions, it is important to note that glutathione fulfils several essential functions in plants. Some of these roles can influence cellular glutathione status and thus the following sections briefly highlight some recent issues and key relevant points concerning the roles of glutathione in plant metabolism and defense.

3.1. Glutathione S-transferases

Compared to other phylogenetic groups, the GST family in plants is notable for its structural and functional diversity (see Dixon and Edwards (2010) for details of GST-encoding genes in Arabidopsis). While traditionally considered as cytosolic enzymes, work in Arabidopsis has revealed that several isoforms may localize at least partly to other compartments, including the chloroplast, peroxisome, and nucleus (Thatcher et al., 2007; Dixon et al., 2009). In many cases, the biochemical functions of specific GSTs remain to be fully elucidated. Several of these enzymes have peroxidase or DHAR activity. At least some GSTs are inducible by H2O2 (Levine et al., 1994; Willekens et al., 1997), and several of the Arabidopsis genes are induced by oxidative stress. For example, some GSTs of the tau family are strongly and rapidly up-regulated in catalasedeficient plants, and can therefore be considered useful markers for increased intracellular availability of H2O2 (Vanderauwera et al., 2005; Queval et al., 2007, 2009; Chaouch et al., 2010). Some GSTs that are induced by jasmonic acid (JA; Yan et al., 2007) or that have been shown to have activity against oxylipins (Mueller et al., 2008) are differentially regulated in response to H2O2 and/ or GR deficiency (Mhamdi et al., 2010a).

Conjugation of metabolites to GSH may be important in biosynthetic and catabolic pathways (Dixon and Edwards, 2010). Recently, evidence has been presented that the reduced sulfur group in glucosinolates is derived from GSH rather than cysteine. For example, the pad2 mutant, which is deficient in glutathione but not cysteine (Parisy et al., 2006), has decreased glucosinolate levels and impaired insect resistance (Schlaeppi et al., 2008). Moreover, LC-MS analysis of tobacco expressing Arabidopsis genes involved in the early steps of glucosinolate synthesis identified a GS-conjugate whose further metabolism was dependent on γ-glutamyl peptidase activity (Geu-Flores et al., 2009). Formation of a GS-conjugate is also part of the camalexin biosynthetic pathway (Böttcher et al., 2009).

3.2. Glyoxylase and Formaldehyde Metabolism

Glyoxalases catalyze the metabolism of potentially toxic oxo-aldehydes, the simplest of which is glyoxal, to hydroxyacids. The most common physiological substrate is thought to be methylgloxyal, which is notably formed from triose phosphate, either non-enzymically or as an intermediate in the triose-phosphate isomerase reaction (Maiti et al., 1997; Marasinghe et al., 2005). Metabolism of such oxo-aldehydes through the glyoxalase system involves conjugation to GSH, catalyzed by glyoxalase 1. A single gene (At1g11840) is annotated as a glyoxalase I homolog in Arabidopsis. Glyoxalase 2 then cleaves the glutathione to yield the corresponding free hydroxyacid (lactate in the case of methylglyoxal). Five genes are annotated to encode glyoxalase 2 in Arabidopsis (GLX2-1 to GLX2-5; At2g43430, At3g10850, At1g53580, At1g06130, At2g31350). Three of the encoded proteins are predicted to be mitochondrial enzymes that may have novel functions as well as or instead of glyoxalase 2 activity (Limphong et al., 2009). Tobacco plants overexpressing glyoxalases 1 and 2 or both glyoxalases together showed increased tolerance to both exogenous methylglyoxal and salt (Singla-Pareek et al., 2003). Similar strategies were reported to enhance salt tolerance in rice and Arabidopsis (Deb Roy et al., 2008; Singla-Pareek et al., 2008).

As well as the glyoxalase pathway, glutathione may also be involved in detoxification of formaldehyde via S-formylglutathione, producing formate which may be converted to CO2 or enter C1 metabolism (Martinez et al., 1996; Haslam et al., 2002; Achkor et al., 2003). GSHdependent formaldehyde dehydrogenase activity in Arabidopsis can result from the product of the At5g43940 gene. However, this gene encodes an enzyme with GSNO reductase activity (Sakamoto et al., 2002; Diaz et al., 2003), which may be its most important physiological function (see section 8.5).

3.3. Phytochelatins and Heavy Metal Resistance

One of the best characterized functions of glutathione is as a precursor in the synthesis of phytochelatins, which are required for responses to cadmium and other heavy metals (Grill et al., 1987, 1989; Cobbett and Goldsbrough, 2002; Rea et al., 2004). In plants and fungal species such as fission yeast, phytochelatins are produced from glutathione or homologs by phytochelatin synthase (PCS), a cytosolic enzyme (Blum et al., 2010). Two PCS genes are annotated in Arabidopsis (Ha et al., 1999; Cazalé and Clemens, 2001). The essential role of PCS1 (At5g44070) is evidenced by its identification as the gene affected by the cad1 mutation (Howden and Cobbett, 1992; Ha et al., 1999). PCS2 (At1g03980) also encodes a functional enzyme but is expressed in most tissues at much lower levels than PCS1 (Cazalé and Clemens, 2001; Blum et al., 2010). An alternative route of phytochelatin synthesis involving vacuolar carboxypeptidases has been reported in budding yeast (Wünschmann et al., 2007). Increased sensitivity to cadmium in the cad2 mutant is caused by impaired activity of γ-glutamylcysteine synthetase (γ-ECS), the first enzyme of the glutathione biosynthesis pathway (Howden et al., 1995; Cobbett et al., 1998). This results in a glutathione content of about 25% wild-type levels.

While transgenic up-regulation of glutathione synthesis has been reported to increase heavy metal tolerance, effects of overexpressing PCS have been less clear (Zhu et al., 1999a,b; Lee et al., 2003). Attempts have been made to improve cadmium sequestration and/or tolerance by engineering leaf-specific or ectopic expression of PCS in the chloroplast (Peterson and Oliver, 2006; Picault et al., 2006). Evidence is accumulating that PCS may be involved in cellular processes other than phytochelatin synthesis (Rea et al., 2004; Clemens and Peršoh, 2009), possibly including turnover of GS-conjugates (Blum et al., 2007, 2010). Ongoing work is identifying phytochelatin-independent mechanisms of heavy metal resistance (Hugouvieux et al., 2009). The role of glutathione in heavy metal responses could be multiple, as many of these elements are considered to directly or indirectly result in enhanced cellular oxidation. Interestingly, in Brassica napus exposed to cadmium, phytochelatins as well as GS-cadmium complexes were found at high concentrations in the phloem whereas phytochelatin concentrations were much lower in the xylem (Mendoza-Cózatl et al., 2008).

3.4. Glutathione and Sulfur Assimilation

Cysteine is the first organic product of sulfur assimilation in plants and is notably used for synthesis of protein amino acids, either as cysteine itself or after synthesis of methionine (Leustek et al., 2000; Saito, 2000; Droux, 2004). The production of these two primary S products requires interactions with both N and C metabolism, as does the production of the tripeptide, glutathione, as discussed previously by ourselves and others (Noctor et al., 1998a,b; Kopriva and Rennenberg, 2004). Glutathione is a significant non-protein sink for reduced sulfur. Sulfur supply has been positively correlated with resistance to some pathogens, a phenomenon termed sulfur-induced resistance (SIR; Bloem et al., 2007). Although the causes underlying SIR remain to be identified, tissue contents of glutathione or precursors may be among the factors linking sulfur nutrition to the responses of plants to fungal and viral infection (Gullner et al., 1999; Bloem et al., 2007; Zechmann et al., 2007; Höller et al., 2010).

3.4.1. Glutathione regulation of sulfur assimilation

As a major storage and transport form of reduced sulfur and non-protein cysteine, glutathione has been implicated in the regulation of sulfur metabolism in Arabidopsis, as in other plants (Kopriva and Rennenberg, 2004). Several components involved in sulfur assimilation are regulated by glutathione. Inhibition of sulfate uptake by GSH occurs in tobacco (Herschbach and Rennenberg, 1994). Similarly, glutathione has been reported to repress sulfur assimilation and uptake activities of sulfate transporters in a number of species (Buchner et al., 2004). In Arabidopsis, GSH inhibits the expression of AST68, a sulfate transporter, and ATP sulfurylase 1 (APS1), accompanied by a decrease of sulfate influx and APS activity in the roots (Lappartient et al., 1999). All three isoforms of the second enzyme of the sulfate assimilation pathway, adenosine 5′-phosphosulfate reductase (APR), seem to be particularly sensitive to inhibition by glutathione at the levels of transcripts and activity in Arabidopsis roots (Vauclare et al., 2002).

Available evidence suggests that GSH is the most important reductant for APR (Leustek, 2002). This enzyme contains a carboxyl domain with homology to TRX that appears to act as a functional GRX, as it uses GSH to reduce adenosine 5′-phosphosulfate to sulfite with apparent KMGSH values of 0.6 to 1 mM (Bick et al., 1998). Post-translational redox control of APR (particularly APR1) marks it out as a significant control point in response to glutathione oxidation (Leustek, 2002).

3.4.2. Manipulation of sulfur assimilation and effects on glutathione

Sulfur availability limits GSH accumulation in plants (Nikiforova et al., 2003). An Arabidopsis mutant defective in the Sultr1;2 gene, encoding a high affinity sulphate transporter, has decreased levels of glutathione (Maruyama-Nakashita et al., 2003) while constitutive expression of a bacterial APR in Arabidopsis increases the contents of both cysteine and glutathione (Tsakraklides et al., 2002). Overexpression of the cysteine synthesis pathway enhances cysteine and glutathione contents in Arabidopsis and tobacco (Harms et al, 2000; Noji and Saito, 2002; Wirtz and Hell, 2007). Environmental stresses such as ozone concomitantly increase cysteine and glutathione levels, associated with the posttranslational activation of APR1 (Bick et al., 2001). Genes encoding all three APRs are also induced by intracellular H2O2, as are cysteine, γ-EC, and glutathione (Queval et al., 2009).

4. GLUTATHIONE BIOSYNTHESIS

4.1. Enzymes, Mutants, and Inhibitors

The committed pathway of glutathione synthesis in plants has been well established for many years (Rennenberg, 1982; Noctor et al., 2002; Mullineaux and Rausch, 2005). The metabolic route is essentially similar to the pathway in most other organisms containing GSH (Meister, 1988b). From glutamate, cysteine, and glycine, two ATP-dependent enzymes produce GSH via the intermediate, γ-EC (Figure 3). Plant sequences for the single genes that encode each of these enzymes were first reported in Arabidopsis (May and Leaver, 1994; Ullman et al., 1996). The first enzyme, γ-ECS, is encoded by the GSH1 gene (At4g23100). In wheat leaves, γ-ECS activities are strongly associated with chloroplasts (Noctor et al., 2002) and subcellular localization studies of the Arabidopsis protein have convincingly demonstrated that the enzyme is restricted to plastids (Wachter et al., 2005). In contrast, the GSH2 gene (At5g27380) encodes two transcripts. The most abundant is the shorter form, which encodes a cytosolic GSH-S, while the longer form encodes a chloroplast-targeted protein (Wachter et al., 2005). Thus, the first step of glutathione synthesis is located in the plastid while the second step can occur in both chloroplasts and cytosol. Plastidic γ-ECS and cytosolic GSH-S are linked by γ-EC export across the chloroplast envelope and transporters have recently been described that are competent in both γ-EC and GSH transport (Maughan et al., 2010).

4.1.1. γ-ECS mutants

Knockout mutations in GSH1 produce an embryo-lethal phenotype (Cairns et al., 2006). Less severe mutations in this gene, which produce partial decreases in glutathione contents, have been extremely useful in elucidating the functions of glutathione in plants. The rml1 (rootmeristemless1) mutant has less than 5% of wild-type glutathione contents. It fails to develop a root apical meristem because the cells arrest at the G1 phase of the cell cycle (Vernoux et al., 2000). This mutant is much less affected in its shoot phenotype. By combining the rml1 mutation with mutations in the two genes encoding dual-addressed cytosolic/mitochondrial NADPH-thioredoxin reductases (NTRA, At2g17420; NTRB At4g35460), it was shown that there is functional redundancy between glutathione and TRX systems in the control of shoot apical meristem function (Reichheld et al., 2007). Moreover, the combination of the cad2 mutation with ntra and ntrb resulted in perturbed auxin transport and metabolism, together with a loss of apical dominance and reduced secondary root production (Bashandy et al., 2010).

Several mutations in GSH1 decrease glutathione to a residual but still significant level (about 25 to 50% of wild-type). In themselves, these decreases do not markedly affect Arabidopsis development. They do, however, cause alterations in responses to stress and the environment. Of these mutants, cad2 was identified by its enhanced sensitivity to cadmium, rax1 by altered expression of APX2 (At3g09640) and pad2 by decreased camalexin contents and enhanced sensitivity to pathogens (Howden et al., 1995; Cobbett et al., 1998; Ball et al., 2004; Parisy et al., 2006).

4.1.2. Glutathione synthetase

As stated above, GSH-S is encoded by a single gene, GSH2, but alternative splicing results in localization of the gene product in both the chloroplast and cytosol (Wachter et al., 2005). While GSH1 knockouts are embryo-lethal, GSH2 knockout mutants are seedling-lethal, probably reflecting partial replacement of GSH functions by γ-EC, which accumulates to very high levels in these plants (Pasternak et al., 2008). The wild-type phenotype can be restored in gsh2 by complementing the mutant with targeted expression of the enzyme to the cytosol alone (Pasternak et al., 2008). This provides further evidence that GSH can be imported from the cytosol into the plastid, consistent with radiolabelling studies of isolated wheat chloroplasts (Noctor et al., 2002).

Figure 3.

Pathway of glutathione biosynthesis and principal methods of regulation.

(A) Scheme showing production of glutathione from constituent amino acids and cysteine synthesis enzymes that are likely important in regulating glutathione concentration. Important regulatory mechanisms are shown as red lines (feedback inhibition) or green lines (up-regulation in response to oxidation or other factors). Green dotted lines indicate likely transcriptional activation mechanisms while solid lines indicate post-translational processes. The role of GSSG and the exact mechanisms involved in the activation of γ-ECS by thiol oxidation remain to be established. For simplicity, not all enzymes involved in sulfate reduction are shown. APR, adenosine phosphosulfate reductase. γ-ECS, γ-glutamylcysteine synthetase. GRX, glutaredoxin. GSH-S, glutathione synthetase. JA, jasmonic acid. OAS, O-acetylserine thiol lyase. SAT, serine acetyl transferase.

(B) Expression of GSH1 and GSH2 in response to a range of environmental stresses ( www.genevestigator.com; Hruz et al., 2008). Examples of significant responses are highlighted. Botrytis, rosettes infected with B.cinerea or B.graminis versus uninfected rosettes. Drought, leaves from draughted plants versus watered plants. Light, light-grown seedlings compared to dark-grown seedlings. MeJa, methyl jasmonate-treated cell cultures versus solvent-treated cell cultures.

4.1.3. Buthionine sulfoximine

As well as mutants, one tool that has proved very useful in functional analysis of glutathione in plants is buthionine sulfoximine (BSO), a specific inhibitor that binds to the γ-ECS active site (Griffith and Meister, 1979). This compound is structurally homologous to methionine sulfoximine (MSO), which inhibits both γ-ECS and glutamine synthetase. The reactions catalyzed by both enzymes involve ATP-dependent ligation of an amine group (ammonia or the amino group of cysteine) to the γ-carboxy group of glutamate. Unlike MSO, BSO does not inhibit glutamine synthetase and so can be used in studies specifically focused on the role of glutathione synthesis and concentration in plants (eg, Cobbett et al., 1998; Kocsy et al., 2000; Meyer et al., 2007; Reichheld et al., 2007). Recently, a screen for BSO-insensitive Arabidopsis mutants revealed the identity of the chloroplast γ-EC and GSH transporter, called chloroquinone-like transporter (CLT; Maughan et al., 2010). These transporters are encoded by three genes: CLT1 (At5g19380), CLT2 (At4g24460), and CLT3 (At5g12160/12170). This finding, taken with the available information on γ-ECS and GSH-S localization, leads to the view that the synthesis of γ-EC is restricted to the plastid but that GSH synthesis can also occur in the cytosol following γ-EC export across the plastid envelope (Wachter et al., 2005; Maughan et al., 2010).

4.2. Regulation and Manipulation of Glutathione Synthesis

The production of glutathione is regulated at multiple levels, with the most important limitations being γ-ECS activity and cysteine concentration. This has been shown by overexpression of the two enzymes of glutathione synthesis and of enzymes involved in cysteine synthesis (Strohm et al., 1995; Noctor et al., 1996, 1998a; Creissen et al., 1999; Harms et al, 2000; Noji and Saito, 2002; Wirtz and Hell, 2007). Although these factors are considered to be the major controls, regulation is possible at other levels, such as the availability of glycine and ATP (Buwalda et al., 1990; Noctor et al., 1997; Ogawa et al., 2004). Even though the overall flux through the glutathione synthesis pathway is relatively low in comparison to primary metabolism, the multiplicity of controls suggests that the pathway is highly regulated and exquisitely sensitive to the cellular redox environment. This complexity can also be understood in terms of the different physiological functions of modified glutathione synthesis. In certain cases, such as heavy metal exposure, GSH is consumed and accelerated synthesis may be accompanied by unchanged or decreased glutathione pools. In other stresses with a strong oxidative component, increased synthesis occurs within a context of accumulating glutathione, a response that has presumably evolved to offset changes in redox potential caused by increased GSSG (see section 8.1.2).

4.2.1. Expression of glutathione synthesis genes

Relatively few conditions cause very marked induction of GSH1 or GSH2 transcripts. These genes are induced by JA and heavy metals (Xiang and Oliver, 1998; Sung et al., 2009) and also respond to light and some stress conditions such as drought and certain pathogens (Figure 3B). Increases in H2O2 availability are known to trigger accumulation of glutathione in several plant species (Smith et al., 1984; May and Leaver, 1993; Willekens et al., 1997; Queval et al., 2009). However, neither exogenous H2O2 nor intracellularly generated H2O2 cause increased abundance of GSH1 or GSH2 transcripts in Arabidopsis (Xiang and Oliver, 1998; Sánchez-Fernández et al., 1998; Queval et al., 2009). Some evidence has been presented that production of the γ-ECS protein is regulated at the level of translation (Xiang and Bertrand, 2000).

4.2.2. Post-translational regulation of γ-ECS

Feedback inhibition of enzyme activity by GSH has long been established in animals (Richman and Meister, 1975) and also occurs in plants (Hell and Bergmann, 1990; Noctor et al., 2002). Other types of regulation were proposed based on the failure of Arabidopsis γ-ECS to fully restore wild-type GSH levels in yeast mutants (May et al., 1998b). The tobacco enzyme was shown to be inhibited by dithiols (Hell and Bergmann, 1990), and this effect has also been described for the wheat and Arabidopsis enzymes (Noctor et al., 2002; Jez et al., 2004). The plant enzyme forms a homodimer linked by two disulfide bonds (Hothorn et al., 2006), one of which is involved in redox regulation (Hicks et al., 2007; Gromes et al., 2008). Though the exact mechanistic details are still subject to debate, activation through disulfide formation very probably contributes to the well-known up-regulation of glutathione synthesis in response to oxidative stress conditions. It remains unclear whether this mode of regulation interacts mechanistically with feedback inhibition.

4.2.3. Cysteine availability

Up-regulation of the cysteine synthesis pathway occurs concomitantly with that of GSH synthesis. Two levels of co-regulation have been described under conditions in which oxidative stress triggers accumulation of glutathione in Arabidopsis. In catalase-deficient mutants, accumulation of transcripts encoding APR and serine acetyltransferase (SAT) was observed (Queval et al., 2009). Like γ-ECS, APRs are localized in plastids, whereas the two enzymes of cysteine synthesis are found in several compartments (Droux, 2004; Kopriva and Rennenberg, 2004; Martin et al., 2005; Kawashima et al., 2005). Recent studies show that the mitochondrial SAT (encoded by SERAT2.2; At3g13110) makes the major contribution to cysteine synthesis under standard conditions, and that knocking out this enzyme affects flux into glutathione and leaf glutathione contents (Haas et al., 2008; Watanabe et al., 2008). However, transcripts for the chloroplast SAT (SERAT2.1; At1g55920) were the most strongly induced during H2O2-triggered up-regulation of glutathione in Arabidopsis (Queval et al., 2009).

Exposure of Arabidopsis to ozone activates at least one APR at the post-translational level through formation of a disulfide bond (Bick et al., 2001). Although much more is known about thiol-disulfide regulation mediated by TRX (Buchanan and Ballmer, 2005), GRX are also important players in redox regulation (see section 8.3). One simple explanation of the link between decreased GSH:GSSG and enhanced glutathione synthesis is that increases in the chloroplast glutathione redox potential allow GRX-mediated oxidative activation of both APR and γ-ECS (Figure 3A). This would be consistent with a model for APR regulation in which the enzyme is maintained in a reduced inactive form by TRX and activated by GSSG (Leustek, 2002). Accordingly, H2O2-triggered accumulation of glutathione in barley is associated with appreciable increases in chloroplast GSSG content (Smith et al., 1985), and in situ analysis of glutathione in leaf mesophyll cells strongly points to a similar effect in Arabidopsis (Queval et al., 2011).

4.2.4. Overexpression of the glutathione synthesis pathway

Despite the multiplicity of controls over glutathione synthesis, plant cells can accommodate marked increases in overall glutathione content. Studies using expression of E.coli γ-ECS targeted to the poplar chloroplast or cytosol or to the tobacco chloroplast produced increases in leaf glutathione of up to four-fold (Noctor et al., 1996, 1998a; Creissen et al., 1999). Even higher contents were achieved by combined overexpression of γ-ECS and GSH-S (Creissen et al., 1999). Overexpression of γ-ECS in Arabidopsis also enhanced GSH, though less markedly (Xiang et al., 2001). Initial studies of several independent lines of poplar overexpressing γ-ECS showed no phenotypic effects (Noctor et al., 1998a). One of these lines has been reported to show symptoms of early leaf senescence (Herschbach et al., 2009). Clear effects of γ-ECS overexpression in tobacco included lesion formation and high levels of GSSG accumulation (Creissen et al., 1999), whereas marked negative phenotypic effects were absent in poplar and Brassica juncea expressing the same gene (Noctor et al., 1998a; Zhu et al., 1999a). Recently, expression of a bifunctional feedback-insensitive γ-ECS/GSH-S enzyme from the bacterium Streptococcus thermophilus in tobacco has shown that the glutathione pool can be boosted by up to 20-fold without apparent marked effects on plant growth and development (Liedschulte et al., 2010). In view of the information from several recent studies (see section 8.6), these plants present a very interesting system in which to analyze responses to biotic and other stresses. Increased glutathione in plants overexpressing γ-ECS is associated with enhanced resistance to heavy metals and certain herbicides (Zhu et al., 1999a; Gullner et al., 2001). Overexpression of γ-ECS in the poplar chloroplast increased not only glutathione but also levels of several free amino acids, such as leucine, isoleucine, leucine, tyrosine, and lysine (Noctor et al., 1998a). The causes underlying these effects remain unclear, though it is interesting to note that several enzymes involved in synthesis or metabolism of some of these amino acids are among those that have been identified as potential TRX targets through redox protemics approaches (Montrichard et al., 2009).

Compared to boosting γ-ECS capacity, overexpression of GSH-S has less marked effects on glutathione contents. No effect of substantial increases in GSH-S activity was observed in poplar (Foyer et al., 1995; Noctor et al., 1998a), though enhanced synthesis of GSH was induced when γ-EC was supplied to leaf tissue from the transformants (Strohm et al., 1995). Thus, GSH-S may become limiting when γ-EC supply is increased. Brassica juncea lines overexpressing GSH-S showed improved resistance to cadmium (Zhu et al., 1999b). Expression of a soybean homoGSH-S was used as part of an interesting strategy to confer herbicide resistance in tobacco, which is sensitive to the herbicide, fomesafen (Skipsey et al., 2005b). When expressed together with a homoGSH-preferring GST from soybean, which is fomesafen-tolerant, enhanced resistance was observed in the transformed tobacco lines (Skipsey et al., 2005b).

5. GLUTATHIONE BREAKDOWN

Overall rates of glutathione catabolism in Arabidopsis leaves have been estimated at about 30 nmol.g-1 FW h-1 (Ohkamu-Ohtsu et al., 2008), suggesting that some pools of glutathione could turn over several times daily. In mammals, degradation of glutathione to dipeptides and amino acids is part of the γ-glutamyl cycle (Meister, 1988b). This involves hydrolysis or transpeptidation of GSH at the plasmembrane by γ-glutamyl transpeptidase (GGT), and intracellular metabolism of γ-glutamyl amino acid derivatives by γglutamyl cyclotransferases (GGC) and 5-oxoprolinase (5-OPase) to produce free glutamate. These enzyme activities were characterized in tobacco cells by Rennenberg and co-workers (Rennenberg et al., 1981; Steinkamp and Rennenberg, 1984; Steinkamp et al., 1987). Two GGTs purified from tomato were shown to have broad substrate specificity against GSH, GSSG and GS-conjugates and to localize to the cell wall or outer surface of the plasmalemma (Martin and Slovin, 2000). Both showed KM GSH values of about 0.1 mM (Martin and Slovin, 2000), consistent with their localization in a compartment where GSH concentrations are relatively low (Vanacker et al., 1998; Ohkamu-Ohtsu et al., 2007a). Studies in maize and barley also detected GGT activity in the apoplast (Masi et al., 2007; Ferretti et al., 2009). Interestingly, the pH optimum of the enzyme detected in barley roots was reported to be higher than seven, which suggests that the activity could increase in conditions that induce alkalinisation of the apoplast (Ferretti et al., 2009). Recent studies in Arabidopsis have significantly extended our knowledge in this area while also revealing the potential complexity of the breakdown of GSH, GSSG, and GS-conjugates (Figure 4). Key outstanding questions are cellular/ tissue specificities and activities against GSH compared to GS-conjugates or GSSG, which can be considered as a glutathioneglutathione conjugate.

Figure 4.

Possible routes of glutathione breakdown in Arabidopsis.

For clarity, the term γ-EC, used elsewhere in this chapter, is replaced here by γ-Glu-Cys. Emphasis is placed on the localization of the initiating reactions. For simplicity possible downstream reactions are not always shown (eg, GGT action on γ-Glu-Cys conjugates). Only the transpeptidation reaction catalyzed by GGT is shown, but these enzymes could also catalyze hydrolysis of GSH to Glu and Cys-Gly. aa, amino acid(s). CP, carboxypeptidase. GGC, γ-glutamyl cyclotransferase. GGT, γ-glutamyl transpeptidase. 5-OP, 5-oxoproline. 5-OPase, 5-oxoprolinase. PCS, phytochelatin synthase. X, S-conjugated compound or second moiety of disulfide.

5.1. γ-Glutamyl Transpeptidase

Four genes were predicted to encode GGT in Arabidosis (Storozhenko et al., 2002). One of these sequences (GGT1; At4g39640) was identified in a screen of Arabidopsis cDNAs that render yeast tolerant to the thiol-oxidizing agent, diamide (Kushnir et al., 1995). Overexpression of this GGT in tobacco caused an increase in extracellular GGT activity (Storozhenko et al., 2002). The GGT2 sequence (At4g39650) is contiguous and highly similar to GGT1 and is also localized outside the cell, though its expression is largely confined to young siliques whereas GGT1 is more widely expressed (Ohkamu-Ohtsu et al., 2007a; Destro et al., 2010). However, root GGT activity can be enhanced by incubation with GSH and comparative analysis of ggt1 and wild-type plants suggests that GGT2 may contribute to this inducible activity (Destro et al., 2010). Unlike animal GGT, GGT1 and GGT2 are probably bound to the cell wall rather than anchored to the plasmalemmma (Martin et al., 2007; Ohkamu-Ohtsu et al., 2007a). Knockout mutants for GGT1 show a phenotype of decreased growth and early senescence, linked to accumulation of GSSG (Ohkamu-Ohtsu et al., 2007a). Thus, one function of extracellular GGTs may be to counteract oxidative stress or to salvage excreted GSSG (Ohkamu-Ohtsu et al., 2007a; Ferretti et al., 2009). A third GGT, called GGT4 (At4g29210), originally called GGT3 by Ohkamu-Ohtsu et al. (2007a,b), is localized in the vacuole where its major role is considered to be the metabolism of GS-conjugates (Grzam et al., 2007; Martin et al., 2007; Ohkamu-Ohtsu et al., 2007b). The fourth sequence (At1g69820), now called GGT3, is considered unlikely to encode a functional GGT but transcripts have been detected in some tissues (Martin et al., 2007; Destro et al., 2010).

5.2. Glutathione Breakdown by Other Enzymes

While several studies have focused on GGTs, a recent analysis suggests that these enzymes may not be the major route for GSH breakdown in plants (Ohkamu-Ohtsu et al., 2008). γ-Glutamyl peptides such as those that result from GGT activity, can be degraded to 5-oxoproline by GGC activity. No gene has yet been identified for GGC in Arabidopsis but a single gene (OXP1; At5g37830) shows high similarity to animal 5-OPase and the encoded protein is likely found in the cytosol (Ohkamu-Ohtsu et al., 2008). Knockout oxp1 mutants accumulate high levels of 5-oxoproline, and this accumulation is not affected in triple oxp1 ggt1 ggt4 mutants that are deficient in the major GGT activities as well as 5-OPase (Ohkamu-Ohtsu et al., 2008). Based on these and other observations, it was proposed that cytosolic GGC, rather than vacuolar or extracellular GGT, is the major enzyme initiating GSH degradation and that this route could be particularly important in the intercellular transport function of GSH (Ohkamu-Ohtsu et al., 2008).

Another cytosolic route for glutathione degradation could involve PCS, which may be particularly important in GS-conjugate breakdown (Blum et al., 2007, 2010). In Arabidopsis leaf discs, no evidence was found that PCS contributes to metabolism of GS-conjugates, though this route might become significant when PCS is activated by heavy metals (Grzam et al., 2006). Work in Arabidopsis suspension cells suggests that cytosolic PCS is a potentially significant route for GS-conjugate metabolism, and the relative importance of the different pathways could be dependent on tissue or cell type (Blum et al., 2007; Brazier-Hicks et al., 2008). Yet another route of glutathione degradation would involve cleavage of the glycine by a carboxypeptidase activity (Steinkamp and Rennenberg, 1985). A vacuolar carboxypeptidase activity has been described that degrades GS-conjugates in barley (Wolf et al., 1996). Although not shown in Figure 4, other proteins may also metabolize GSH, GSSG, or GS-conjugates. These include proteins such as GGP1, an enzyme with a putative glutamine amidotransferase domain that has been implicated in the removal of the Glu residue from a GS-conjugate during glucosinolate synthesis (Geu-Flores et al., 2009). In addition to GGC, Arabidopsis genes remain to be identified for the vacuolar carboxypeptidase and for dipeptidases that could cleave Cys-Gly. In conclusion, there is still no consensus on the compartmentation of glutathione degradation. Most attention continues to be paid to apoplastic and vacuolar routes. A third possibility involves cytosolic degradation of GS-conjugates although evidence for appreciable accumulation of these compounds in this compartment is lacking.

6. COMPARTMENTATION AND TRANSPORT

6.1. Intercellular Transport

Glutathione synthesized in the chloroplast and cytosol is transported between cells and organs. Together with S-methylmethionine, glutathione is considered to be a major form of organic sulfur translocated in the phloem (Herschbach and Rennenberg, 1994, 1995; Bourgis et al., 1999; Mendoza-Cózatl et al., 2008). Tobacco cells and broad bean protoplasts are able to import both GSH and GSSG (Schneider et al., 1992; Jamaï et al., 1996). One group of proteins that could be responsible for transport across the plasmalemma is the oligopeptide transporter (OPT) family. First identified in yeast, where the protein is named HGT1, these proteins are able to transport GSH, GSSG, GS-conjugates as well as other small peptides (Bourbouloux et al., 2000). Heterologous expression in yeast of a homologous B. juncea gene, BjGT1, showed that the protein was competent in glutathione uptake (Bogs et al., 2003) while a rice homolog, OsGT1, can transport GSH, GSSG, and GS-conjugates (Zhang et al., 2004). In Arabidopsis, there are nine annotated OPT genes (Koh et al., 2002). OPT6 (Atg27730) may be able to transport GS-conjugates and GS-cadmium complexes as well as GSH and GSSG (Cagnac et al., 2004). Another study reported that GSH but not GSSG was transported by OPT6 (Pike et al., 2009). Consistent with its proposed role in long-distance transport, the OPT6 gene is highly expressed in the vasculature (Cagnac et al., 2004). However, OPT6 does not seem to be a glutathione-specific transporter: it is active against a broad spectrum of peptides, including peptides that may have signaling functions in plants (Pike et al., 2009). Knockout opt6 mutants are aphenotypic, suggesting gene redundancy (Pike et al., 2009).

6.2. Transport Between the Plastid and the Cytosol

Direct analysis of 35S-GSH uptake by wheat chloroplasts suggested the existence of components able to translocate glutathione across the inner chloroplast envelope (Noctor et al., 2002). The recent identification of the CLT family of transporter proteins that are targeted to the plastid envelope and that transport both γ-EC and glutathione (Maughan et al., 2010) is consistent with the exclusive location of γ-ECS in the plastid and the major location of GSH-S in the cytosol (Figure 5; Wachter et al., 2005). The triple clt1 clt2 clt3 knockout has depleted cytosolic GSH and enriched chloroplastic GSH (Maughan et al., 2010). The first of these transporters was identified in a screen for mutants resistant to BSO. There is as yet no direct evidence that the CLT transports the inhibitor itself. However, its closest homolog, the chloroquinone transporter in the malaria parasite Plasmodium, has recently been shown to bind BSO (Martin et al., 2009). Chloroplast uptake of GSH by this transporter family likely explains the observation that gsh2 knockout mutants can be rescued by expression of GSH-S in the cytosol alone (Pasternak et al., 2008).

6.3. Transport into the Nucleus and Mitochondria

A significant proportion of cellular GSH has been detected in the nucleus using in situ immunolocalization techniques (Zechmann et al., 2008). Indeed, glutathione is recruited into the nucleus early in the cell cycle in both mammals and plants (Markovic et al., 2007; Diaz-Vivancos et al., 2010a). This has led to the concept of a nuclear redox cycle that occurs during mitosis in which GSH moves into the nucleus during G1, promoting oxidation of the cytosol and enhanced glutathione accumulation (Diaz-Vivancos et al., 2010b). In Arabidopsis cells growing in culture, GSH localization in the nucleus at G1 is accompanied by enhanced levels of ROS and lowering of the oxidative defense shield, as reflected by transcriptomic changes (Diaz-Vivancos et al., 2010a). The accumulated GSH is divided between the daughter cells, in which the process begins again. These observations suggest the presence of proteins in plants that are able to alter the permeability of nuclear pores, facilitating GSH sequestration in the nucleus. While in plants these proteins remain to be identified, in mammalian tissues it is considered that the anti-apoptotic factor Bcl-2 is a crucial component regulating GSH transport into the nucleus, as it is in mitochondria (Voehringer et al., 1998). As noted above, high concentrations of glutathione have been detected in leaf mitochondria (Zechmann et al., 2008), despite the absence of evidence for GSH synthesis in this organelle. It is thus tempting to suggest that pore-regulating proteins operate to control the mitochondrial and nuclear GSH concentration in a manner similar to that observed in animal cells (Figure 5).

Figure 5.

Subcellular compartmentation of glutathione synthesis and glutathione transport.

BAG, Bcl2-associated anathogene. CLT, chloroquinone-like transporter. MRP multidrug resistance-associated protein. OPT, oligopeptide transporter.

6.4. Transport into the Vacuole

GS-conjugates are known to be transported into the vacuole by tonoplast MRP (Multidrug Resistance-associated Protein) transporters of the ABC (ATP-Binding Cassette) type (Martinoia et al., 1993; Rea, 1999; Foyer et al., 2001). Barley vacuoles rapidly take up GSSG but GSH is taken up much more slowly (Tommasini et al., 1993). Arabidopsis MRPs can transport GSSG as well as GS-conjugates when heterologously expressed in yeast (Lu et al., 1998). Vacuolar import of GSSG may play a role in maintaining cytosolic glutathione redox status (Tommasini et al., 1993). Recent observations in Arabidopsis suggest that vacuolar accumulation of GSSG is a physiologically important part of the response to oxidative stress (Queval et al., 2011). Glutathione was increased about four-fold in catalase-deficient cat2 mutants (mainly as GSSG), but most of this increase was accounted for by changes in chloroplast and vacuolar glutathione, with the cytosolic and nuclear glutathione pools remaining relatively stable (Queval et al., 2011). The KMGSSG for uptake into barley vacuoles was found to be 0.4 to 0.6 mM (Tommasini et al., 1993) while values for AtMRPI (At1g30400) and AtMRP2 (At2g34660) were reported to be 219 and 73 µM, respectively (Lu et al., 1998). In the absence of stress conditions, cytosolic GSSG concentrations may be well below these values (see section 8.1). This suggests that the activity of MRP proteins will be stimulated in response to stress-induced increases in cytosolic GSSG accumulation and function to dampen such increases. Further, specific MRPs are induced by oxidative stress (Sánchez-Fernández et al., 1998) and also in Arabidopsis mutants that accumulate high amounts of GSSG (Mhamdi et al., 2010a).

7. REDOX TURNOVER

The global tissue glutathione pool is largely reduced under optimal conditions. Divergence from this highly reduced state is an indicator of stress. However, it should be noted that certain compartments, such as the endoplasmic reticulum, are maintained in an oxidized state (Hwang et al., 1992; Enyedi et al., 2010), as are also some cell types (eg, those of the root quiescent center) or dormant tissues like seeds (Kranner and Grill, 1996; Kranner et al., 2002; Kranner et al., 2006). Decreases in GSH:GSSG may be produced by compromised regeneration of GSH, direct reaction with ROS and other oxidants, or the activities of several types of peroxidases. The principal enzymes involved in these reactions are now discussed.

7.1. Peroxidation Reactions

There is a close relationship between intracellular H2O2 availability and the status of the glutathione pool. This is most notably evident from the effect of inhibiting leaf catalase activities, either pharmacologically or genetically. While such inhibition causes modest or undetectable increases in H2O2 contents, it triggers a marked perturbation of both the glutathione pool size and redox state (Smith et al., 1984; May and Leaver, 1993; Willekens et al., 1997; Noctor et al., 2002; Riszhky et al., 2002; Queval et al., 2007, 2009; Chaouch et al., 2010). The effects of increased intracellular H2O2 availability appear to be quite specific to glutathione, and do not involve readily detectable changes in the status of other major antioxidants or redox couples (Mhamdi et al., 2010a).

Apart from catalase, various peroxidases can metabolize H2O2. Since the 1970's the predominant view has been that the major antioxidative H2O2-metabolizing peroxidases in plants are APXs, rather than GSH-dependent peroxidases, which are important in mammalian cells. However, GPXs have also been characterized in plants (Eshdat et al., 1997). Unlike the animal enzymes, plants GPXs are selenium-independent, because their catalytic action depends on cysteine rather than selenocysteine (Eshdat et al., 1997). Plant GPXs can use both GSH and TRX as reductants (Herbette et al., 2002) but TRX is the more efficient reducing substrate (Iqbal et al., 2006). Thus, the enzymes can functionally be considered to be peroxiredoxins (PRX) rather than GPXs (Navrot et al., 2006). Several of these genes, such as GPX6 (At4g11600), are induced by oxidative stress (eg, Mhamdi et al., 2010a). In addition to possible antioxidant functions, GPXs may be involved in redox signal transduction in plants (Miao et al., 2006), as they are in yeast (Delauney et al., 2002). They have also been described as playing a role in plant immune responses as well as photooxidative stress (Chang et al., 2009). Unlike APX, these enzymes can reduce both H2O2 and organic peroxides with similar efficiency (Iqbal et al., 2006; Navrot et al., 2006). Compared to APX (Asada, 1999), the catalytic rate and affinity for H2O2 of GPXs are rather low. This lends support to the notion that the main physiological roles of these enzymes is in regulating the concentrations of lipid peroxides that are poorly or not metabolized by catalase or APX, or in peroxide sensing.

Athough annotated plant GPXs use TRXs more efficiently than GSH, at least two other types of enzyme could be important in peroxidation of glutathione (Figure 6). Certain GSTs show GSH-dependent peroxidase activity against H2O2 and organic peroxides (Wagner et al., 2002; Dixon et al., 2009). Work in poplar has shown that glutathione could also be involved in peroxide metabolism though GRX-dependent PRX (Rouhier et al., 2002). It has been reported in yeast that GRXs can also directly reduce peroxides to the corresponding alcohol (Collinson et al., 2002).

7.2. Ascorbate Regeneration

The classic route by which GSH oxidation is coupled to H2O2 metabolism is the ascorbate-glutathione pathway (Foyer and Halliwell, 1976). In this pathway, GSH contributes to ascorbate regeneration by reducing DHA, either chemically or via enzymes called DHA reductases, which constitute a class of GSTs (Dixon et al., 2002). Overexpression studies have confirmed the role of DHAR in maintaining a high tissue ascorbate redox state, and have highlighted the roles of GSH-dependent ascorbate pools in a number of physiological responses, eg, the control of stomatal closure (Chen et al., 2003).

It remains unclear whether direct GSH peroxidation or oxidation by DHA is the major route for H2O2-dependent glutathione oxidation (Figure 6). Some relevant physiological evidence comes from analysis of catalase-deficient plants which, as noted above, are useful systems for probing the potential functions of reductive H2O2-metabolizing pathways in vivo (Dat et al., 2001; Mhamdi et al., 2010c). The major leaf catalase isoform in Arabidopsis is CAT2 (At4g35090; Queval et al., 2007; Mhamdi et al., 2010c). In Arabidopsis cat2 knockouts, DHAR2 (At1g75270), which is predicted to encode a cytosolic enzyme, is induced along with cytosolic APX1 (At1g07890; Mhamdi et al., 2010a). This observation is consistent with studies of Arabidopsis apx1 knockouts (Davletova et al., 2005), and suggests that the cytosolic ascorbate-glutathione pathway is important in metabolizing H2O2 originating in other compartments. However, tobacco plants deficient in both catalase and cytosolic APX show a less marked phenotype than the parent lines (Riszhsky et al., 2002), whereas Arabidopsis double cat2 gr1 mutants have a more severe phenotype than cat2 single mutants (Mhamdi et al., 2010a). While a more direct comparison of interactions between these components is required, the above results seem to be at odds with the notion that GR activity is solely required for the classical ascorbate-glutathione pathway. If this were the case, APX and GR1 deficiency should not produce what appear to be opposite effects in a catalase-deficient background. This and other observations of Mhamdi et al. (2010a) provide in vivo evidence that the GR-glutathione system could be important in ascorbate-independent H2O2 metabolism (eg, through some GSTs) as well as for the ascorbate-glutathione pathway. Further, such observations are consistent with specificity of redox functions for ascorbate and glutathione, as discussed elsewhere (Noctor et al., 2000; Potters et al., 2002; Foyer and Noctor, 2011).

Figure 6.

Major links between glutathione and peroxide metabolism.

The scheme is intentionally simplified. Oxidation of glutathione could also be caused, directly or via ascorbate, through direct chemical reactions with compounds such as superoxide. APX, ascorbate peroxidase. ASC, ascorbate. DHA(R), dehydroascorbate (reductase). GR, glutathione reductase. GRX, glutaredoxin. GST, glutathione S-transferase. PRX, peroxiredoxin. ROOH, organic peroxide.

7.3. Glutathione Reductase

7.3.1. Genes and Subcellular Localization

The high cellular GSH:GSSG ratio is maintained by GRs, which are flavoproteins with high affinity for both GSSG and NADPH (Halliwell and Foyer, 1978; Edwards et al., 1990). Classical fractionation studies described high GR activities in chloroplasts (Foyer and Halliwell, 1976) but GR has also been detected in the cytosol, mitochondria, and peroxisomes (Edwards et al., 1990; Rasmusson and Møller, 1990; Jiménez et al., 1997; Stevens et al., 2000; Romero--Puertas et al., 2006). Two genes encode GR in Arabidopsis. The first, GR2 (At3g54660) encodes an enzyme corresponding to the pea GR that was the first protein shown to be dual-addressed to plastids and mitochondria (Creissen et al., 1995). Dual localization of the gene product in these two compartments has also been demonstrated in Arabidopsis (Chew et al., 2003). Cytosolic GR activity is encoded by GR1 (At3g24170) but proteomic analysis suggested that this gene also encodes peroxisomal GR activity (Kaur et al., 2009). It has recently been shown that import of GR1 into the peroxisomes occurs through the Peroxisomal Targeting Sequence 1 (PTS1) pathway, though targeting efficiency was considered to be relatively weak (Kataya and Reumann, 2010). From results obtained in pea, it seems that the peroxisomal enzyme represents a relatively minor part of the GR1 gene product (Romero-Puertas et al., 2006), which overall accounts for 30 to 60% of the total leaf enzyme activity in Arabidopsis (Marty et al., 2009; Mhamdi et al., 2010a). Available information therefore suggests that GR1 is mainly targeted to the cytosol, though an intriguing question concerns the physiological significance of its localization in the peroxisomes, which may be important sources of ROS and related signals (Foyer and Noctor, 2003; del Río et al., 2006; Nyathi and Baker, 2006; Mhamdi et al., 2010c).

7.3.2. Overexpression Studies

Glutathione reductase has been overexpressed in a number of plant species. In all cases the reported effects of enhanced GR activity in the chloroplast or cytosol are relatively modest, indicating that under many conditions, activity is sufficient to maintain glutathione pools and function (Aono et al., 1993; Broadbent et al., 1995; Foyer et al., 1991, 1995; Korneyev et al., 2005; Ding et al., 2009). However, increases in GR activity do increase the reduction state of the ascorbate pool (Foyer et al., 1995). These observations are entirely consistent with efficient coupling of the reactions of the ascorbate-glutathione pathway. Nevertheless, as noted above, there are specific situations such as increased availability of H2O2 where the glutathione pool becomes preferentially oxidized, suggesting that GR may become limiting for GSSG reduction. This could be a key property of pathway regulation, allowing altered glutathione status to transmit signaling information, and is discussed further below (section 8.1.1).

7.3.3. Mutagenesis Studies in Arabidopsis

Arabidopsis T-DNA mutants for GR2 are embryo-lethal (Tzafrir et al., 2004), presumably reflecting the crucial importance of maintaining appropriate glutathione reduction states in plastids, mitochondria, or both, at least during certain developmental stages. In contrast, gr1 knockout mutants are aphenotypic, despite a significant decrease in both total extractable GR activity and GSH:GSSG ratios (Marty et al., 2009; Mhamdi et al., 2010a). The aphenotypic nature of gr1 mutants is due to the ‘back-up’ GSSG regeneration activity by the cytosol-located NTR-TRX system (Marty et al., 2009). The TRX system also replaces GR in insects and yeast (Kanzok et al., 2001 ; Meyer et al., 2008). Triple mutants in which both NTRs and GR1 are knocked out show a male-sterile phenotype, demonstrating the requirement for at least one of the systems for plant survival (Marty et al., 2009). Further, it was reported that gr1 mutants were little affected in abiotic stress responses, and no increase in sensitivity to exogenously supplied H2O2 was observed (Marty et al., 2009). Based on this study, the specificity of GR1 or NTR-TRX function with regard to glutathione reduction remained unclear. However, when the gr1 mutation is introduced into the cat2 background, in which intracellular H2O2 places an increased load on regeneration of GSH, the resulting double mutants show a much exacerbated phenotype compared to cat2 (Mhamdi et al., 2010a). This effect is linked to dramatic oxidation and accumulation of glutathione in cat2 gr1, while the gr1 mutation is associated with compromised responses to pathogens and altered expression of genes involved in defense hormone signaling (Mhamdi et al., 2010a). These observations strongly suggest that, although not required for growth in controlled environment conditions, GR1 plays a specific role in stress, particularly biotic stress. They also underline the importance of the compartmentation of H2O2 production in the analysis of potential roles of specific components such as GR1 in redox homeostasis and stress responses. Further, transcriptomic patterns suggest that GR-GRX systems may function largely independently of TRXs in oxidative stress signaling, at least at the level of transcript abundance (Mhamdi et al., 2010a). However, the results of Marty et al. (2009) suggest that, under conditions in which GSSG accumulates, the resulting engagement of the NTR-TRX system could be one mechanism potentially contributing to TRX oxidation and signaling.

8. GLUTATHIONE IN REDOX SIGNALING

While thiol-disulfide redox signaling through the chloroplast TRX system was well established by the end of the 1970's, attention has been focused on glutathione-linked redox signaling only more recently (Foyer et al., 1997; May et al., 1998a). Both for TRX and for glutathione, essential information was generated by the sequencing and annotation of the Arabidopsis genome. This advance revealed the great number and diversity of TRXs and GRXs (Laloi et al., 2001; Lemaire, 2004; Lemaire et al., 2007; Meyer et al., 2008; Rouhier et al., 2008; Rouhier, 2010), and was the starting-point for the ongoing characterization of glutathione-mediated redox signaling systems in plants.

8.1. Glutathione Redox Potential and NADPH: Theoretical Considerations and Potential Importance in Signaling

A key outstanding issue is to what extent GR activity ensures that the glutathione and NADP(H) couples are in thermodynamic equilibrium (at the same redox potential). At cytosolic glutathione concentrations of 2 to 5 mM (Meyer et al., 2001; Queval et al., 2011), and an NADP(H) redox potential of -320 mV (ie, NADP+:NADPH = 1), redox equilibrium would require the GSH:GSSG ratio to be in the range of 105 to 106. In plant tissues such as leaves in the absence of stress, measured GSH:GSSG ratios in extracts are much lower, typically about 20 (eg, Queval and Noctor, 2007; Mhamdi et al., 2010a), while intracellular ratios in other organisms are often cited as being in the range of 102 to 103 (eg, Trotter and Grant, 2002). Chung et al. (1991) proposed that GSH inhibition of GR prevents the two redox couples from reaching equilibrium. Another possibility is that global cellular or tissue measurements, while a useful indicator of redox perturbation, do not precisely reflect GSH:GSSG ratios in certain intracellular compartments, and there is accumulating evidence that this is indeed the case.

Figure 7.

Scheme of glutaredoxin(GRX)-dependent oxidative signaling potentially mediated by changes in glutathione redox potential. The relationship between glutathione redox potential and GSH:GSSG is shown for two total glutathione concentrations, assuming a midpoint potential of -230 mV and constant pH. NADPH:NADP* ratios at two glutathione redox potentials are shown assuming thermodynamic equilibrium between NADP(H) and glutathione. The horizontal lines on the graph indicate a redox potential span of about 100 mV corresponding to NADPH:NADP+ values of 20 (top) to 0.01 (bottom). GRXred, reduced glutaredoxin. GRXox, oxidized glutaredoxin. For further discussion, see text.

8.1.1. Insights from studies with redox-sensitive GFP

In vivo roGFP studies suggest that the glutathione redox potential in wild-type Arabidopsis in the absence of stress is lower than -300 mV in the cytosol and in other compartments (Meyer et al., 2007; Schwarzlander et al., 2008; Jubany-Mari et al., 2010). Given that the cytosolic NADPH:NADP+ ratio is not far removed from one (Igamberdiev and Gardeström, 2003), these redox potential values suggest that the two couples are close to redox equilibrium in this compartment. Such a situation could confer high sensitivity on the signaling function of glutathione redox potential mediated through GRX-dependent changes in protein thiol-disulfide status. A change in redox potential of about 50 mV is sufficient to significantly alter the balance between oxidized and reduced forms of TRX-regulated proteins (Setterdahl et al., 2002). For example, an increase in redox potential from -350 to -300 mV converts the TRX-regulated chloroplast glucose-6-phosphate dehydrogenase 1 (At5g35790) from almost completely inactive to active (Née et al., 2009).

The relationship between glutathione redox potential and redox state is shown in Figure 7. Knockout mutants for cytosolic NADP-isocitrate dehydrogenase (At1g65930) do not show a significant difference in whole leaf NADP+:NADPH ratios, but the absence of this enzyme does modify the GSH:GSSG ratio under conditions of increased H2O2 availability (Mhamdi et al., 2010b). This is consistent with the notion that changes in NADP reduction state that are undetectable in whole tissues may be sufficient in vivo to allow significant changes in glutathione redox potential, and hence signal amplification. Such effects may be of physiological significance, especially if localized to specific cells in certain conditions, eg, during plant responses to pathogens. Given the low KMof GR for NADPH (Smith et al., 1989; Edwards et al., 1990) compared to likely cytosolic NADPH concentrations (around 150 µM: Igamberdiev and Gardeström, 2003), this could occur through adjustment of reactant concentrations in the NADP-glutathione equilibrium rather than kinetic limitation of GR activity by NADPH.

The above discussion suggests that glutathione oxidation during stress is unlikely to be caused by kinetic limitation of GR by reductant but that decreases in NADPH:NADP+ could impact thermodynamically on the glutathione pool. Moreover, a substantial body of literature data shows that the extractable activity of GR is the lowest of the classical ascorbate-glutathione pathway enzymes. Given that enzymes other than DHAR may also contribute to GSH oxidation under oxidative stress conditions (Figure 6), the relatively low capacity of GR may be important to allow GSSG accumulation to fulfil signaling functions. This may involve gradual uncoupling of ascorbate and glutathione reduction states as flux through the pathway accelerates. This notion is consistent with the behaviour of the two pools observed in several studies of catalase-deficient plants.

8.1.2. The significance of glutathione concentration

A potentially important factor in redox signaling is glutathione concentration. This is because the actual glutathione redox potential is related to [GSH]2:GSSG (Mullineaux and Rausch, 2005; Meyer, 2008). Thus, if the GSH:GSSG ratio remains unchanged, decreases in glutathione concentration in themselves cause the redox potential to increase, ie, become more positive (Figure 7). The cytosolic thiol/disulfide redox potential as detected by roGFP was more positive in both cad2 and gr1 mutants (Meyer et al., 2007; Marty et al., 2009), underscoring the contributions of concentration (decreased in cad2) and GSH:GSSG (decreased in gr1) to the measured redox potential. Quantification of the glutathione redox potential in triple clt1 clt2 clt3 mutants, perturbed in chloroplast envelope glutathione transport, revealed an increase of about 20 mV in the cytosol while the plastidial potential was not affected (Maughan et al., 2010). Decreased reduction state of GRXs caused by a more positive redox potential could explain observations in plants that both lack NTR and are deficient in glutathione (Reichheld et al., 2007; Bashandy et al., 2010).

Given that glutathione concentration can clearly increase in response to decreases in GSH:GSSG, it is unclear why the GSH:GSSG ratio does not seem to increase to compensate for decreases in glutathione concentration. One possibility is that the GSH:GSSG ratio is already extremely high in some compartments, as discussed above, and so cannot be increased further in response to decreases in glutathione concentration. Because the KMGSSG values of GRs have been determined to be in the 10–50 µM range (Smith et al., 1989; Edwards et al., 1990), this may partly reflect a limitation over redox equilibrium with NADPH at very low GSSG concentrations.

8.2. Glutathione in the Regulation of Plant Development

Cell identity has a profound influence on signaling processes that govern cell fate (Jiang et al., 2006b) and responses to abiotic stress (Dinneny et al., 2008). Several lines of evidence demonstrate that glutathione status is important in plant development. The most convincing has come from analysis of the phenotypes of GSH-deficient Arabidopsis mutants, which has shown that glutathione has critical functions in embryo and meristem development (Vernoux et al., 2000; Cairns et al., 2006; Reichheld et al., 2007; Frottin et al., 2009; Bashandy et al., 2010). Other evidence has been provided by studies of cell proliferation using cell cultures or analysis of quiescence in meristems (Kerk and Feldman, 1995; Potters et al., 2002; 2004; Diaz Vivancos et al., 2010a,b). Glutathione status has also been implicated in influencing the regeneration efficiency of somatic embryos (Belmonte et al., 2005). The role of glutathione in the cell cycle has recently been reviewed (Diaz-Vivancos et al., 2010b).

8.3. Glutaredoxins

Thiol-disulfide exchange reactions are a fundamental mechanism of regulation of enzyme activity in plants (Buchanan and Balmer, 2005). While several chloroplast enzymes are regulated in this way by TRX, GRX may also couple glutathione redox potential to changes in protein thiol-disulfide status (Figure 7). Compared to many other organisms, terrestrial plants contain a large GRX family. These genes were classified by Lemaire (2004) into subfamilies depending on predicted active site motifs, with additional GRX or similar sequences described more recently (Navrot et al., 2006; Rouhier et al., 2006; Meyer et al., 2008). Based on a four amino acid motif at the active site sequence, the encoded proteins are grouped into ‘classical’ CxxC/S GRX (class I), ‘monothiol’ CGFS GRX (class II), and CC-type GRX (class III), with the last sub-class being specific to land plants (Lemaire, 2004; Rouhier, 2010). While specific nomenclatures have been proposed for some of these GRXs (Ziemann et al., 2009; Rouhier, 2010), much remains to be discovered regarding their functions and those of related sequences that have been described in Arabidopsis.

8.3.1. Class I and Class II Glutaredoxins

The activities of class I GRXs include thiol-disulfide exchange, (de)glutathionylation, and regeneration of PRX and methionine sulfoxide reductase (MSR; Rouhier et al., 2002, 2006; Zaffagnini et al., 2008; Tarrago et al., 2009; Gao et al., 2010). A DHAR activity has long been known for GRX (Wells et al., 1990) and appears to be specific to this type, but its physiological importance remains unclear. Members of both classes of GRX have been shown to play roles in assembly of iron-sulfur clusters (Rouhier et al., 2007; Bandyopadhyay et al., 2008), though this activity is most strongly associated with class II GRX (Rouhier, 2010).

Arabidopsis class II GRXs have been identified as interacting with ion channels or as involved in oxidative stress responses (Cheng and Hirschi, 2003; Cheng et al., 2006). The plastidic glutaredoxin CXIP1 (AtGRXcp1; At3g54900) is able to rescue the sensitivity of yeast GRX5 mutants to H2O2 and Arabidopsis mutants for this GRX have slightly increased sensitivity to exogenous H2O2 (Cheng et al., 2006). Manipulation of the expression of a tomato GRX with a CGFS active site in both Arabidopsis and tomato also modulates the antioxidative system and alters stress tolerance (Guo et al., 2010). Expression of a GRX from an arsenic-hyperaccumulating fern was also reported to increase thermotol-erance in Arabidopsis (Sundaram and Rathinasabapathi, 2010).

As discussed further below, both class I and class II GRX have been shown to be competent in protein deglutathionylation in vitro (Gao et al., 2009a). Although members of the CGFS (class II) sub-family are sometimes referred to as ‘monothiol’ GRX, in at least some cases catalytic activity involves formation of an intramolecular disulfide bond with a second cysteine located distally from the GRX active site motif. These types of GRX may be dependent on TRX reductase rather than GSH for their reduction (Zaffagnini et al., 2008; Gao et al., 2009a).

8.3.2. Class III Glutaredoxins

Compared to the other GRX, relatively little is known about the biochemistry of the plant-specific class III GRX (CC-type or ROXY GRX). However, certain members of this class are the best characterized GRX in terms of physiological function, with three members implicated in plant development and defense. GRX480/ROXY19 (At1g28480) overexpression represses the JA marker gene, PDF1.2 (Ndamukong et al., 2007) while ROXY1 (At3g02000) and ROXY2 (At5g14070) are involved in petal and anther development (Li et al., 2009). Both of these functions involve interaction with TGA transcription factors. Complementation experiments showed that the rice orthologs of ROXY1 or ROXY2 can rescue the defect in petal development in Arabidopsis roxy1 mutants, as can GRX480 (Li et al., 2009; Wang et al., 2009). Overexpression of Arabidopsis or rice ROXY1 in Arabidopsis also increased susceptibility to Botrytis cinerea (Wang et al., 2009). This biochemical redundancy suggests that the key factor that determines the specificity of Class III GRX function may be temporal and/or spatial expression patterns (Ziemann et al., 2009). Both this notion and the proposed roles of this type of GRX in ROS signaling receive support from the analysis of plants with perturbations in glutathione status (Mhamdi et al., 2010a). H2O2-triggered alteration of glutathione status in cat2 and cat2 gr1 mutants modifies the abundance of transcripts encoding four class III GRXs but not other GRX types. These genes included GRX480 as part of a generalized effect on JA-associated genes, while ROXY1 or ROXY2 expression were not affected (Mhamdi et al., 2010a).

8.4. Protein S-Glutathionylation

Glutathione can form stable mixed disulfide bonds with protein cysteine residues (protein S-glutathionylation or thiolation). S-Glutathionylation is a ubiquitous redox-sensitive and reversible modification of cysteinyl residues that can directly regulate the activity of the affected protein. Isolated reports of protein S-glutathionylation in plants have appeared over the years, eg, for seed acyl carrier protein (Butt and Ohlrogge, 1991), but it is only over the last decade that systematic attention has been paid to the significance and potential of this process. Elucidation of this phenomenon remains technically challenging, but several techniques have been used to identify target proteins (lto et al., 2003; Dixon et al. 2005; Michelet et al., 2005, 2008; Zaffagnini et al., 2007; Holtgrefe et al., 2008; Gao et al., 2009a). To date, the information derived is largely qualitative and probably preferentially detects relatively abundant proteins. Further, the in vivo significance of the identified modifications remains in many cases to be established.

An intriguing finding is that f-type TRX is glutathionylated on a cysteine residue external to the active site, decreasing its activity against target proteins such as plastidial NADP-glyceraldehyde3-phosphate dehydrogenase (GAPDH; Michelet et al., 2005). The activity of another TRX-independent isoform of GAPDH is also decreased by glutathionylation (Zaffignini et al., 2007). Therefore, in conditions leading to glutathionylation in the chloroplast, the activities of both types of GAPDH are likely to be decreased, either directly or indirectly, through glutathionylation of TRX f. These observations point to a new layer of complexity in thiol-disulfide regulation of chloroplast metabolism. Interestingly, some of the proteins regulated by TRXs and glutathione, including the glutathionylated GAPDH isoform, are known to associate into complexes, eg, with CP12, in a TRX-dependent manner (Howard et al., 2008).

Recent studies also point to thiol-disulfide regulation of metabolism outside the chloroplast. The activity of the mitochondrial enzyme, glycine decarboxylase (GDC), is sensitive to oxidative stress (Taylor et al., 2002), and this enzyme has recently been shown to be inhibited by thiol modifications on several cysteine residues (Palmieri et al., 2010). Given that NADH production through this enzyme is highly active during photorespiration, this mechanism may act together with other thiol modifications, such as those that regulate mitochondrial alternative oxidase activity (Vanlerberghe et al., 1995; Gelhaye et al., 2005), as part of redox regulation related to ROS production by the respiratory electron transport chain. Glycolytic NAD-dependent GAPDH is a cytosolic enzyme that is also partly associated with mitochondria and the nucleus (Giegé et al., 2003; Anderson et al., 2004). This enzyme is inhibited by oxidation (Hancock et al., 2006), and has been identified as a glutathionylated protein (Dixon et al., 2005; Holtgrefe et al., 2008).

The in vivo mechanisms underlying S-glutathionylation remain to be unequivocally established. They may involve thiol-disulfide exchange between protein thiol groups and GSSG, perhaps catalyzed by certain GRX, occur through conversion of cysteine thiol groups to thiyl radicals or sulfenic acids, or be mediated by GSNO (Dixon et al., 2005; Gao et al., 2009b; Holtgrefe et al., 2008; Palmieri et al., 2010). Several possible routes of deglutathionylation are also possible, notably including GRX-dependent monothiol and dithiol mechanisms (Gao et al., 2009b). Biochemical analysis of Chlamydomonas proteins has shown that some class I and class II GRX are competent in deglutathionylation reactions, though there appears to be specificity within both groups (Zaffagnini et al., 2008; Gao et al., 2009a, 2010). Some class II GRXs appear to catalyze deglutathionylation through a monothiol mechanism, the intermediate glutathionylated GRX likely being resolved by formation of an intramolecular disulfide between the active site Cys and a partially conserved C-terminal Cys. This disulfide was shown to possess a low redox potential and to be reduced by thioredoxin reductase rather than by GSH (Zaffagnini et al., 2008). For some class II GRXs, the reductant of the intermediate glutathionylated GRX is TRX, not GSH, and the dithiol mechanism involves a disulfide bond with a Cys residue distal from the active site motif (Zaffagnini et al., 2008). Monothiol deglutathionylation mechanisms include CPYC-dependent reactions involving only the active site Cys that is closer to the N-terminus, as well as deglutathionylation via a poplar GRX with a CSYS active site (Couturier et al., 2009). GRX-catalyzed deglutathionylation through a monothiol mechanism is also likely part of the catalytic cycle of Arabidopsis MSRB1 (At1g53670) whereas MSRB2 (At4g21860) is regenerated through a TRX-dependent dithiol mechanism (Tarrago et al., 2009). While TRXs are not considered to be directly active against mixed disulfides, a role for yeast TRX in deglutathionylation has recently been suggested (Greetham et al., 2010).

Although several class III GRX showed modified transcript abundance in response to disulfide stress in Arabidopsis, no other type of GRX or TRX was affected, even when GSSG accumulation was very marked (Mhamdi et al., 2010a). This suggests either that transcriptional control of these types of proteins is not an important part of the response or that accumulation of GSSG is not in itself sufficient to induce significant protein S-glutathionylation.

8.5. Protein S-Nitrosylation and GSNO Reductase

Another glutathione-linked modification of protein cysteine residues is S-nitrosylation (Lindermayr et al., 2005; Romero-Puertas et al., 2008; Lindermayr and Durner, 2009). GSNO may be an important physiological NO donor, for example, in the regulation of NPR1, where nitrosylation of proteine cysteine residues acts antagonistically to reduction by TRX (Tada et al., 2008). Inhibition of a methionine adenosyltransferase by GSNO-triggered S-nitrosylation has been proposed to be an important regulatory mechanism in ethylene synthesis (Lindermayr et al., 2006), while modification of PRX by S-nitrosylation has been implicated in the regulation of ROS signaling (Romero-Puertas et al., 2007).

Cytosolic GAPDH can be S-nitrosylated as well as S-glutathionylated (Lindermayr et al., 2006). Similar to the GDC P protein studied in leaf mitochondrial extracts (Palmieri et al., 2010), cysteine groups on purified GAPDH were found to undergo both modifications after GSNO treatment (Holtgrefe et al., 2008; Palmieri et al., 2010). Both for GDC and GAPDH, the modifications were associated with a loss of activity. The inhibition of GDC by thiol modifications adds to the growing awareness of the potential importance of photorespiration-linked redox reactions in plant stress responses (Foyer et al., 2009; Petershansel et al., 2010) while oxidative inhibition of GAPDH could contribute to regulating relative fluxes through the oxidative pentose pathway and glycolysis during stress (Holtgrefe et al., 2008).

Specific NO-producing enzymes in plants remain to be fully characterized, but GSNO reductase could be influential in regulating GSNO-dependent reactions. This enzyme is encoded by a single gene in Arabidopsis (At5g43940), and catalyzes the NADH-dependent reduction of GSNO to S-aminoglutathione, which then decomposes to ammonia and free GSH (Sakamoto et al., 2002; Feechan et al., 2005; Díaz et al., 2006; Barroso et al., 2006; Lee et al., 2008; Chen et al., 2009). As well as interactions with NO, glutathione has been implicated in the possible signaling functions of carbon monoxide in Medicago sativa (Han et al., 2007).

8.6. Glutathione in Environmental Responses

8.6.1. Light Signaling

Evidence for the role of glutathione in irradiance-dependent signaling came from the identification of the Arabidopsis rax1 mutant. This line contains less than 50% wild-type glutathione contents, and shows enhanced constitutive expression of the high light-induced gene, APX2 (Ball et al., 2004). Interestingly, overexpression of a tomato CGFS-type (class II) GRX in Arabidopsis also enhances expression of APX2 (Guo et al., 2010). Several studies in Arabidopsis and other species point to a potential link between growth daylength, light signaling and glutathione status (Becker et al., 2006; Queval et al., 2007; Bartoli et al., 2008). Genetic evidence of a link between glutathione and photoreceptor signaling has been provided by identification of arsenic-tolerant mutants, called ars4 and ars5 (Sung et al., 2007). While ars4 was shown to be a phytochrome A (phyA) mutant (Sung et al., 2007), ars5 is affected in a component of the 26S proteasome (Sung et al., 2009). Mutants for phyA showed increased resistance to BSO as well as to arsenic (Sung et al., 2007), and ars5 had enhanced levels of glutathione when exposed to arsenic, accompanied by increased GSH1 and GSH2 transcripts (Sung et al., 2009).

8.6.2. Biotic Interactions, Salicylic Acid, and Phytoalexin Synthesis

It is well established that exogenous GSH acts similarly to fungal elicitors in activating the expression of defense-related genes (Wingate et al., 1988; Dron et al., 1990) including PATHOGENESIS-RELATED1 (PR1, Gomez et al., 2004b; Senda and Ogawa, 2004). Moreover, accumulation of glutathione is triggered by pathogen infection (Edwards et al., 1991; May et al., 1996a), and this can involve characteristic transient changes in glutathione redox state (Vanacker et al., 2000; Parisy et al., 2006). Similar changes have also been reported following exogenous application of the defense-related hormone salicylic acid (SA), or biologically active SA analogs (Mou et al., 2003; Mateo et al., 2006; Koornneef et al., 2008). Glutathione perturbation in catalasedeficient cat2 is also linked to hypersensitive response (HR)-like lesions as part of a wide spectrum of defense responses that are conditionally induced in this line (Chaouch et al., 2010).

Thiol-disulfide status is clearly involved in the regulation of the SA-dependent NPR1 pathway (Déspres et al., 2003; Mou et al., 2003; Rochon et al., 2006; Tada et al., 2008). It was proposed that SA-dependent PR1 induction requires reduction of disulfide bonds of the oligomeric form of NPR1 that could be mediated by TRX or GRX in vivo (Mou et al., 2003) A more recent study reported that monomerization of the NPR1 protein to the active form was catalyzed by TRX but that glutathione is involved through GSNO-dependent nitrosylation of NPR1 monomers (Tada et al., 2008). The reoligomerization was postulated to be necessary to prevent protein depletion (Tada et al., 2008). The identification of NPR1 as a target of S-nitrosylation underpins the involvement of GSH in plant disease resistance, in which protein S-nitrosylation is considered to be a key post-translational modification. The Atgsnor1 mutant for the enzyme GSNO reductase has compromised resistance to both virulent and avirulent pathogens (Feechan et al., 2005).

Further evidence that glutathione is involved in defense against biotic stress came from the studies of Arabidopsis mutants altered in GSH synthesis or metabolism. Studies on the effect of GSH deficiency have reported conflicting results. Analysis of responses of cad2 to virulent and avirulent strains of the oomycete Peronospora parasitica and the bacterium Pseudomonas syringae did not reveal compromised resistance (May et al., 1996b) but subsequent studies of cad2 and rax1-1 reported increased susceptibility to avirulent strains of P. syringae (Ball et al., 2004). Triple mutants defective in glutathione transport across the chloroplast envelope showed decreased expression of PR1 and resistance to the oomycete Phytophthora brassicae (Maughan et al., 2010). The pad2 mutant is deficient in the Arabidopsis phytoalexin camalexin (Glazebrook and Ausubel, 1994) and shows enhanced susceptibility to various bacterial, fungal and oomycete pathogens as well as to insects (Ferrari et al., 2003; Parisy et al., 2006; Schlaeppi et al., 2008). The affected gene in pad2 was identified as GSH1 and the line was shown to have GSH contents that were lower than those in cad2 (Parisy et al., 2006). A comparative study of the allelic mutants rax1, cad2, and pad2 revealed an inverse correlation between resistance to P. brassicae and glutathione contents, though only pad2 was markedly affected in camalexin contents and its response to P. syringae (Parisy et al., 2006). These observations suggest that there is a critical glutathione status below which the accumulation of pathogen-defense related molecules and consequently disease resistance, is impaired. This level could vary according to conditions and pathogen specificity. In the case of camalexin synthesis, it is not yet clear whether GSH acts as a regulatory component or as a precursor of the thiazole ring, or both (Parisy et al., 2006; Glawischnig, 2007; Nafisi et al., 2007). However, knockout mutants for GR1, which are not glutathione-deficient but which have a decreased GSH:GSSG ratio, accumulate less SA than wild-type and show increased sensitivity to virulent P. syringae (Mhamdi et al., 2010a). This observation suggests that glutathione redox potential could be one factor that determines some of the effects reported in glutathione-deficient mutants.

The components that link changes in glutathione status to pathogen responses remain for the most part unidentified. As well as GRX, GSNO, and S-glutathionylation reactions, GSTs may play important roles (Lieberherr et al., 2003; Sappl et al., 2004; Davoine et al., 2006). Induction of GSTs by SA and other hormones has been systematically studied (Sappl et al., 2009). An active process occuring at the pathogen entry site consists of conjugation of GSH to insect-deterrent indole glucosinolate hydrolysis products to mediate the accumulation of other antifungal glucosinolate compounds (Bednarek et al., 2009). An activation of both GST transcripts and protein activity was also observed in cryptogein-induced tobacco HR together with the conjugation of oxylipins to GSH. This was suggested to be an HR-linked event with a signaling role during plant defense (Davoine et al., 2006).

Another plant-microbe interaction in which GSH plays a significant role is symbiosis between legumes and nitrogen-fixing rhizobia, which involves the production of nodules. These organs contain high levels of both GSH and homoglutathione, and deficiency in these molecules inhibits nodule formation (Frendo et al., 2005; Pauly et al., 2006).

8.6.3. Glutathione and Cell Death

It has been proposed that the glutathione redox state is a key determinant of cell death and dormancy (Kranner et al., 2006). According to this model, an increase in redox potential above a threshold reducing value necessarily causes death and/or growth arrest. Increased glutathione leaf contents triggered by γ-ECS overexpression in the tobacco chloroplast triggered lesion formation and induced PR gene expression (Creissen et al., 1999). In tobacco deficient in catalase, lesions also appear under conditions that cause accumulation of GSSG (Chamnongpol et al., 1996; Willekens et al., 1997). Double Arabidopsis knockouts for catalase and GR1 (cat2 gr1) show dramatic oxidation of glutathione (more than 2 µmol GSSG g-1 FW, representing 95% of the leaf glutathione pool). While cat2 gr1 undergoes leaf bleaching when grown in long days, this effect is associated with repression rather than stimulation of SA-dependent pathogen responses (Mhamdi et al., 2010a), even though a wide range of these responses is associated with the less marked glutathione oxidation observed in cat2 single mutants (Chaouch et al., 2010). Moreover, H2O2-triggered lesions on the leaves of cat2 can be reverted by blocking SA synthesis through the isochorismate pathway or by treating leaves with myo-inositol, but in neither case does the suppression of cell death correlate with less oxidized tissue glutathione status (Chaouch and Noctor, 2010; Chaouch et al., 2010). Glutathione perturbation is equally or more marked in cat2 gr1 in short day as in long day growth conditions, but neither bleaching nor necrotic lesions are observed in the double mutant in short days (Mhamdi et al., 2010a). These observations suggest that perturbation of glutathione status is not in itself sufficient to cause bleaching or cell death. Accumulation of GSSG in cat2 grown from seed in air is associated with a stunted phenotype, possibly related to the roles of thiols in the plant cell cycle, growth and development (Reichheld et al., 1999, 2007; Vernoux et al., 2000). However, significant GSSG accumulation is also observed in gr1 knockout mutants, which are aphenotypic (Marty et al., 2009; Mhamdi et al., 2010a). Together with the recent study of Liedschulte et al. (2010), the available data suggest that high levels of glutathione (as GSH or GSSG) can be tolerated by plants and are not in themselves sufficient to trigger cell death pathways. As well as cellular tolerance of shifts in glutathione status, subcellular compartmentation could be a key issue.

One explanation of the difference between the effects of overexpressing monofunctional γ-ECS and bifunctional γ-ECS/ GSH-S in the tobacco chloroplast is that lesions are caused by γ-EC accumulation. This metabolite can accumulate to high levels when γ-ECS is overexpressed alone, either in the cytosol or the chloroplast (Noctor et al., 1996, 1998a; Creissen et al., 1999). Support for this idea comes from the observation that tobacco overexpressing both γ-ECS and GSH-S had an ameliorated phenotype compared to tobacco expressing the first enzyme alone (Creissen et al., 1999). Differences in γ-EC accumulation may therefore explain the differences between the studies of Creissen et al. (1999) and Liedschulte et al. (2010). However, the extent of γ-EC accumulation in tobacco overexpressing the bifunctional bacterial enzyme has not yet been documented, and there are several lines of evidence that argue against this interpretation. The first is that γ-EC was increased rather than decreased in tobacco overexpressing both enzymes of glutathione synthesis (Creissen et al., 1999). Second, abolition of photorespiratory glycine production by high CO2 enhances γ-EC accumulation in poplar expressing E.coli γ-ECS in the cytosol or chloroplast. In this condition, γ-EC can accumulate to more than 1 µmol.g-1 FW without triggering lesions (Noctor et al., 1997, 1999). Third, H2O2activated glutathione accumulation in Arabidopsis cat2 plants is accompanied by increased γ-EC contents and this does not in itself promote lesions (Queval et al., 2009). Currently available information suggests that neither total glutathione nor GSSG accumulation necessarily induces cell death in tobacco, poplar or Arabidopsis, and that further work is required to establish the significance of γ-EC accumulation in lesion formation. Despite this, there is a clear though apparently complex link between glutathione status and SA-dependent cell death.

8.6.4. Jasmonic Acid and Responses to Herbivores

While glutathione status is clearly implicated in SA signaling, several strands of evidence point to an equally influential role in JA signaling. The first indication of a link was provided by Xiang and Oliver (1998), who reported JA induction of transcripts encoding enzymes of GSH synthesis and GR. Subsequent work has shown that other antioxidative genes can be induced by JA (Sasaki-Sekimoto et al., 2005) while this hormone has been implicated, together with thiol status, in resistance to selenite (Tamaoki et al., 2008). Overexpression of GRX480 represses JA-triggered induction of the defensin, PDF1.2 (Ndamukong et al., 2007) and SA-dependent repression of this marker gene is correlated with changes in glutathione status and opposed by inhibition of glutathione synthesis (Koornneef et al., 2008).

Analysis of cat2-deficient plants provided some evidence that H2O2 may oppose JA-regulated gene expression (Vanderauwera et al., 2005). A comparative analysis of cat2 and gr1 knockouts suggests that the H2O2-JA interaction is daylength-conditioned and may be at least partly mediated by glutathione status (Mhamdi et al., 2010a). Based on JA- or oxylipin-dependent expression patterns (Yan et al., 2007; Mueller et al., 2008), the peroxidatic or conjugation activities of GSTs against electrophilic metabolites (Wagner et al., 2002; Davoine et al., 2006), and the known interaction of CC-type GRX with transcription factors (Ndamukong et al., 2007; Li et al., 2009), glutathione may influence JA signaling through several GSTs and GRX. Specific genes within the GST family and the class III GRX subfamily show modified expression in response to changes in glutathione status in caf2 and gr1 mutants (Mhamdi et al., 2010a). Interestingly, GSNO reductase transcripts are decreased by both wounding and JA (Diaz et al., 2003).

Resistance to some insects is compromised in the glutathione-deficient pad2 mutant and this is linked to decreased glucosinolates (Schlaeppi et al., 2008). The effect could be reverted by GSH treatment but not DTT, possibly suggesting that it is linked to a requirement for GSH in biosyntheses rather than mediated via redox changes (Schlaeppi et al., 2008). This is consistent with studies in tobacco engineered to express the Brassica glucosinolate pathway (Geu-Flores et al., 2009). It therefore seems that changes in glutathione status could impact functionally on JA-dependent stress responses through both a redox signaling effect, possibly mediated by GRX, and a requirement for GSH in glucosolinate synthesis.

9. CONCLUSION AND PERSPECTIVES

The list of cellular processes that are dependent on or influenced by glutathione in plants continues to grow. Probably the most exciting development over the last decade is the emergence of glutathione as an important signaling molecule. It is now clear that glutathione status can interact with several phytohormone-dependent pathways (Figure 8). The view of signal transduction operating in isolation through linear pathways is gradually being replaced by one that takes fuller account of cellular complexity. Arguably, the integration of signal transduction mechanisms in plants can be seen in at least some cases as analogous to metabolic control theory, where no single component is ever fully dominant. From a physiological ‘whole plant’ perspective, it is perhaps in this context that major redox buffers like glutathione act as an interface between the environment and the cell. Through its interactions with sulfur and nitrogen metabolism, with ROS and ascorbate via peroxidases and DHAR, with NADPH via GR, with protein thiol-disulfide groups through GRX, and with signaling metabolites and hormones via GSTs, glutathione is an outstanding candidate as a central “hub” molecule whose status can modulate other signaling pathways.

Figure 8.

Hormone-dependent growth and stress responses affected by glutathione status.

ABA, abscisic acid. JA, jasmonic acid. SA, salicylic acid. Other abbreviations as defined in other legends.

Many of the most significant breakthroughs have been made possible by studies of glutathione-deficient Arabidopsis mutants. While there is a clear requirement for glutathione in antioxidant metabolism as well as specific features of abiotic stresses involving detoxification of some heavy metals and xenobiotics or other potentially toxic compounds, current information concerning biotic stress points to at least two ways in which glutathione availability influences responses in Arabidopsis (Figure 8). The first is as a co-factor or sulfur source for the production of compounds such as camalexin and glucosinolates. The second involves a signaling function related to glutathione redox status and/or concentration. Regarding the second, a key issue is the importance of changes in glutathione status that may be monitored as glutathione redox potential (thermodynamic signaling). However, it is still unclear whether the redox potential of glutathione is a principal player or whether other factors such as modified abundance of specific glutathione-dependent proteins or reactive glutathione derivatives are more important. To date, documented changes in glutathione redox potential, whether in mutants or in stress conditions, are of the order of 20–30 mV, and it remains to be established whether these are sufficient to greatly modify the biological activity of dependent proteins. A related issue is whether the processes affected in studies of glutathione-deficient mutants simply reflect the requirement for a threshold glutathione concentration. Or are these processes also modified by dynamic changes in glutathione that can be induced by the environment or at specific developmental stages or in specific cells?

Despite the insights from studies of Arabidopsis mutants over the last decade, the components that link glutathione to phytohormone signaling remain in many cases to be identified. Although the signaling scheme shown in Figure 7 is based on a classical thiol-disufide exchange reaction, novel types of redox regulation may await discovery, eg, involving the plant-specific class III GRX. An important question concerns the relationship between the biological activity of these proteins and glutathione redox potential or concentration. Similarly, much remains to be described concerning the regulation of the intracellular distribution of glutathione and glutathione-dependent proteins.

While recent studies have established redundancy between glutathione and other thiol systems such as TRX, the physiological significance of this overlap remains to be resolved. Another issue is protein S-glutathionylation. The inventory of potential targets is currently being described, as are the mechanisms potentially regulating glutathionylation and deglutathionylation. An even more challenging task will be to elucidate the functional impact of such changes, both in terms of their quantitative influence on the biological activity of the affected protein, as well as their fractional control over the integrated network of subcellular and cellular regulation. The relative importance of the various mechanisms regulating glutathione synthesis in different conditions also remains to be established, as does the integration of these controls with changes in transport and degradation.

Research effort is gradually turning toward species of obvious agronomic importance. Part of the knowledge generated on glutathione in Arabidopsis may not be completely transferable to other species (eg, because of specificity of defense pathways). Nevertheless, our view is that much of the information generated from studies in Arabidopsis is generic, ie, it concerns cellular processes that cut across taxonomic boundaries. Apart from the well documented advantages of Arabidopsis as a model system, the plethora of specific genetic resources and tools now available should ensure that this species continues to feature strongly in studies on the roles of glutathione and glutathione-dependent processes in plant function.

ACKNOWLEDGMENTS

G.N. is grateful to Bernd Zechmann (University of Graz, Austria) for discussion and comments on glutathione concentrations in different subcellular compartments, and for providing the electron micrograph shown in Figure 2. He also thanks Stéphane Lemaire (Institut de Biologie Physico-Chimique, Paris, France) for discussion and comments on the sections on glutaredoxins and protein glutathionylation reactions. Recent research in this area in the authors' laboratories was partly funded by the ANR-Genoplante project “Redoxome” (Orsay), ANR project “Vulnoz” (Orsay), and the EU Marie-Curie ITN “COSI” project (Orsay, Leeds).

REFERENCES

1.

H. Achkor , M. Díaz , M.R. Fernández , J.A. Biosca , X. Parés and M.C. Martínez (2003). Enhanced formaldehyde detoxification by overexpression of glutathione-dependent formaldehyde dehydrogenase from Arabidopsis. Plant Physiol. 132: 2248–2255 Google Scholar

2.

R.G. Alscher (1989). Biosynthesis and antioxidant function of glutathione in plants. Physiol. Plant 77: 457–464. Google Scholar

3.

L.E. Anderson , M.R. Ringenberg , and A.A. Carol (2004) Cytosolic glyceraldehyde-3-P dehydrogenase and the B subunit of the chloroplast enzyme are present in the pea leaf nucleus. Protoplasma 223: 33–43 Google Scholar

4.

M. Aono , A. Kubo , H. Saji , K. Tanaka , and N. Kondo (1993). Enhanced tolerance to photooxidative stress of transgenic Nicotiana tabacum with high chloroplastic glutathione reductase activity. Plant Cell Physiol. 34: 129–135. Google Scholar

5.

K. Asada (1999). THE WATER-WATER CYCLE IN CHLOROPLASTS: scavenging of active oxygens and dissipation of excess photons. Annu. Rev. Plant Physiol. Plant Mol. Biol. 50: 601–639. Google Scholar

6.

F. Åslund , and J. Beckwith (1999). Bridge over troubled waters: sensing stress by disulfide bond formation. Cell 96: 751–753. Google Scholar

7.

L. Ball , G. Accotto , U. Bechtold , G. Creissen , D. Funck , A. Jimenez , B. Kular , N. Leyland , J. Mejia-Carranza , H. Reynolds , S. Karpinski , and P.M. Mullineaux (2004). Evidence for a direct link between glutathione biosynthesis and stress defense gene expression in Arabidopsis. Plant Cell 16: 2448–2462. Google Scholar

8.

S. Bandyopadhay , F. Gama , M.M. Molina-Navarro , J.M. Gualberto , R. Claxton , S.G. Naik , B.H. Huynh , E. Herrero , J.P. Jacquot , M. Johnson , and N. Rouhier (2008). Chloroplast monothiol glutaredoxins as scaffold proteins for the assembly and delivery of [2Fe-S] clusters. EMBO J. 27: 1122–1133. Google Scholar

9.

J.B. Barroso , F.J. Corpas , A. Carreras , M. Rodríguez-Serrano , F.J. Esteban , A. Fernández-Ocaňa , M. Chaki , M. Romero-Puertas , R. Valderrama , L.M. Sandalio , and L.A. del Río (2006) Localization of S-nitrosoglutathione and expression of S-nitrosoglutathione reductase in pea plants under cadmium stress. J. Exp. Bot. 57: 1785–1793. Google Scholar

10.

C.G. Bartoli , E.A. Tambussi , F. Diego , and C.H. Foyer (2009). Control of ascorbic acid synthesis and accumulation and glutathione by the incident light red/far red ratio in Phaseolus vulgaris leaves. FEBS lett. 583: 118–122. Google Scholar

11.

T. Bashandy , J. Guilleminot , T. Vernoux , D. Caparros-Ruiz , K. Ljung , Y. Meyer , J.P. Reichheld (2010). Interplay between the NADP-linked thioredoxin and glutathione systems in Arabidopsis auxin signaling. Plant Cell 22 : 376–391. Google Scholar

12.

B. Becker , S. Holtgrefe , S. Jung , C. Wunrau , S. Kandlbinder , M. Baier , K.J. Dietz , J.E. Backhausen , and R. Scheibe (2006). Influence of the photoperiod on redox regulation and stress responses in Arabidopsis thaliana L. (Heynh.) plants under long- and short-day conditions. Planta 224: 380–393. Google Scholar

13.

P. Bednarek , M. Pislewska-Bednarek , A. Svatos , B. Schneider , J. Doubsky , M. Mansurova , M. Humphry , C. Consonni , R. Panstruga , A. Sanchez-Vallet , A. Molina , and P. Schulze-Lefert (2009). A glucosinolate metabolism pathway in living plant cells mediates broadspectrum antifungal defense. Science 323: 101–106. Google Scholar

14.

M.F. Belmonte , G. Donald , D.M. Reid , E.C. Yeung , and C. Stasolla (2005) Alterations of the glutathione redox state improve apical meristem structure and somatic embryo quality in white spruce (Picea glauca). J. Exp. Bot. 56: 2355–2364. Google Scholar

15.

J.A. Bick , F. Aslund , Y. Chen , and T. Leustek (1998). Glutaredoxin function for the carboxyl-terminal domain of the plant-type 5′-adenylylsulfate reductase. Proc. Natl. Acad. Sci. USA 95, 8404–8409. Google Scholar

16.

J.A. Bick , A.T. Setterdahl , D.B. Knaff , Y. Chen , L.H. Pitcher , B.A. Zilinskas , and T. Leustek (2001). Regulation of the plant-type 5′-adenylyl sulfate reductase by oxidative stress. Biochemistry 40, 9040–9048. Google Scholar

17.

E. Bloem , S. Haneklaus , I. Salac , P. Wickenhäuser , and E. Schnug (2007). Facts and fiction about sulfur metabolism in relation to plantpathogen interactions. Plant Biol. 9: 596–607. Google Scholar

18.

R. Blum , A. Beck , A. Korfte , A. Stengel , T. Letzel , K. Lendzian , and E. Grill (2007). Function of phytochelatin synthase in catabolism of glutathione-conjugates. Plant J. 49: 740–749. Google Scholar

19.

R. Blum , K.C. Meyer , J. Wünschmann , K.J. Lendzian , and E. Grill (2010). Cytosolic action of phytochelatin synthase. Plant Physiol. 153: 159–169. Google Scholar

20.

J. Bogs , A. Bourbouloux , O Cagnac , A. Wachter, D. Rausch , and S. Delrot (2003). Functional chacterization and expression analysis of a glutathione transporter, BjGT1, from Brassica juncea: evidence for regulation by heavy metal exposure. Plant Cell Environ. 26: 1703–1711. Google Scholar

21.

C. Böttcher , L. Westphal , C. Schmotz , E. Prade , D. Scheel , and E. Glawischnig (2009) The multifunctional enzyme CYP71B15 (PHYTOALEXIN DEFICIENT3) converts cysteine-indole-3-acetonitrile to camalexin in the indole-3-acetonitrile metabolic network of Arabidopsis thaliana. Plant Cell 21: 1830–1845. Google Scholar

22.

A. Bourbouloux , P. Shahi , A. Chakladar , S. Delrot , and A.K. Bachhawat (2000). Hgt1p, a high affinity glutathione transporter from the yeast Saccharomyces cerevisiae. J. Biol. Chem. 275: 13259–13265. Google Scholar

23.

F. Bourgis , S. Roje , M.L. Nuccio , D.B. Fisher , M.C. Tarczynski , C. Li , C. Herschbach , H. Rennenberg , M.J. Pimenta , T.L. Shen , D.A. Gage , and A.D. Hanson (1999). S-methylmethionine plays a major role in phloem sulfur transport and is synthesized by a novel type of methyltransferase. Plant Cell 11: 1485–1497. Google Scholar

24.

M. Brazier-Hicks , K.M. Evans , O.D. Cunningham , D.R.W. Hodgson , P.G. Steel , and R. Edwards (2006) Catabolism of glutathione-conjugates in Arabidopsis thaliana. Role in metabolic reactivation of the herbicide safener fenchlorim. J. Biol. Chem. 283: 21102–21112. Google Scholar

25.

P. Broadbent , G.P. Creissen , B. Kular , A.R. Wellburn , and P.M. Mullineaux (1995). Oxidative stress responses in transgenic tobacco containing altered levels of glutathione reductase activity. Plant J. 8: 247–255. Google Scholar

26.

B.B. Buchanan , and Y. Balmer (2005). Redox regulation: A broadening horizon. Annu. Rev. Plant Biol. 56: 187–220. Google Scholar

27.

P. Buchner , C.E. Stuiver , S. Westerman , M. Wirtz , R. Hell , M.J. Hawkesford , and L.J. De Kok (2004). Regulation of sulfate uptake and expression of sulfate transporter genes in Brassica oleracea as affected by atmospheric H2S and pedospheric sulfate nutrition. Plant Physiol. 136, 3396–3408. Google Scholar

28.

A.D. Butt , and J.B. Ohlrogge (1991). Acyl carrier protein is conjugated to glutathione in spinach seed. Plant Physiol. 96: 937–942. Google Scholar

29.

F. Buwalda , I. Stulen , L. de Kok , and P.J.C. Kuiper (1990) Cysteine, γglutamylcysteine and glutathione contents of spinach leaves as affected by darkness and application of excess sulfur. II. Glutathione accumulation in detached leaves exposed to H2S in the absence of light is stimulated by the supply of glycine to the petiole. Physiol. Plant. 74, 663–668. Google Scholar

30.

O. Cagnac , A. Bourbouloux , D. Chakrabarty , M.Y. Zhang , and S. Delrot (2004). AtOPT6 transports glutathione derivatives and is induced by primisulfuron. Plant Physiol. 135: 1378–1387. Google Scholar

31.

N.G. Cairns , M. Pasternak , A. Wachter , C.S. Cobbett , and A.J. Meyer (2006). Maturation of Arabidopsis seeds is dependent on glutathione biosynthesis within the embryo. Plant Physiol. 141: 446–455. Google Scholar

32.

A.C. Cazalé , and S. Clemens (2001). Arabidopsis thaliana expresses a second functional phytochelatin synthase. FEBS lett. 507: 215–219. Google Scholar

33.

S. Chamnongpol , H. Willekens , W. Moeder , C. Langebartels , H. Sandermann , M. Van Montagu , D. Inze , and W. Van Camp (1998). Defense activation and enhanced pathogen tolerance induced by H2O2 in transgenic tobacco. Proc. Natl. Acad. Sci. USA 95: 5818–5823. Google Scholar

34.

CC Chang , I Slesak, L Jordá , A Sotnikov , M Melzer , Z Miszalski , P.M. Mullineaux, J.E. Parker , B Karpinska , and S. Karpinski (2009) Arabidopsis chloroplastic glutathione peroxidases play a role in cross talk between photooxidative stress and immune responses. Plant Physiol. 150: 670–683. Google Scholar

35.

S. Chaouch and G. Noctor (2010). Myo-inositol abolishes salicylic aciddependent cell death and pathogen defence responses triggered by peroxisomal H2O2. New Phytol. 188: 711–718 Google Scholar

36.

S. Chaouch , G. Queval , S. Vanderauwera , M. Vandorpe , M. Langlois-Meurinne , F. Van Breusegem , P. Saindrenan , and G. Noctor (2010) Peroxisomal hydrogen peroxide is coupled to biotic defense responses by isochorismate synthase 1 in a daylength-related manner. Plant Physiol. 153: 1692–1705. Google Scholar

37.

Z. Chen , T.E. Young , J. Ling , S.C. Chang & D.R. Gallie (2003). Increasing vitamin C content of plants through enhanced ascorbate recycling. Proc. Natl Acad. Sci. USA 100: 3525–3530. Google Scholar

38.

R. Chen , S. Sun , C. Wang , Y. Li , Y. Liang , F. An , C. Li , H. Dong , X. Yang , J. Zhang , and J. Zuo (2009). The Arabidopsis PARAQUAT RESISTANT2 gene encodes an S-nitrosoglutathione reductase that is a key regulator of cell death. Cell Research 19: 1377–1387. Google Scholar

39.

N.H. Cheng , and K.D. Hirschi (2003) Cloning and characterization of CXIP1, a novel PICOT domain-containing Arabidopsis protein that associates with CAX1. J. Biol. Chem. 278: 6503–6509. Google Scholar

40.

N.H. Cheng , J.Z. Liu , A. Brock , R.S. Nelson , and K.D. Hirschi (2006). AtGRXcp, an Arabidopsis chloroplastic glutaredoxin, is critical for protection against protein oxidative damage. J. Biol. Chem. 281: 26280–26288. Google Scholar

41.

O. Chew , J. Whelan , and A.H. Millar (2003). Molecular definition of the ascorbate-glutathione cycle in Arabidopsis mitochondria reveals dual targeting of antioxidant defenses in plants. J. Biol. Chem. 278: 46869– 46877. Google Scholar

42.

P.M. Chung , R.E. Cappel , H.F. Gilbert (1991). Inhibition of glutathione disulfide reductase by glutathione. Arch. Biochem. Biophys. 288: 48–53. Google Scholar

43.

S. Clemens , and D. Paršoh (2009) Multi-tasking phytochelatin synthases. Plant Sci. 177: 266–271. Google Scholar

44.

C Cobbett , and P. Goldsbrough (2002). Phytochelatins and metallothioneins: roles in heavy metal detoxifaction and homeostasis. Annu. Rev. Plant Biol. 53: 159–182. Google Scholar

45.

C.S. Cobbett , M.J. May , R. Howden , and B. Rolls (1998). The glutathione-deficient, cadmium-sensitive mutant, cad2-1, of Arabidopsis thaliana is deficient in γ-glutamylcysteine synthetase. Plant J. 16: 73–78. Google Scholar

46.

E.J. Collinson , G.L. Wheeler , E.O. Garrido , A.M. Avery , S.V. Avery , and C.M. Grant (2002). The yeast glutaredoxins are active as glutathione peroxidases. J. Biol. Chem. 277: 16712–16717. Google Scholar

47.

J. Couturier , C.S. Koh , M. Zaffagnini , A.M. Winger , J.M. Gualberto , C. Corbier , P. Decottignies , J.P. Jacquot , S.D. Lemaire , C. Didierjean , and N. Rouhier (2009). Structure-function relationship of the chloroplastic glutaredoxin S12 with an atypical WCSYS active site. J. Biol. Chem. 284: 9299–9210. Google Scholar

48.

G. Creissen , H. Reynolds , Y.B. Xue , and P. Mullineaux (1995). Simultaneous targeting of pea glutathione reductase and of a bacterial fusion protein to chloroplasts and mitochondria. Plant J. 8: 167–175. Google Scholar

49.

G. Creissen , J. Firmin , M. Fryer , B. Kular , M. Leyland , H. Reynolds , G. Pastori , F. Wellburn , N. Baker , A. Wellburn , and P. Mullineaux (1999). Elevated glutathione biosynthetic capacity in the chloroplasts of transgenic tobacco paradoxically causes increased oxidative stress. Plant Cell 11: 1277–1291. Google Scholar

50.

J.F. Dat , D. Inzé , and F. Van Breusegem (2001). Catalase-deficient tobacco plants: tools for in planta studies on the role of hydrogen peroxide. Redox Rep. 6: 37–42. Google Scholar

51.

S. Davletova , L. Rizhsky , H. Liang , Z. Shengqiang , D.J. Oliver , J. Coutu , V. Shulaev , K. Schlauch and R. Mittler (2005). Cytosolic ascorbate peroxidase 1 is a central component of the reactive oxygen gene network of Arabidopsis. Plant Cell 17: 268–281. Google Scholar

52.

C. Davoine , O. Falletti , T. Douki , G. Iacazio , N. Ennar , J.L. Montillet , and C. Triantaphylides (2006). Adducts of oxylipin electrophiles to glutathione reflect a 13 specificity of the downstream lipoxygenase pathway in the tobacco hypersensitive response. Plant Physiol. 140: 1484–1493. Google Scholar

53.

J. de Rey-Pailhade (1928). Le philothéion, diastase d'hydrogénation. C. R. Soc. Biol. XCIX: 1700. Google Scholar

54.

S. Deb Roy , M. Saxena , P.S. Bhomkar , M. Pooggin , T. Hohn , and N. Bhalla-Sarin (2008). Generation of marker free salt tolerant transgenic plants of Arabidopsis thaliana using the gly I gene and cre gene under inducible promoters. Plant Cell Tiss. Organ Cult. 95: 1–11. Google Scholar

55.

L.A. del Río , L.M. Sandalio , F.J. Corpas , J.M. Palma , and J.B. Barroso (2006). Reactive oxygen species and reactive nitrogen species in peroxisomes. Production, scavenging, and role in cell signaling. Plant Physiol. 141: 330–335. Google Scholar

56.

A. Delauney , D. Pflieger , M.B. Barrault , J. Vinh , and M.B. Toledano (2002). A thiol peroxidase is an H2O2 receptor and redox-transducer in gene activation. Cell 111: 1–11. Google Scholar

57.

C. Després , C. Chubak , A. Rochon , R. Clark , T. Bethune , D. Desveaux , and P.R. Fobert (2003). The Arabidopsis NPR1 disease resistance protein is a novel cofactor that confers redox regulation of DNA binding activity to the basic domain/leucine zipper transcription factor TGA1. Plant Cell 15: 2181–2191. Google Scholar

58.

T. Destro , D. Prasad , D. Martignago , I.L. Bernet A.R. Trentin, I.K. Renu , M. Ferretti , and A. Masi (2010). Compensatory expression and substrate inducibility of γ-glutamyl transferase GGT2 isoform in Arabidopsis thaliana. J. Exp. Bot. 62: 805–814. Google Scholar

59.

M. Díaz , H. Achkor , E. Titarenko , and M.C. Martínez (2006) The gene encoding glutathione-dependent formaldehyde dehydrogenase/GSNO reductase is responsive to wounding, jasmonic acid and salicylic acid. FEBS lett. 543: 136–139. Google Scholar

60.

P. Diaz-Vivancos , Y.-P. Dong , K. Ziegler , J. Markovic , F. Pallardó , T.K. Pellny , P. Verrier , and C.H. Foyer (2010a). Recruitment of glutathione into the nucleus during cell proliferation adjusts whole cell redox homeostasis in Arabidopsis thaliana and lowers the oxidative defence shield. Plant J. 64: 825–838. Google Scholar

61.

P. Diaz Vivancos , T. Wolff , J. Markovic , F.V. Pallard'O and C.H. Foyer (2010b). A nuclear glutathione cycle within the cell cycle. Biochem. J. 431: 169–178. Google Scholar

62.

J.R. Dinneny , T.A. Long , J.Y. Wang , D. Mace , S. Pointer , C. Barron , S.M. Brady , J. Schiefelbein , and P.N. Benfey (2008). Cell identity mediates the response of Arabidopsis roots to abiotic stress. Science 32: 942–945. Google Scholar

63.

S.H. Ding , Q.T. Lu , Y. Zhang , Z.P. Yang , X.G. Wen , L.X. Zhang , and C.M. Lu (2009). Enhanced sensitivity to oxidative stress in transgenic tobacco plants with decreased glutathione reductase activity leads to a decrease in ascorbate pool and ascorbate redox state. Plant Mol. Biol. 69: 577–592. Google Scholar

64.

D.P. Dixon , and R. Edwards (2010) Glutathione S-transferases: May 08, 2010. The Arabidopsis Book. Rockville, MD: American Society of Plant Biologists. doi: 10.1199:tab.0131  http://www.aspb.org/publications/arabidopsis/ Google Scholar

65.

D.P. Dixon , B.G. Davis , and R. Edwards (2002). Functional divergence in the glutathione transferase superfamily in plants. Identification of two classes with putative functions in redox homeostasis in Arabidopsis thaliana. J. Biol. Chem. 277: 30859–30869. Google Scholar

66.

D.P. Dixon , M. Skipsey , N.M. Grundy , and R. Edwards (2005). Stressinduced protein S-glutathionylation in Arabidopsis. Plant Physiol. 138: 2233–2244. Google Scholar

67.

D.P. Dixon , T. Hawkins , P.J. Hussey , and R. Edwards (2009) Enzyme activities and subcellular localization of members of the Arabidopsis glutathione transferase superfamily. J. Exp. Bot. 60: 1207–1218. Google Scholar

68.

M. Dron , S.D. Clouse , R.A. Dixon , M.A. Lawton , and C.J. Lamb (1988) Glutathione and fungal elicitor regulation of a plant defense gene promoter in electroporated protoplasts. Proc. Natl Acad. Sci. USA 85: 6738–6742. Google Scholar

69.

M. Droux (2004). Sulfur assimilation and the role of sulfur in plant metabolism: a survey. Photosynth. Res. 79, 331–348. Google Scholar

70.

E.A. Edwards , S. Rawsthorne , and P.M. Mullineaux (1990). Subcellular distribution of multiple forms of glutathione reductase in pea (Pisum sativum L.). Planta 180: 278–284. Google Scholar

71.

R. Edwards , J.W. Blount , and R.A. Dixon (1991). Glutathione and elicitation of the phytoalexin response in legume cell cultures. Planta 184: 403–409. Google Scholar

72.

B. Enyedi , P. Várnai , and M. Geiszt (2010). Redox state of the endoplasmic reticulum is controlled by Ero1 L-alpha and intraluminal calcium. Antiox. Redox Signal. 13: 721–729. Google Scholar

73.

Y. Eshdat , D. Holland , Z. Faltin , and G. Ben-Hayyim (1997). Plant glutathione peroxidases. Physiol. Plant. 100: 234–240. Google Scholar

74.

E. Esterbauer , and D. Grill (1978). Seasonal variation of glutathione and glutathione reductase in needles of Picea abies. Plant Physiol. 61: 119– 121. Google Scholar

75.

R.C. Fahey , G.L. Newton , D. Randel , and E.M. Kosower (1981). Analysis of biological thiols: quantitative determination of thiols at the picomole level upon derivatization with monobromobimanes and separation by cation-exchange chromatography. Anal. Biochem. 111: 357–365. Google Scholar

76.

A.H. Fairlamb , P. Blackburn , P. Ulrich , B.T. Chait , and A. Cerami (1985). Trypanothione: a novel bis(glutathionyl)spermidine cofactor for glutathione reductase in trypanosomatids. Science 227: 1485–1487. Google Scholar

77.

A. Feechan , E. Kwon , B.W. Yun , Y. Wang , J.A. Pallas , and G.J. Loake (2005). A central role for S-nitrosothiols in plant disease resistance. Proc. Natl Acad. Sci. USA 102: 8054–8059. Google Scholar

78.

S. Ferrari , J.M. Plotnikova , G. De Lorenzo , and F.M. Ausubel (2003). Arabidopsis local resistance to Botrytis cinerea involves salicylic acid and camalexin and requires EDS and PAD2, but not SID2, EDS5 or PAD4. Plant J. 35: 193–205. Google Scholar

79.

M. Ferretti , T. Destro , S.C.E. Tosatto , N. La Rocca , N. Rascio , and A. Masi (2009). Gamma-glutamyl transferase in the cell wall participates in extracellular glutathione salvage from the root apoplast. New Phytol. 181: 115–126. Google Scholar

80.

C.H. Foyer , and B. Halliwell (1976). The presence of glutathione and glutathione reductase in chloroplasts: a proposed role in ascorbic acid metabolism. Planta 133: 21–25. Google Scholar

81.

C.H. Foyer , and P.M. Mullineaux (1998). The presence of dehydroascorbate and dehydroascorbate reductase in plant tissues. FEBS Lett. 425: 528–529. Google Scholar

82.

C.H. Foyer , and G. Noctor (2003). Redox sensing and signalling associated with reactive oxygen in chloroplasts, peroxisomes and mitochondria. Physiol Plant. 119: 355–364 Google Scholar

83.

C.H. Foyer , and G. Noctor (2005). Redox homeostasis and antioxidant signaling: a metabolic interface between stress perception and physiological responses. Plant Cell 17: 1866–1875. Google Scholar

84.

C.H. Foyer , and G. Noctor (2011). Ascorbate and glutathione: the heart of the redox hub. Plant Physiol. 155: 2–18. Google Scholar

85.

C.H. Foyer , H. Lopez-Delgado , J.F. Dat , I.M. Scott (1997). Hydrogen peroxide- and glutathione-associated mechanisms of acclimatory stress tolerance and signalling. Physiol. Plant 100: 241–254. Google Scholar

86.

C.H. Foyer , F.L. Theodoulou , and S. Delrot (2001) The functions of intercellular and intracellular glutathione transport systems in plants. Trends Plant Sci. 6: 486–492. Google Scholar

87.

C.H. Foyer , M. Lelandais , C. Galap , and K.J. Kunert (1991). Effects of elevated cytosolic glutathione reductase activity on the cellular glutathione pool and photosynthesis in leaves under normal and stress conditions. Plant Physiol. 97: 863–872. Google Scholar

88.

C.H. Foyer , N. Souriau , S. Perret , M. Lelandais , K.J. Kunert , C. Pruvost , and L. Jouanin (1995). Overexpression of glutathione reductase but not glutathione synthetase leads to increases in antioxidant capacity and resistance to photoinhibition in poplar trees. Plant Physiol. 109: 1047–1057. Google Scholar

89.

C.H. Foyer , A.J. Bloom , G. Queval., and G. Noctor (2009). Photorespiratory metabolism: genes, mutants, energetics, and redox signaling. Annu. Rev. Plant Biol. 60: 455–484. Google Scholar

90.

P. Frendo , MJ.H. Jiménez , C. Mathieu , L. Duret , D. Gallesi , G. Van de Sype , D. Hérouart , and A. Puppo (1999). A Medicago trunculata homoglutathione synthetase is derived from glutathione synthetase by gene duplication. Plant Physiol. 126: 1706–1715. Google Scholar

91.

P. Frendo , J. Harrison , C. Norman , M.J. Hernandez-Jimenez, G. Van de Sype , A. Gilabert , and A. Puppo (2005). Glutathione and homoglutathione play a critical role in the nodulation process of Medicago truncatula. Mol. Plant Microb. Int. 18: 254–259. Google Scholar

92.

F. Frottin , C. Espagne , J.A. Traverso , C. Mauve , B. Valot , C. Lelarge-Trouverie , M. Zivy , G. Noctor , T. Meinnel , and C. Giglione (2009). Cotranslational proteolysis dominates glutathione homeostasis for proper growth and development. Plant Cell 21: 3296–3314. Google Scholar

93.

X.H. Gao , M. Bedhomme , D. Veyel , M. Zaffagnini , and S.D. Lemaire (2009a). Methods for analysis of protein glutathionylation and their application to photosynthetic organisms. Mol. Plant 2: 218–235. Google Scholar

94.

X.H. Gao , M. Bedhomme , M. Michelet , M. Zaffagnini , and S.D. Lemaire (2009b). Glutathionylation in photosynthetic organisms. Adv. Bot. Res. 52: 363–403. Google Scholar

95.

X.H. Gao , M. Zaffagnini , M. Bedhomme , L. Michelet , C. Cassier-Chauvat , P. Decottignies , S.D. Lemaire (2010). Biochemical characterization of glutaredoxins from Chlamydomonas reinhardtii: Kinetics and specificity in deglutathionylation reactions. FEBS Lett. 584: 2242–2248. Google Scholar

96.

E. Gelhaye , N. Rouhier , J. Gérard , Y. Jolivet , J. Gualberto , N. Navrot , I. Ohlsson , G. Wingsle , M. Hirasawa , D.B. Knaff , H. Wang , P. Dizengremel , Y. Meyer , and J.P. Jacquot (2005) A specific form of thioredoxin h occurs in plant mitochondria and regulates the alternative oxidase. Proc. Natl Acad. Sci. USA 101: 14545–14550. Google Scholar

97.

F. Geu-Flores , M.T. Nielsen , M. Nafisi , M.E. Møldrup , C.E. Olsen , M.S. Motawia , and B.A. Halkier (2009) Glucosinolate engineering identifies a γ-glutamyl peptidase. Nat. Chem. Biol. 5: 575–577. Google Scholar

98.

P. Giegé , J.L. Heazlewood , U. Roessner-Tunali , A.H. Millar , A.R. Fernie , C.J. Leaver , and L.J. Sweetlove (2003) Enzymes of glycolysis are functionally associated with the mitochondrion in Arabidopsis cells. Plant Cell 15: 2140–2151. Google Scholar

99.

E. Glawischnig (2007). Camalexin. Phytochemistry 68: 401–406. Google Scholar

100.

J. Glazebrook , and F.M. Ausubel (1994). Isolation of phytoalexin-deficient mutants of Arabidopsis thaliana and characterization of their interactions with bacterial pathogens. Proc. Natl Acad. Sci. USA 91: 8955– 8959. Google Scholar

101.

L. Gomez , H. Vanacker , P. Buchner , G. Noctor , and C.H. Foyer (2004a) The intercellular distribution of glutathione synthesis and its response to chilling in maize. Plant Physiol. 134: 1662–1671. Google Scholar

102.

L.D. Gomez , G. Noctor , M. Knight , and C.H. Foyer (2004b). Regulation of calcium signaling and gene expression by glutathione. J. Exp. Bot. 55: 1851–1859. Google Scholar

103.

D. Greetham , J. Vickerstaff , D. Shenton , G.G. Perrone , I.A. Dawes , and C.M. Grant (2010). Thioredoxins function as deglutathionylase enzymes in the yeast Saccharomyces cerevisiae. BMC Biochem. 11:3. Google Scholar

104.

E. Grill , E.L. Winnacker , and M.H. Zenk (1987). Phytochelatins, a class of heavy-metal-binding peptides from plants, are functionally analogous to metallothioneins. Proc. Natl. Acad. Sci. USA 84: 439–443. Google Scholar

105.

E. Grill , LöfflerS. , E.L. Winnacker , and M.H. Zenk (1989) Phytochelatins, the heavy-metal-binding peptides of plants, are synthesized from glutathione by a specific γ-glutamylcysteine dipeptidyl transferase (phytochelatin synthase). Proc. Natl. Acad. Sci. USA 86: 6838–6842. Google Scholar

106.

O.W. Griffith (1980). Determination of glutathione and glutathione disulfide using glutathione reductase and 2-vinylpyhdine. Anal. Biochem. 106: 207–212. Google Scholar

107.

O.W. Griffith , and A. Meister (1979) Potent and specific inhibition of glutathione synthesis by buthionine sulfoximine (S-n-butyl homocysteine sulfoximine). J. Biol. Chem. 254: 7558–7560. Google Scholar

108.

R. Gromes , M. Hothorn , E.D. Lenherr , V. Rybin , K. Sheffzek , and T. Rausch (2008) The redox switch of γ-glutamylcysteine ligase via a reversible monomer-dimer transition is a mechanism unique to plants. Plant J. 54: 1063–1075. Google Scholar

109.

A. Grzam , P. Tennstedt , S. Clemens , R. Hell , and A.J. Meyer (2006). Vacuolar sequestration of glutathione S-conjugates outcompetes a possible degradation of the glutathione moiety by phytochelatin synthase. FEBS Lett. 580: 6384–6390. Google Scholar

110.

A. Grzam , M.N. Martin , R. Hell , and A.J. Meyer (2007). γ-Glutamyl transpeptidase GGT4 initiates vacuolar degradation of glutathione S-conjugates in Arabidopsis. FEBS Lett. 581: 3131–3138. Google Scholar

111.

G. Gullner , I. Tóbiás , J. Fodor , T. Kömives (1999). Elevation of glutathione level of glutathione-related enzymes affect virus infection in tobacco. Free Rad. Res. 31: 155–161. Google Scholar

112.

G. Gullner , T. Komives , and H. Rennenberg (2001). Enhanced tolerance of transgenic poplar plants overexpressing γ-glutamylcysteine synthetase towards chloroacetanilide herbicides. J. Exp. Bot. 52: 971–979. Google Scholar

113.

Y. Guo , C. Huang , Y. Xie , F. Song , and X. Zhou (2010). A tomato glutaredoxin gene SIGRX1 regulates plant responses to oxidative, drought and salt stress. Planta 232: 1499–1509. Google Scholar

114.

S.K. Ha , A.P. Smith , R. Howden , W.M. Dietrich , S. Bugg , M.J. O'Connell , P.B. Goldsbrough , and C.S. Cobbett (1999). Phytochelatin synthase genes from Arabidopsis and Schizosaccharomyces pombe. Plant Cell 11: 1153–1163. Google Scholar

115.

F.H. Haas , C. Heeg , R. Queiroz , A. Bauer , M. Wirtz , and R. Hell (2008) Mitochondrial serine acetyltransferase functions as pacemaker of cysteine synthesis in plant cells. Plant Physiol. 148: 1055–1067. Google Scholar

116.

B. Halliwell , and C.H. Foyer (1978). Properties and physiological function of a glutathione reductase purified from spinach leaves by affinity chromatography. Planta 139: 9–17. Google Scholar

117.

Y. Han , J. Zhang , X. Chen , Z. Gao , W. Xuan , S. Xu , X. Ding , and W. Shen (2007) Carbon monoxide alleviates cadmium-induced oxidative damage by modulating glutathione metabolism in the roots of Medicago sativa. New Phytol. 177: 155–166. Google Scholar

118.

J.T. Hancock , R. Desikan , J. Harrison , J. Bright , R. Hooley , and S.J. Neill (2006). Doing the unexpected: proteins involved in hydrogen peroxide perception. J. Exp. Bot. 57: 1711–1718. Google Scholar

119.

G.T. Hanson , R. Aggeler , D. Oglesbee , M. Cannon , R.A. Capaldi , R.Y. Tsien , and S.J. Remington (2004) Investigating mitochondrial redox potential with redox-sensitive green fluorescent protein. J. Biol. Chem. 279: 13044–13053. Google Scholar

120.

K. Harms , P. Von Ballmoos , C. Brunold , R. Höfgen , and H. Hesse (2000). Expression of a bacterial serine acetyltransferase in transgenic potato plants leads to increased levels of cysteine and glutathione. Plant J. 22: 335–343. Google Scholar

121.

T.N. Hartmann , M.D. Fricker , H. Rennenberg , and A.J. Meyer (2003) Cell-specific measurements of cytosolic glutathione in poplar leaves. Plant Cell Environ. 26: 965–975. Google Scholar

122.

R. Haslam , S. Rust , K. Pallett , D. Cole and J. Coleman (2002). Cloning and characterisation of S-formylglutathione hydrolase from Arabidopsis thaliana: a pathway for formaldehyde detoxification. Plant Physiol. Biochem. 40: 281–288. Google Scholar

123.

R. Hell , and L. Bergmann (1990) γ-Glutamylcysteine synthetase in higher plants: catalytic properties and subcellular localization. Planta. 180: 603–612 Google Scholar

124.

S. Herbette , C. Lenne , N. Leblanc , J.L. Julien , J.R. Drevet , and P. Roeckel-Drevet (2002). Two GPX-like proteins from Lycopersicon esculentum and Helianthus annuus are antioxidant enzymes with phospholipid hydroperoxide glutathione peroxidase and thioredoxin peroxidase activities. Eur. J. Biochem. 269: 2414–2420. Google Scholar

125.

C. Herschbach , and H. Rennenberg (1994). Influence of glutathione (GSH) on net uptake of sulfate and sulfate transport in tobacco plants. J. Exp. Bot. 45, 1069–1076. Google Scholar

126.

C. Herschbach , and H. Rennenberg (1995). Long-distance transport of 35S-sulphur in 3-year-old beech trees (Fagus sylvatica). Physiol. Plant. 95: 379–386. Google Scholar

127.

C. Herschbach , L. Rizzini , S. Mult , T. Hartmann , F. Busch , A.D. Peuke , S. Kopriva , and I. Ensminger (2009). Over-expression of bacterial γ-glutamylcysteine synthetase (GSH1) in plastids affects photosynthesis, growth and sulphur metabolism in poplar (Populus tremula × Populus alba) dependent on the resulting γ-glutamylcysteine and glutathione levels. Plant Cell Environ. 33: 1138–1151. Google Scholar

128.

L.M. Hicks , R.E. Cahoon , E.R. Bonner , R.S. Rivard , J. Sheffield , and J.M. Jez (2007). Thiol-based regulation of redox-active glutamate-cysteine ligase from Arabidopsis thaliana. Plant Cell 19: 2653–2661. Google Scholar

129.

K. Höller , L. Király , A. Künstler , M. Müller , G. Gullner , M. Fattinger , B. andZechmann (2010). Enhanced glutathione metabolism is correlated with sulfur-inducedresistance in tobacco mosiac virus-infected genetically susceptible Nicotiana tabacum plants. Mol. Plant Microb. Int. 23:1448–1459. Google Scholar

130.

S. Holtgrefe , J. Gohlke , J. Starmann , S. Druce , S. Klocke , B. Altmann , J. Wojtera , C. Lindermayr , and R. Scheibe (2008). Regulation of plant cytosolic glyceraldehyde 3-phosphate dehydrogenase isoforms by thiol modifications. Physiol. Plant. 133: 211–218. Google Scholar

131.

F.G. Hopkins (1921) On an autooxidisable component of the cell. Biochem. J. 15: 286–303. Google Scholar

132.

F.G. Hopkins (1929) On glutathione: a reinvestigation. J. Biol. Chem. 84: 269–320. Google Scholar

133.

F.G. Hopkins , and E.J. Morgan (1936) Some relations between ascorbic acid and glutathione. Biochem. J. 30: 1446–1462. Google Scholar

134.

M. Hothorn , A. Wachter , R. Gromes , T. Stuwe , T. Rausch , and K. Scheffzek (2006). Structural basis for the redox control of plant glutamate cysteine ligase. J. Biol. Chem. 281: 27557–27565. Google Scholar

135.

T.P. Howard , M. Metodiev , J.C. Lloyd , and C.A. Raines (2008). Thioredoxin-mediated reversible dissociation of a stromal multiprotein complex in response to changes in light availability. Proc. Natl Acad. Sci. USA 105: 4056–4061. Google Scholar

136.

R. Howden , and C.S. Cobbett (1992). Cadmium-sensitive mutants of Arabidopsis thaliana. Plant Physiol. 99: 100–107. Google Scholar

137.

R. Howden , C.R. Anderson , P.B. Goldsbrough and C.S. Cobbett (1995). A cadmium-sensitive, glutathione-deficient mutant of Arabidopsis thaliana. Plant Physiol. 107: 1067–1073. Google Scholar

138.

T. Hruz , O. Laule , G. Szabo , F. Wessendorp , S. Bleuler , L. Oertle , P. Widmayer , W. Gruissem , and P. Zimmermann (2008). Genevestigator V3: a reference expression database for the meta-analysis of transcriptomes. Advances in Bioinformatics vol. 2008, Article ID 420747, 5 pages. doi:10.1155/2008/420747 Google Scholar

139.

V. Hugouvieux , C. Dutilleul , A. Jourdain , F. Reynaud , V. Lopez , and J. Bourguignon (2009). Arabidopsis putative selenium-binding protein 1 expression is tightly linked to cellular sulfur demand and can reduce sensitivity to stresses requiring glutathione for tolerance. Plant Physiol. 151: 768–781. Google Scholar

140.

G. Hunter , and B.A. Eagles (1927) Non-protein sulfur compounds in blood. II. Glutathione. J. Biol. Chem. 72: 133–146. Google Scholar

141.

C. Hwang , A.J. Sinskey , and H.F. Lodish (1992). Oxidized redox state of glutathione in the endoplasmic reticulum. Science 257: 1496–1502. Google Scholar

142.

A.U. Igamberdlev , and P. Gardeström (2003). Regulation of NAD- and NADP-dependent isocitrate dehydrogenases by reduction levels of pyridine nucleotides in mitochondria and cytosol of pea leaves. Biochim. Biophys. Acta 1606: 117–125. Google Scholar

143.

A. Iqbal , Y. Yabuta , T. Takeda , Y. Nakano , and S. Shigeoka (2006). Hydroperoxide reduction by thioredoxin-specific glutathione peroxidase isoenzymes of Arabidopsis thaliana. FEBS J. 273: 5589–5597. Google Scholar

144.

H. Ito , M. Iwabuchi and K. Ogawa (2003). The sugar-metabolic enzymes aldolase and triose-phosphate isomerase are targets of glutathionylation in Arabidopsis thaliana: detection using biotinylated glutathione. Plant Cell Physiol. 44: 655–660. Google Scholar

145.

A. Jamaï , R. Tommasini , E. Martinoia , and S. Delrot (1996). Characterization of glutathione uptake in broad bean leaf protoplasts. Plant Physiol. 111: 1145–1152. Google Scholar

146.

J.M. Jez , R.E. Cahoon , and S. Chen (2004) Arabidopsis thaliana glutamate-cysteine ligase. Functional properties, kinetic mechanism, and regulation of activity. J. Biol. Chem. 279, 33463–33470. Google Scholar

147.

K. Jiang , C. Schwarzer , E. Lally , S.B. Zhang , S. Ruzin , T. Machen , S.J. Remington , and L. Feldman (2006a). Expression and characterization of a redox-sensing green fluorescent protein (reduction-oxidation-sensitive green fluorescent protein) in Arabidopsis. Plant Physiol. 141: 397–403. Google Scholar

148.

K. Jiang , T. Ballinger , D. Li , S. Zhang , and L. Feldman (2006b) A role for mitochondria in the establishment and maintenance of the maize root quiescent center. Plant Physiol. 140 : 1118–1125. Google Scholar

149.

Y.P. Jiang , Y.H. Zhou. , J.X. Cui , X.J. Xia , W.H. Mao , K. Shi , T. Asami. , Z. Chen , and J.Q. Yu (2010) Brassinosteroids increase CO2 assimilation through redox changes in Cucumis sativus. Manuscript submitted Google Scholar

150.

A. Jiménez , J.A. Hernández , L. del Río , and F. Sevilla (1997). Evidence for the presence of the ascorbate-glutathione cycle in mitochondria and peroxisomes of pea leaves. Plant Physiol. 114: 275–284. Google Scholar

151.

T Jubany-Mari , L Alegre-Batlle , K Jiang , LJ. Feldman (2010). Use of a redox-sensing GFP (c-roGFP1) for real-time monitoring of cytosol redox status in Arabidopsis thaliana water-stressed plants. FEBS lett. 584: 889–897. Google Scholar

152.

S.M. Kanzok , A. Fechner , H. Bauer , J.K. Ulschmid , H.M. Müller , J. Botella-Munoz , S. Schneuwly , R. Schirmer , and K. Becker (2001). Substitution of the thioredoxin system for glutathione reductase in Drosophila melanogaster. Science 291: 643–646. Google Scholar

153.

A.M.R. Kataya , and S. Reumann (2010) Arabidopsis glutathione reductase 1 is dually targeted to peroxisomes and the cytosol. Plant Signal. Behav. 5: 171–175. Google Scholar

154.

N. Kaur , S. Reumann , and J. Hu (2009). Peroxisome biogenesis and function: September 11, 2009. The Arabidopsis Book. Rockville, MD: American Society of Plant Biologists. doi: 10.1199:tab.0123  http://www.aspb.org/publications/arabidopsis/ Google Scholar

155.

C.G. Kawashima , O. Berkowitz , R. Hell , M. Noji , and K. Saito (2005). Characterization and expression analysis of a serine acetyltransferase gene family involved in a key step of the sulfur assimilation pathway in Arabidopsis. Plant Physiol. 137: 220–230. Google Scholar

156.

N. Kerk , and L. Feldman (1995). A biochemical model for the initiation and maintenance of the quiescent center: implications for organization of root meristems. Development 121: 2825–2833. Google Scholar

157.

S. Klapheck (1988). Homoglutathione: isolation, quantification and occurrence in legumes. Physiol. Plant. 74: 727–732. Google Scholar

158.

S. Klapheck , B. Chrost , J. Starke , and H. Zimmermann (1992) γ-Glutamylcysteinylserine — a new homologue of glutathione in plants of the family Poaceae. Bot. Acta 105: 174–179. Google Scholar

159.

G. Kocsy , S. von Ballmoos , M. Suter , A. Ruegsegger , U. Galli , G. Szalai , G. Galiba , and C. Brunold (2000). Inhibition of glutathione synthesis reduces chilling tolerance in maize. Planta 211: 528–536. Google Scholar

160.

S. Koh , A.M. Wiles , J.S. Sharp , F.R. Naider , J.M. Becker , and G. Stacey (2002). An oligopeptide transporter gene family in Arabidopsis. Plant Physiol. 128: 21–29. Google Scholar

161.

D. Kolb , M. Müller , G. Zellnig , and B. Zechmann (2010). Cadmiuminduced changes in subcellular glutathione contents within glandular trichomes of Cucurbita pepo L. Protoplasma 243: 87–94. Google Scholar

162.

A. Koornneef , A. Leon-Reyes , T. Ritsema , A. Verhage , F.C. Den Otter , L.C. Van Loon , and C.M.J. Pieterse (2008). Kinetics of salicylate-mediated suppression of jasmonate signaling reveal a role for redox modulation. Plant Physiol. 147: 1358–1368. Google Scholar

163.

S Kopriva , and H. Rennenberg (2004) Control of sulphate assimilation and glutathione synthesis: interaction with N and C metabolism. J. Exp. Bot. 55: 1831–1842. Google Scholar

164.

D. Kornyeyev , B.A. Logan , R.D. Allen , and A.S. Holaday (2005). Fieldgrown cotton plants with elevated activity of chloroplastic glutathione reductase exhibit no significant alteration of diurnal and seasonal patterns of excitation energy partitioning and CO2 fixation. Field Crops Res. 94: 165–175. Google Scholar

165.

I. Kranner , and D. Grill (1996) Significance of thiol-disulfide exchange in resting stages of plant development. Bot. Acta 109: 8–14. Google Scholar

166.

I. Kranner , R.P. Beckett , S. Wornik , M. Zorn , and H.W. Pfeifhofer (2002). Revival of a resurrection plant correlates with its antioxidant status. Plant J. 31: 13–24. Google Scholar

167.

I. Kranner , S. Birtic , K.M. Anderson and H.W. Pritchard (2006). Glutathione half-cell reduction potential: A universal stress marker and modulator of programmed cell death. Free Rad. Biol. Med. 40: 2155–2165. Google Scholar

168.

S. Krueger , A. Niehl , M.C.L. Martin , D. Steinhauser , A. Donath , T. Hildebrandt , L.C. Romero , R. Hoefgen , C. Gotor , and H. Hesse (2009). Analysis of cytosolic and plastidic serine acetyltransferase mutants and subcellular metabolite distributions suggests interplay of the cellular compartments for cysteine biosynthesis in Arabidopsis. Plant Cell Environ. 32: 349–367. Google Scholar

169.

S. Kushnir , E. Babiychuk , K. Kampfenkel , E. Belles-Boix , M. Van Montagu , and Inzé D. (1995). Characterization of Arabidopsis thaliana cDNAs that render yeasts tolerant toward the thiol-oxidizing drug diamide. Proc. Natl. Acad. Sci. USA 92: 10580–10584. Google Scholar

170.

C. Laloi , N. Rayapuram , Y. Chartier , J.M. Grienenberger , G. Bonnard , and Y. Meyer (2001). Identification and characterization of a mitochondrial thioredoxin system in plants. Proc. Natl Acad. Sci. USA 98: 14144–14149. Google Scholar

171.

A.G. Lappartient , J.J. Vidmar , T. Leustek , A.D. Glass , and B. Touraine (1999). Inter-organ signaling in plants: regulation of ATP sulfurylase and sulfate transporter genes expression in roots mediated by phloemtranslocated compound. Plant J. 18, 89–95. Google Scholar

172.

S. Lee , D. Petros , J.S. Moon , T.S. Ko , P.B. Golsdsbrough , and S.S. Korban (2003). Higher levels of ectopic expression of Arabidopsis phytochelatin synthase do not lead to increased cadmium tolerance and accumulation. Plant Physiol. Biochem. 41: 903–910. Google Scholar

173.

U. Lee , C. Wie , B.O. Fernandez , M. Feelisch , and E. Vierling (2008) Modulation of nitrosative stress by S-nitrosoglutathione reductase is critical for thermotolerance and plant growth in Arabidopsis. Plant Cell 20: 786–802. Google Scholar

174.

S.D. Lemaire (2004). The glutaredoxin family in oxygenic photosynthetic organisms. Photosynth. Res. 79: 305–318. Google Scholar

175.

S. Lemaire , L. Michelet , M. Zaffagnini , V. Massot , and E. Issakidis-Bourguet (2007). Thioredoxins in chloroplasts. Curr. Genet. 51: 343–365. Google Scholar

176.

T. Leustek (2002). Sulfate metabolism: April 4, 2002. The Arabidopsis Book. Rockville, MD: American Society of Plant Biologists, doi: 10.1199/ tab.0017  http://www.aspb.org/publications/arabidopsis/ Google Scholar

177.

T. Leustek , M.N. Martin , J.A. Bick , and J.P. Davies (2000). Pathways and regulation of sulfur metabolism revealed through molecular and genetic studies. Annu. Rev. Plant Physiol. Plant Mol. Biol. 51: 141–165. Google Scholar

178.

A. Levine , R. Tenhaken , R. Dixon , and C. Lamb (1994). H2O2 from the oxidative burst orchestrates the plant hypersensitive disease resistance response. Cell 79: 583–593. Google Scholar

179.

S. Li , A. Lauri , M. Ziemann , A. Busch , M. Bhave , and S. Zachgo (2009). Nuclear activity of ROXY1, a glutaredoxin interacting with TGA factors, is required for petal development in Arabidopsis thaliana. Plant Cell 21: 429–441. Google Scholar

180.

D Lieberherr ,. U. Wagner , P.H. Dubuis , J.P. Metraux , and F. Mauch (2003). The rapid induction of glutathione S-transferases ATGSTF2 and ATGSTF6 by avirulent Pseudomonas syringae is the result of combined salicylic acid and ethylene signaling. Plant Cell Physiol. 44: 750–757. Google Scholar

181.

V. Liedschulte , A. Wachter , A. Zhigang , and T. Rausch (2010). Exploiting plants for glutathione (GSH) production: Uncoupling GSH synthesis from cellular controls results in unprecedented GSH accumulation. J. Plant Biotechnol. 8: 807–820. Google Scholar

182.

P. Limphong , M.W. Crowder , B. Bennett , and C.A. Makaroff (2009) Arabidopsis thaliana GLX2-1 contains a dinculear metal binding site, and is not a glyoxalase 2. Biochem. J. 417: 323–330. Google Scholar

183.

C. Lindermayr , and J. Durner (2009). S-Nitrosylation in plants: Pattern and function. J. Proteomics 73: 1–9. Google Scholar

184.

C. Lindermayr , G. Saalbach , and J. Dürner (2005). Proteomic identification of S-nitrosylated proteins in Arabidopsis. Plant Physiol. 137: 921– 930. Google Scholar

185.

C. Lindermayr , G. Saalbach , G. Bahnweg , and J. Durner (2006). Differential inhibition of Arabidopsis methionine adenosyltransferases by protein S-nitrosylation. J. Biol. Chem. 281: 4285–4291. Google Scholar

186.

P. Lindroth , and K. Mopper (1979). High performance liquid chromatographic determination of subpicomole amounts of amino acids by precolumn fluorescence derivatization with o-phthaldialdehyde. Anal. Chem. 51: 1667–1674. Google Scholar

187.

Y.P. Lu , Z.S. Li , Y.M. Drozdowicz , S. Hörtensteiner , E. Martinoia , and P.A. Rea (1998). AtMRP2, an Arabidopsis ATP binding cassette transporter able to transport glutathione S-conjugates and chlorophyll catabolites: functional comparisons with AtMRP1 Plant Cell 10: 267–282. Google Scholar

188.

P.K. MacNicol (1987). Homoglutathione and glutathione synthetases of legume seedlings: partial purification and substrate specificity. Plant Sci. 53: 229–235. Google Scholar

189.

M.K. Maiti , S. Krishnasamy , H.A. Owen , and C.A. Makaroff (1997). Molecular cloning of glyoxalase II from a higher plant: comparison of mitochondrial and cytoplasmic isozymes. Plant Mol. Biol. 35: 471–481. Google Scholar

190.

G.P.K. Marasinghe , I.M. Sander , B. Bennett , G. Perlyannan , K.W. Yang , C.A. Makaroff , and M.W. Crowder (2005). Structural studies on a mitochondrial glyoxylase II. J. Biol. Chem. 280: 40668–40675. Google Scholar

191.

J. Markovic , C. Borrás , A. Ortega , J. Sastre , J. Viña , and F.V. Pallardó (2007). Glutathione is recruited into the nucleus in early phases of cell proliferation. J. Biol. Chem. 282: 20416–20424. Google Scholar

192.

M.N. Martin , and J.P. Slovin (2000). Purified γ-glutamyl transpeptidases from tomato exhibit high affinity for glutathione and glutathione S-conjugates. Plant Physiol. 122: 1417–1426. Google Scholar

193.

M.N. Martin , M.C. Tarczynski , B. Shen , and T. Leustek (2005) The role of 5′-adenylylsulfate reductase in controlling sulfate reduction in plants. Photosynth. Res. 86: 309–323. Google Scholar

194.

M.N. Martin , P.H. Saladores , E. Lambert , A.O. Hudson , and T. Leustek (2007). Localization of members of the gamma-glutamyl transpeptidase family identifies sites of glutathione and glutathione S-conjugate hydrolysis. Plant Physiol. 144: 1715–1732. Google Scholar

195.

R.E. Martin , R.V. Marchetti , A.I. Cowan , S.M. Howitt , S. Bröer , and K. Kirk (2009). Chloroquinone transport via the malaria parasite's chloroquine resistance transporter. Science 325: 1680–1682. Google Scholar

196.

M.C. Martínez , H. Achkor, B. Persson , M.R. Fernández , J. Shafqat , J. Farrés , H. Jörnvall and X. Parés (1996). Arabidopsis formaldehyde dehydrogenase. Molecular properties of plant class III alcohol dehydrogenase provide further insights into the origins, structure and function of plant class p and liver class I alcohol dehydrogenases. Eur. J. Biochem. 241: 849–857 Google Scholar

197.

E. Martinoia , E. Grill , R. Tommasini , K. Kreuz , and N. Amrhein (1993). ATP-dependent glutathione S-conjugate ‘export’ pump in the vacuolar membrane of plants. Nature 364: 247–249. Google Scholar

198.

L. Marty , W. Siala , M. Schwarzländer , M.D. Fricker , M. Wirtz , L.J. Sweetlove , Y. Meyer , A.J. Meyer , J.P. Reichheld , and R. Hell (2009). The NADPH-dependent thioredoxin system constitutes a functional backup for cytosolic glutathione reductase in Arabidopsis. Proc. Natl. Acad. Sci. USA 106: 9109–9114. Google Scholar

199.

A. Maruyama-Nakashita , E. Inoue , A. Watanabe-Takahashi , T. Yamaya , and H. Takahashi (2003). Transcriptome profiling of sulfur-responsive genes in Arabidopsis reveals global effects of sulfur nutrition on multiple metabolic pathways. Plant Physiol. 132, 597–605. Google Scholar

200.

A. Masi , T. Destro , L. Turetta , S. Varotto , G. Caporale , and M. Ferretti (2007). Localization of gamma-glutamyl transferase activity and protein in Zea mays organs and tissues. J. Plant Physiol. 164: 1527–1535. Google Scholar

201.

A. Mateo , D. Funck , P. Mühlenbock , B. Kular , P.M. Mullineaux , and S. Karpinski (2006). Controlled levels of salicylic acid are required for optimal photosynthesis and redox homeostasis. J. Exp. Bot. 57: 1795– 1807. Google Scholar

202.

S.C. Maughan , M. Pasternak , N. Cairns , G. Kiddle , T. Brach , R. Jarvis , F. Haas , J. Nieuwland , B. Lim , C. Müller , E. Salcedo-Sora , C. Kruse , M. Orsel , R. Hell , A.J. Miller , P. Bray , C.H. Foyer , J.A.H. Murray , A.J. Meyer and C.S. Cobbett (2010) Plant homologs of the Plasmodium falciparum chloroquinone-resistance transporter, PfCRT, are required for glutathione homeostasis and stress responses. Proc. Natl. Acad. Sci. USA 107: 2331–2336. Google Scholar

203.

G. Maulucci , G. Pani , V. Labate , M. Mele , E. Panieri , M. Papi , G. Arcovito , T. Galeotti , and M. De Spirito (2009). Investigation of the spatial distribution of glutathione redox-balance in live cells by using fluorescence ratio imaging microscopy. Biosens. Bioelectron. 25: 682–687. Google Scholar

204.

M.J. May , and C.J. Leaver (1993). Oxidative stimulation of glutathione synthesis in Arabidopsis thaliana suspension cultures. Plant Physiol. 103: 621–627. Google Scholar

205.

M.J. May , and C.J. Leaver (1994). Arabidopsis thaliana γ-glutamylcysteine synthetase is structurally unrelated to mammalian, yeast and Escherichia coli homologs. Proc. Natl. Acad. Sci. USA 91: 10059–10063. Google Scholar

206.

M.J. May , K.E. Hammond-Kosack , and J.D.G. Jones (1996a). Involvement of reactive oxygen species, glutathione metabolism and lipid peroxidation in the Cf-gene-dependent defence response of tomato cotyledons induced by race-specific elicitors of Cladosporium fulvum. Plant Physiol. 110: 1367–1379. Google Scholar

207.

M.J. May , J.E. Parker , M.J. Daniels , C.J. Leaver , and C.S. Cobbett (1996b) An Arabidopsis mutant depleted in glutathione shows unaltered responses to fungal and bacterial pathogens. Mol. Plant Microb. Int. 9: 349–356 Google Scholar

208.

M.J. May , T. Vernoux , C. Leaver , M. Van Montagu , and D. Inzé (1998a). Glutathione homeostasis in plants: implications for environmental sensing and plant development. J. Exp. Bot. 49: 649–667. Google Scholar

209.

M.J. May , T. Vernoux , R. Sanchez-Fernandez , M. Van Montagu , and D. Inzé (1998b). Evidence for post-transcriptional activation of γ-glutamylcysteine synthetase during plant stress responses. Proc. Natl. Acad. Sci. USA 95: 12049–12054. Google Scholar

210.

A. Meister (1988a). On the discovery of glutathione. Trends Biochem. Sci 13: 185–188. Google Scholar

211.

A. Meister (1988b). Glutathione metabolism and its selective modification. J. Biol. Chem. 263: 17205–17208. Google Scholar

212.

J. Mendoza , T. Garrido , R. Riveros , C. Gonzalez , and J. Parada (2008). Simultaneous determination of glutathione and glutathione disulfide in an acid extract of plant shoot and root by capillary electrophoresis. J. Chil. Chem. Soc. 53: 1626–1630. Google Scholar

213.

D.G. Mendoza-Cózatl , E. Butko , F. Springer , J.W. Torpey , E.A. Komives , J. Kehr , and J.I. Schroeder (2008). Identification of high levels of phytochelatins, glutathione and cadmium in the phloem sap of Brassica napus. A role for thiol-peptides in the long-distance transport of cadmium and the effect of cadmium on iron translocation. Plant J. 54: 249–259. Google Scholar

214.

P. Meuwly , P. Thibault , and W.E. Rauser (1993) γ-Glutamylcysteinylglutamic acid — a new homologue of glutathione in maize seedlings exposed to cadmium. FEBS lett. 336: 472–476. Google Scholar

215.

A.J. Meyer (2008). The integration of glutathione homeostasis and redox signaling. J. Plant Physiol. 165: 1390–1403. Google Scholar

216.

A.J. Meyer , and M. Fricker (2002). Control of demand-driven biosynthesis of glutathione in green Arabidopsis suspension culture cells. Plant Physiol. 130:1927–1937. Google Scholar

217.

A.J. Meyer , M.J. May , and M. Fricker (2001) Quantitative in vivo measurement of glutathione in Arabidopsis cells. Plant J. 27: 67–78. Google Scholar

218.

A.J. Meyer , T. Brach , L. Marty , S. Kreye , N. Rouhier , J.P. Jacquot , and R. Hell (2007). Redox-sensitive GFP in Arabidopsis thaliana is a quantitative biosensor for the redox potential of the cellular glutathione redox buffer. Plant J. 52: 973–986. Google Scholar

219.

Y. Meyer , W. Siala , T. Bashandy , C. Riondet , F. Vignols , and J.P. Reichheld (2008). Glutaredoxins and thioredoxins in plants. Biochim. Biophys. Acta 1783: 589–600. Google Scholar

220.

A. Mhamdi , J. Hager , S. Chaouch , G. Queval , Y. Han , L. Taconnat , P. Saindrenan , H. Gouia , E. Issakidis-Bourguet , J.P. Renou , and G. Noctor (2010a) Arabidopsis GLUTATHIONE REDUCTASE 1 plays a crucial role in leaf responses to intracellular H2O2 and in ensuring appropriate gene expression through both salicylic acid and jasmonic acid signaling pathways. Plant Physiol. 153: 1144–1160. Google Scholar

221.

A. Mhamdi , C. Mauve , H. Gouia , P. Saindrenan , M. Hodges , and G. Noctor (2010b). Cytosolic NADP-dependent isocitrate dehydrogenase contributes to redox homeostasis and the regulation of pathogen responses in Arabidopsis leaves. Plant Cell Environ. 33: 1112–1123. Google Scholar

222.

A. Mhamdi , G. Queval , S. Chaouch , S. Vanderauwera , F. Van Breusegem , and G. Noctor (2010c). Catalase function in plants: a focus on Arabidopsis mutants as stress-mimic models. J. Exp. Bot. 61: 4197–4220 Google Scholar

223.

Y. Miao , D. Lv , P. Wang , X.C. Wang , J. Chen , C. Miao , et al. (2006) An Arabidopsis glutathione peroxidase functions as both a redox transducer and a scavenger in abscisic acid and drought responses. Plant Cell 18: 2749–2766. Google Scholar

224.

L. Michelet , M. Zaffagnini , C. Marchand , V. Collin , P. Decottignies , P. Tsan , J.M. Lancelin , P. Trost , M. Miginiac-Maslow , G. Noctor and S.D. Lemaire (2005). Glutathionylation of chloroplast thioredoxin f is a redox signaling mechanism in plants. Proc. Natl. Acad. Sci. USA 102: 16478–16483. Google Scholar

225.

L. Michelet , M. Zaffagnini , H. Vanacker , P. Le Maréchal , C. Marchand , M. Schroda , S.D. Lemaire , and P. Decottignies (2008) In vivo targets of S-thiolation in Chlamydomonas reinhardtii. J. Biol. Chem. 283: 21571–21578. Google Scholar

226.

R. Mittler , S. Vanderauwera , M. Gollery , and F. Van Breusegem (2004). Reactive oxygen gene network of plants. Trends Plant Sci. 9: 490–498. Google Scholar

227.

F. Montrichard , F. Alkhalfioui , H. Yano , W.H. Vensel , W.J. Hurkman , and B.B. Buchanan (2009). Thioredoxin targets in plants: the first 30 years. J. Proteomics 72: 452–474. Google Scholar

228.

Z. Mou , W. Fan , and X. Dong (2003). Inducers of plant systemic acquired resistance regulate NPR1 function through redox changes. Cell 113: 935–944. Google Scholar

229.

I. Muckenschnabel , B. Williamson , B.A. Goodman , G.D. Lyon , D Stewart , and N. Deighton (2001). Markers for oxidative stress associated with soft rots in French beans (Phaseolus vulgaris) infected by Bortytis cinerea. Planta 212: 376–381. Google Scholar

230.

S. Mueller , B. Hilbert , K. Dueckershoff , T. Roitsch , M. Krischke , M.J. Mueller , S. Berger (2008). General detoxification and stress responses are mediated by oxidized lipids through TGA transcription factors in Arabidopsis. Plant Cell 20: 768–785. Google Scholar

231.

M. Müller , M. Tausz , H. Guttenberger , and D. Grill (1999a). Histochemical localization of glutathione in plant tissues: a comparison of two fluorescence staining methods. Phyton 39: 69–73. Google Scholar

232.

M. Müller , M. Tausz , A. Wonisch , and D. Grill (1999b). Effects of an oxidizing agent (hydrogen peroxide) on the glutathione system in epidermal cells of Allium cepa L. investigated by histochemical staining. Free Rad. Res. 31: S121–127. Google Scholar

233.

P.M. Mullineaux , and T. Rausch (2005). Glutathione, photosynthesis and the redox regulation of stress-responsive gene expression. Photosynth. Res. 86: 459–474. Google Scholar

234.

M. Nafisi , S. Goregaoker , C.J. Botanga , E. Glawischnlg , C.E. Olsen , B.A. Halkier , and J. Glazebrook (2007). Arabidopsis cytochrome P450 monooxygenase 71A13 catalyzes the conversion of indole-3-acetaldoxime in camalexin synthesis. Plant Cell 19: 2039–2052. Google Scholar

235.

N. Navrot , V. Collin , J. Gualberto , E. Gelhaye , M. Hirasawa , P. Rey , D.B. Knaff , E. lssakidis , J.P. Jacquot , and N. Rouhier (2006) Plant glutathione peroxidases are functional peroxiredoxins distributed in several subcellular compartments and regulated during biotic and abiotic stress. Plant Physiol. 142: 1364–1379. Google Scholar

236.

I. Ndamukong , A. Al Abdallat , C. Thurow , B. Fode , M. Zander , R. Weigel , and C. Gatz (2007). SA-inducible Arabidopsis glutaredoxin interacts with TGA factors and suppresses JA-responsive PDF1.2 transcription. Plant J. 50: 128–139. Google Scholar

237.

G. Née , M. Zaffagnini , P. Trost and E. Issakldis-Bourguet (2009). Redox regulation of chloroplastic glucose-6-phosphate dehydrogenase: A new role for f-type thioredoxin. FEBS lett. 583: 2827–2832. Google Scholar

238.

G.L. Newton , and B. Javor (1985). γ-Glutamylcysteine and thiosulfate are the major low-molecular-weight thiols in halobacteria. J. Bacteriol. 161: 438–441. Google Scholar

239.

V. Nikiforova , J. Freitag , S. Kempa , Adamik M. , H. Hesse , and R. Hoefgen (2003). Transcriptome analysis of sulfur depletion in Arabidopsis thaliana: interlacing of biosynthetic pathways provides response specificity. Plant J. 33, 633–650. Google Scholar

240.

G. Noctor , and C.H. Foyer (1998) Simultaneous measurement of foliar glutathione, γ-glutamylcysteine and amino acids by high—performance liquid chromatography: comparison with two other assay methods for glutathione. Anal. Biochem. 264: 98–110 Google Scholar

241.

G. Noctor , S. Strohm , L. Jouanin , K-J. Kunert , C.H. Foyer , and H. Rennenberg (1996). Synthesis of glutathione in leaves of transgenic poplar overexpressing γ-glutamylcysteine synthetase. Plant Physiol. 112: 1071–1078. Google Scholar

242.

G. Noctor , A.C.M. Arisi , L. Jouanin , M.H. Valadier , Y. Roux , and C.H. Foyer (1997). The role of glycine in determining the rate of glutathione synthesis in poplars. Possible implications for glutathione production during stress. Physiol. Plant. 100: 255–263. Google Scholar

243.

G. Noctor , A.C.M. Arisi , L. Jouanin , and C.H. Foyer (1998a). Manipulation of glutathione and amino acid biosynthesis in the chloroplast. Plant Physiol. 118: 471–482. Google Scholar

244.

G Noctor , ACM Arisi, L Jouanin , KJ Kunert , H Rennenberg , CH Foyer (1998b) Glutathione: biosynthesis, metabolism and relationship to stress resistance explored in transformed plants. J. Exp. Bot. 49: 623–647 Google Scholar

245.

G. Noctor , S. Veljovic-Jovanovic , and C.H. Foyer (2000). Peroxide processing in photosynthesis: antixoxidant coupling and redox signalling. Proc. Roy. Soc. Lond. B 355: 1465–1475. Google Scholar

246.

G. Noctor , L. Gomez , H. Vanacker , and C.H. Foyer (2002). Interactions between biosynthesis, compartmentation and transport in the control of glutathione homeostasis and signalling. J. Exp. Bot. 53: 1283–1304. Google Scholar

247.

M. Noji , and K. Saito (2002). Molecular and biochemical analysis of serine acetyltransferase and cysteine synthase towards sulfur metabolic engineering in plants. Amino Acids 22: 231–243. Google Scholar

248.

Y. Nyathi , and A. Baker (2006). Plant peroxisomes as a source of signalling molecules. Biochim Biophys Acta 1763: 1478–1495. Google Scholar

249.

K. Ogawa , A. Hatano-lwasaki , M. Yanagida , and M. Iwabuchi (2004) Level of glutathione is regulated by ATP-dependent ligation of glutamate and cysteine through photosynthesis in Arabidopsis thaliana: mechanism of strong interaction of light intensity with flowering. Plant Cell Physiol. 45: 1–8. Google Scholar

250.

N. Ohkamu-Ohtsu , S. Radwan , A. Peterson , P. Zhao , A.F. Badr , C. Xiang , and D.J. Oliver (2007a). Characterization of the extracellular γ-glutamyl transpeptidases, GGT1 and GGT2, in Arabidopsis. Plant J. 49: 865–877. Google Scholar

251.

N. Ohkamu-Ohtsu , P. Zhao , C. Xiang , and D.J. Oliver (2007b). Glutathione conjugates in the vacuole are degraded by γ-glutamyl transpeptidase GGT3 in Arabidopsis. Plant J. 49: 878–888. Google Scholar

252.

N. Ohkamu-Ohtsu , A. Oikawa , P. Zhao , C. Xiang , K. Saito , and D.J. Oliver (2008) A γ-glutamyl transpeptidase-independent pathway of glutathione catabolism to glutamate via 5-oxoproline in Arabidopsis. Plant Physiol. 148: 1603–1613. Google Scholar

253.

R. Okumura , Y. Koizumi , and J. Sekiya (2003). Synthesis of hydroxymethylglutathione from glutathione and L-serine catalyzed by carboxy-peptidase Y. Biosci. Biotechnol. Biochem. 67: 434–437. Google Scholar

254.

M. Oven , K. Raith , R.H.H. Neubert , T.M. Kutchan , and M.H. Zenk (2001). Homo-phytochelatins are synthesized in response to cadmium in Azuki beans. Plant Physiol. 126: 1275–1280. Google Scholar

255.

M.C. Palmieri , C. Lindermayr , H. Bauwe , C. Steinhauser , and J. Durner (2010). Regulation of plant glycine decarboxylase by S-nitrosylation and glutathionylation. Plant Physiol. 152: 1514–1528. Google Scholar

256.

V. Parisy , B. Poinssot , L. Owsianowski , A. Buchala , J. Glazebrook and F. Mauch (2006). Identification of PAD2 as a γ-glutamylcysteine synthetase highlights the importance of glutathione in disease resistance in Arabidopsis. Plant J. 49: 159–172. Google Scholar

257.

M. Pasternak , B. Lim , M. Wirtz , R. Hell , C.S. Cobbett , and A.J. Meyer (2008). Restricting glutathione biosynthesis to the cytosol is sufficient for normal plant development. Plant J. 53: 999–1012. Google Scholar

258.

N. Pauly , C. Pucciariello , K. Mandon , G. Innocenti , A. Jamet , E. Baudouin , D. Hérouart , P. Frendo , and A. Puppo (2006). Reactive oxygen and nitrogen species and glutathione: Key players in the legumerhizobium symbiosis. J. Exp. Bot. 57: 1769–1776. Google Scholar

259.

C. Petershansel , I. Horst , M. Niessen , C. Blume , R. Kebeish , S. Kürkcüoglu , and F. Kreuzaler (2010). Photorespiration: March 23, 2010. The Arabidopsis Book. Rockville, MD: American Society of Plant Biologists, doi: 10.1199/tab.0130  http://www.aspb.org/publications/arabidopsis/ Google Scholar

260.

A. Peterson , and D.J. Oliver (2006). Leaf-targeted phytochelatin synthase in Arabidopsis thaliana. Plant Physiol. Biochem. 44: 885–892. Google Scholar

261.

N. Picault , Cazalé A.C. , A. Beyly , S. Cuiné , P. Carrier , D.T. Luu , C. Forestier , and G. Peltier (2006). Chloroplast targeting of phytochelatin synthase in Arabidopsis: effects on heavy metal tolerance and accumulation. Biochimie 88: 1743–1750. Google Scholar

262.

S. Pike , A. Patel , G. Stacey , and W. Gassmann (2009). Arabidopsis OPT6 is an oligopeptide transporter with exceptionally broad substrate specificity. Plant Cell Physiol. 50: 1923–1932. Google Scholar

263.

G. Potters , L. De Gara , H. Asard , and N. Horemans (2002). Ascorbate and glutathione: guardians of the cell cycle, partners in crime? Plant Physiol. Biochem. 40: 537–548. Google Scholar

264.

G. Potters , N. Horemans , S. Bellone , R.J. Caubergs , P. Trost , Y. Guisez and H. Asard (2004). Dehydroascorbate influences the plant cell cycle through a glutathione-independent reduction mechanism. Plant Physiol. 134: 1479–1487. Google Scholar

265.

G. Queval and G. Noctor (2007). A plate-reader method for the measurement of NAD, NADP, glutathione and ascorbate in tissue extracts. Application to redox profiling during Arabidopsis rosette development. Anal. Biochem. 363: 58–69. Google Scholar

266.

G. Queval , E. Issakidis-Bourguet , F.A. Hoeberichts , M. Vandorpe , B. Gakière , H. Vanacker , M. Miginiac-Maslow , F. Van Breusegem , and G. Noctor (2007). Conditional oxidative stress responses in the Arabidopsis photorespiratory mutant cat2 demonstrate that redox state is a key modulator of daylength-dependent gene expression and define photoperiod as a crucial factor in the regulation of H2O2-induced cell death. Plant J. 52: 640–657. Google Scholar

267.

G. Queval , D. Thominet , H. Vanacker , M. Miginiac-Maslow , B. Gakière , and G. Noctor (2009). H2O2-activated up-regulation of glutathione in Arabidopsis involves induction of genes encoding enzymes involved in cysteine synthesis in the chloroplast. Mol. Plant 2: 344–356. Google Scholar

268.

G. Queval , D. Jaillard , B. Zechmann , and G. Noctor (2011) Increased intracellular H2O2 availability preferentially drives glutathione accumulation in vacuoles and chloroplasts. Plant Cell Environ. 34: 21–32. Google Scholar

269.

A.G. Rasmusson , and I.M. Møller (1990) NADP-utilizing enzymes in the matrix of plant mitochondria. Plant Physiol. 94: 1012–1018. Google Scholar

270.

W.E. Rauser , R. Schupp , and H. Rennenberg (1991) Cysteine, γ-glutamylcysteine and glutathione levels in maize seedlings. Plant Physiol. 97: 128–134. Google Scholar

271.

M.R. Rawlins , C.J. Leaver , and M.J. May (1995). Characterisation of an Arabidopsis thaliana cDNA encoding glutathione synthetase. FEBS lett. 376: 81–86. Google Scholar

272.

P.A. Rea (1999). MRP subfamily transporters from plants and yeast. J. Exp. Bot. 50: 895–913. Google Scholar

273.

P.A. Rea , O.K. Vatamaniuk , and D.J. Rigden (2004). Weeds, worms, and more. Papain's long-lost cousin, phytochelatin synthase. Plant Physiol. 136: 2463–2474. Google Scholar

274.

JP Reichheld , T Vernoux, F Lardon , M Van Montagu , D Inzé (1999) Specific checkpoints regulate plant cell cycle progression in response to oxidative stress. Plant J. 17: 647–656. Google Scholar

275.

J.P. Reichheld , M. Khafif , C. Riondet , M. Droux , G. Bonnard , and Y. Meyer (2007). Inactivation of thioredoxin reductases reveals a complex interplay between thioredoxin and glutathione pathways in Arabidopsis development. Plant Cell 19: 1851–1865. Google Scholar

276.

H. Rennenberg (1982). Glutathione metabolism and possible biological roles in higher plants. Phytochemistry 21: 2771–2781. Google Scholar

277.

H. Rennenberg , R. Steinkamp , and J. Kesselmeier (1981). 5-oxo-prolinase in Nicotiana tabacum: catalytic properties and subcellular localization. Physiol. Plant. 62: 211–216. Google Scholar

278.

G.C. Rice , E.A. Bump , D.C. Shrieve , W. Lee , and M. Kovacs (1986). Quantitative analysis of cellular glutathione by flow cytometry utilizing monochlorobimane: some applications to radiation and drug resistance in vitro and in vivo. Cancer Res. 46: 6105–6110. Google Scholar

279.

P.G. Richman , and A. Meister (1975) Regulation of γ-glutamylcysteine synthetase by nonallosteric feedback inhibition by glutathione. J. Biol. Chem. 250, 1422–1426. Google Scholar

280.

L. Rizhsky , E. Hallak-Herr , F. Van Breusegem , S. Rachmilevitch , J.E. Barr , S. Rodermel , D. Inzé and R. Mittler (2002). Double antisense plants lacking ascorbate peroxidase and catalase are less sensitive to oxidative stress than single antisense plants lacking ascorbate peroxidase or catalase. Plant J. 32: 329–342. Google Scholar

281.

A. Rochon , P. Boyle , T. Wignes , P.R. Fobert , and C. Després (2006). The coactivator function of Arabidopsis NPR1 requires the core of its BTB/POZ domain and the oxidation of C-terminal oxidases. Plant Cell 18: 3670–3685. Google Scholar

282.

M.A. Rodriguez Milla , A. Maurer , A.R. Huete and J.P. Gustafson (2003). Glutathione peroxidase genes in Arabidopsis are ubiquitous and regulated by abiotic stresses through diverse signaling pathways. Plant J. 36: 602–615. Google Scholar

283.

M.C. Romero-Puertas , F.J. Corpas , L.M. Sandalio , M. Leterrier , M. Rodriguez-Serrano , L.A. Del Rio , and J.M. Palma (2006). Glutathione reductase from pea leaves: response to abiotic stress and characterization of the peroxisomal isozyme. New Phytol. 170: 43–52. Google Scholar

284.

M.C. Romero-Puertas , M. Laxa , A. Matte , F. Zaninotto , I. Finkemeier , A.M.E. Jones , M. Perazzolli , E. Vandelle , K.J. Dietz , and M. Delledonne (2007). S-nitrosylation of peroxiredoxin II E promotes peroxynitrite-mediated tyrosine nitration. Plant Cell 19: 4120–4130. Google Scholar

285.

M.C. Romero-Puertas , N. Campostrini , A. Matté , P.G. Righetti , M. Perazzolli , L. Zoila , P. Roepstorff , and M. Delledonne (2008). Proteomic analysis of S-nitrosylated proteins in Arabidopsis thaliana undergoing hypersensitive response. Proteomics 8: 1459–1469. Google Scholar

286.

N. Rouhier (2010) Plant glutaredoxins: pivotal players in redox biology and iron-sulphur centre assembly. New Phytol. 186: 365–372. Google Scholar

287.

N. Rouhier , E. Gelhaye , and J.P. Jacquot (2002). Glutaredoxin dependent peroxiredoxin from poplar : protein-protein interaction and catalytic mechanism. J. Biol. Chem. 277: 13609–13614. Google Scholar

288.

N. Rouhier , J. Couturier , and J.P. Jacquot (2006). Genome-wide analysis of plant glutaredoxin systems. J. Exp. Bot. 57: 1685–1696. Google Scholar

289.

N. Rouhier , H. Unno , S. Bandyopadhay , L. Masip , S.K. Kim , M. Hirazawa , J.M. Gualberto , V. Lattard , M. Kusunoki , D.B. Knaff , G. Goergiou , T. Hase , M.K. Johnson , and J.P. Jacquot (2007). Functional, structural, and spectroscopic characterization of a glutathione-ligated [2Fe-2S] cluster in poplar glutaredoxin C1. Proc. Natl. Acad. Sci. USA 104: 1373–1379. Google Scholar

290.

N. Rouhier , S. Lemaire , and J.P. Jacquot (2008). The role of glutathione in photosynthetic organisms: emerging functions for glutaredoxins and glutathionylation. Annu. Rev. Plant Biol. 59: 143–166. Google Scholar

291.

K. Saito (2000). Regulation of sulfate transport and synthesis of sulfurcontaining amino acids. Curr. Opin. Plant. Biol. 3: 188–195. Google Scholar

292.

A. Sakamoto , M. Ueda , and M. Morikawa (2002). Arabidopsis glutathione-dependent formaldehyde dehydrogenase is an S-nitrosoglutathione reductase. FEBS Lett. 515: 20–24. Google Scholar

293.

R. Sánchez-Fernández , W. Ardiles-Diaz , M. Van Montagu , D. Inzé , and M.J. May (1998). Cloning and expression analyses of AtMRP4, a novel MRP-like gene from Arabidopsis thaliana. Mol. Gen. Genet. 258: 655–662. Google Scholar

294.

P.G. Sappl , L. Onate-Sanchez , K.B. Singh , and A.H. Millar (2004). Proteomic analysis of glutathione S-transferases of Arabidopsis thaliana reveals differential salicylic acid-induced expression of the plant-specific phi and tau classes. Plant Mol. Biol. 54: 205–219. Google Scholar

295.

P.G. Sappl , A.J. Carroll , R. Clifton , R. Lister , J. Whelan , A.H. Millar , and K.B. Singh (2009). The Arabidopsis glutathione transferase gene family displays complex stress regulation and co-silencing of multiple genes results in altered metabolic sensitivity to oxidative stress. Plant J. 58: 53–68. Google Scholar

296.

Y. Sasaki-Sekimoto , N. Taki , T. Obayashi , M. Aono , F. Matsumoto , N. Sakurai , H. Suzuki , M.Y. Hirai , M. Noji , K. Saito , T. Masuda , K. Takamiya , D. Shibata , and H. Ohta (2005). Coordinated activation of metabolic pathways for antioxidants and defence compounds by jasmonates and their roles in stress tolerance in Arabidopsis. Plant J. 44: 653–668. Google Scholar

297.

K. Schlaeppi , N. Bodenhausen , A. Buchala , F. Mauch , and P. Reymond (2008). The glutathione-deficient mutant pad2-1 accumulates lower amounts of glucosinolates and is more susceptible to the insect herbivore Spodoptera littoralis. Plant J. 55: 774–786. Google Scholar

298.

A. Schneider , T. Schatten , and H. Rennenberg (1992). Reduced glutathione (GSH) transport in cultured tobacco cells. Plant Physiol Biochem. 30: 29–38. Google Scholar

299.

R. Schupp , and H. Rennenberg (1988). Diurnal changes in the glutathione content of spruce needles (Picea abies L.). Plant Sci. 57: 113–117. Google Scholar

300.

M. Schwarzlander , M.D. Fricker , C. Muller , L. Marty , T. Brach , J. Novak , L.J. Sweetlove , R. Hell , and A.J. Meyer (2008). Confocal imaging of glutathione redox potential in living plant cells. J. Microsc. 231: 299–316. Google Scholar

301.

A. Sen Gupta , R.G. Alscher , and D. McCune (1991). Response of photosynthesis and cellular antioxidants to ozone in Populus leaves. Plant Physiol. 96: 650–655. Google Scholar

302.

K. Senda , and K. Ogawa (2004). Induction of PR-1 accumulation accompanied by runaway cell death in the Isd1 mutant of Arabidopsis is dependent on glutathione levels but independent of the redox state of glutathione. Plant Cell Physiol. 45: 1578–1585. Google Scholar

303.

A.T. Setterdahl , P.T. Chivers , M. Hirasawa , S.D. Lemaire , E. Keryer , M. Miginiac-Maslow , S.K. Kim , J. Mason , J.P. Jacquot , C.C. Longbine , F. de Lamotte-Guery , and D.B. Knaff (2003). Effect of pH on the oxidation-reduction properties of thioredoxins. Biochemistry 42: 14877– 14884. Google Scholar

304.

S.L. Singla-Pareek , M.K. Reddy , and S.K. Sopory (2003). Genetic engineering of the glyoxalase pathway in tobacco leads to enhanced salinity tolerance. Proc. Natl. Acad. Sci. USA 100: 14672–14677. Google Scholar

305.

S.L. Singla-Pareek , S.K. Yadav , A. Pareek , M.K. Reddy , and S.K. Sopory (2008). Enhancing salt tolerance in a crop plant by overexpression of glyoxalase II. Trans. Res. 17: 171–180. Google Scholar

306.

M. Skipsey , C.J. Andrews , J.K. Townson , I. Jepson , and R. Edwards (2005a). Diversification in substrate usage by glutathione synthetases from soya bean (Glycine max), wheat (Triticum aestivum) and maize (Zea mays). Biochem J. 391: 567–574. Google Scholar

307.

M. Skipsey , I. Cummins , C.J. Andrews , I. Jepson , and R. Edwards (2005b). Manipulation of plant tolerance to herbicides through co-ordinated metabolic engineering of a detoxifying glutathione transferase and thiol cosubstrate. Plant Biotechnol. J. 3: 409–420. Google Scholar

308.

I.K. Smith , A.C. Kendall , A.J. Keys , J.C. Turner , and P.J. Lea (1984). Increased levels of glutathione in a catalase-deficient mutant of barley (Hordeum vulgare L.). Plant Sci. Lett. 37: 29–33. Google Scholar

309.

I.K. Smith , A.C. Kendall , A.J. Keys , J.C. Turner , and P.J. Lea (1985). The regulation of the biosynthesis of glutathione in leaves of barley (Hordeum vulgare L.). Plant Sci. 41: 11–17. Google Scholar

310.

I.K. Smith , T.L. Vierheller , and C.A. Thorne (1989). Properties and functions of glutathione reductase in plants. Physiol. Plant. 77: 449–456. Google Scholar

311.

R. Steinkamp , and H. Rennenberg (1984). γ-Glutamyltranspeptidase in tobacco suspension cultures: catalytic properties and subcellular localization. Physiol. Plant. 61: 251–256. Google Scholar

312.

R. Steinkamp , and H. Rennenberg (1985). Degradation of glutathione in plant cells: evidence against the participation of a γ-glutamyltranspeptidase. Z. Naturforsch. C 40: 29–33. Google Scholar

313.

R. Steinkamp , B. Schweihofen , and H. Rennenberg (1987). γ-Glutamylcyclotransferase in tobacco suspension cultures: catalytic properties and subcellular localization. Physiol. Plant. 69: 499–503. Google Scholar

314.

R.G. Stevens , G.P. Creissen , and P.M. Mullineaux (2000). Characterisation of pea cytosolic glutathione reductase expressed in transgenic tobacco. Planta 211: 537–545. Google Scholar

315.

S. Storozhenko , E. Belles-Boix , E. Babiychuk , D. Hérouart , M.R. Davey , L. Slooten , M. Van Montagu , D. Inzé , and S. Kushnir (2002). γ-Glutamyl transpeptidase in transgenic tobacco plants. Cellular localization, processing, and biochemical properties. Plant Physiol. 128: 1109–1119. Google Scholar

316.

M. Strohm , L. Jouanin , K.J. Kunert , C. Pruvost , A. Polle , C.H. Foyer , and H. Rennenberg (1995) Regulation of glutathione synthesis in leaves of transgenic poplar (Populus tremula x P.alba) overexpressing glutathione synthetase. Plant J. 7, 141–145. Google Scholar

317.

S. Sundaram , and B. Rathinasabapathi (2010). Transgenic expression of fern Pteris vittata glutaredoxin PvGrx5 in Arabidopsis thaliana increases plant tolerance to high temperature stress and reduces oxidative damage to proteins. Planta 231: 361–369. Google Scholar

318.

D.Y. Sung , D. Lee , H. Harris , A. Raab , J. Feldmann , A. Meharg , B. Kumabe , E.A. Komives , and J.I. Schroeder (2007). Identification of an arsenic tolerant double mutant with a thiol-mediated component and increased arsenic tolerance in phyA mutants. Plant J. 49: 1064–75. Google Scholar

319.

D.Y. Sung , T.H. Kim , E.A. Komives , D.G. Mendoza-Cózatl , and J.I. Schroeder (2009). ARS5 is a component of the 26S proteasome complex, and negatively regulates thiol biosynthesis and arsenic tolerance in Arabidopsis. Plant J. 59: 802–812. Google Scholar

320.

A. Szent-Györgyi (1931) On the function of hexuronic acid in the respiration of the cabbage leaf. J. Biol. Chem. 90: 385–393. Google Scholar

321.

Y. Tada , S.H. Spoel , K. Pajerowska-Mukhtar , Z. Mou , J. Song , C. Wang , J. Zuo , and X. Dong (2008). Plant immunity requires conformational charges of NPR1 via S-nitrosylation and thioredoxins. Science 321: 952–956. Google Scholar

322.

M. Tamaoki , J.L. Freeman , and E.A.H. Pilon-Smits (2008). Cooperative ethylene and jasmonic acid signaling regulates selenite resistance in Arabidopsis. Plant Physiol. 146: 1219–1230. Google Scholar

323.

L. Tarrago , E. Laugier , M. Zaffagnini , C. Marchand , P. Le Maréchal , N. Rouhier , S.D. Lemaire , and P. Rey (2009). Regeneration mechanisms of Arabidopsis thaliana methionine sulfoxide reductases B by glutaredoxins and thioredoxins. J. Biol. Chem. 284: 18963–18971. Google Scholar

324.

M. Tausz , H. Sircelj , and D. Grill (2004). The glutathione system as a stress marker in plant ecophysiology: is a stress-response concept valid? J. Exp. Bot. 55: 1955–1962. Google Scholar

325.

N.L. Taylor , D.A. Day , and A.H. Millar (2002) Environmental stress causes oxidative damage to plant mitochondria leading to inhibition of glycine decarboxylase. J. Biol. Chem. 277: 42663–42668. Google Scholar

326.

L.F. Thatcher , C. Carrie , C.R. Andersson , K. Sivasithamparam , J. Whelan , and K.B. Singh (2007). Differential gene expression and subcellular targeting of Arabidopsis glutathione-S-transferase F8 is achieved through alternative transcription start sites. J. Biol. Chem. 282: 28915– 28928. Google Scholar

327.

F. Tietze (1969). Enzymic method for quantitative determination of nanogram amounts of total and oxidised glutathione. Applications to mammalian blood and other tissues. Anal. Biochem. 27: 502–522. Google Scholar

328.

R. Tommasini , E. Martinoia , E. Grill , K-J. Dietz and N. Amrhein (1993). Transport of oxidized glutathione into barley vacuoles: evidence for the involvement of the glutathione-S-conjugate ATPase. Z. Naturforsch. 48c: 867–871. Google Scholar

329.

E.W Trotter , and C.M. Grant (2002). Thioredoxins are required for protection against a reductive stress in the yeast Saccharomyces cerevisaie. Mol. Microbiol. 46: 869–878. Google Scholar

330.

G. Tsakraklides , M. Martin , R. Chalam , M.C. Tarczynski , A. Schmidt , and T. Leustek (2002). Sulfate reduction is increased in transgenic Arabidopsis thaliana expressing 5′-adenylylsulfate reductase from Pseudomonas aeruginosa. Plant J. 32, 879–889. Google Scholar

331.

I. Tzafrir , R. Pena-Muralla , A. Dickerman , M. Berg , R. Rogers , S. Hutchens , T.C. Sweeney , J. McElver , G. Aux , D. Patton , and D. Meinke (2004). Identification of genes required for embryo development in Arabidopsis. Plant Physiol. 135: 1206–1220. Google Scholar

332.

P. Ullman , L. Gondet , S. Potier , and T.J. Bach (1996) Cloning of Arabidopsis thaliana glutathione synthetase (GSH2) by functional complementation of a yeast gsh2 mutant. Eur. J. Biochem. 236: 662–669. Google Scholar

333.

H. Vanacker , T.L.M. Carver , and C.H. Foyer (1998). Pathogen-induced changes in the antioxidant status of the apoplast in barley leaves. Plant Physiol. 117: 1103–1114. Google Scholar

334.

H. Vanacker , T.L.W. Carver , and C.H. Foyer (2000). Early H2O2 accumulation in mesophyll cells leads to induction of glutathione during the hypersensitive response in the barley-powdery mildew interaction. Plant Physiol. 123: 1289–1300. Google Scholar

335.

S. Vanderauwera , P. Zimmermann , S. Rombauts , S. Vandenabeele , C. Langebartels , W. Gruissem , D. Inzé , and F. Van Breusegem (2005). Genome-wide analysis of hydrogen peroxide-regulated gene expression in Arabidopsis reveals a high light-induced transcriptional cluster involved in anthocyanin biosynthesis. Plant Physiol. 139: 806–821. Google Scholar

336.

G.C. Vanlerberghe , D.A. Day , J.T. Wiskich , A.E. Vanlerberghe , and L. Mclntosh (1995) Alternative oxidase activity in tobacco leaf mitochondria. Dependence on tricarboxylic acid cycle-mediated redox regulation and pyruvate activation. Plant Physiol. 109: 353–361. Google Scholar

337.

P. Vauclare , S. Kopriva , D. Fell , M. Suter , L. Sticher , P. von Ballmoos , U. Krahenbuhl , R.O. den Camp , and C. Brunold (2002). Flux control of sulphate assimilation in Arabidopsis thaliana: adenosine 5′-phosphosulphate reductase is more susceptible than ATP sulphurylase to negative control by thiols. Plant J. 31, 729–740. Google Scholar

338.

S.D. Veljovic-Jovanovic , C. Pignocchi , G. Noctor , and C.H. Foyer (2001). Low ascorbic acid in the vtc-1 mutant of Arabidopsis is associated with decreased growth and intracellular redistribution of the antioxidant system. Plant Physiol. 127: 426–435. Google Scholar

339.

T. Vernoux , R.C. Wilson , K.A. Seeley , J. P. Reichheld , S. Muroy , S. Brown , S.C. Maughan C.S. Cobbett, M. Van Montagu , D. Inzé , M.J. May , and Z.R. Sung (2000). The ROOT MERISTEMLESS1/CADMIUM SENSITIVE2 gene defines a glutathione-dependent pathway involved in initiation and maintenance of cell division during postembryonic root development. Plant Cell 12: 97–110. Google Scholar

340.

D.W. Voehringer , D.J. McConkey , T.J. McDonnell , S. Brisbay , and R.E. Meyn (1998). Bcl-2 expression causes redistribution of glutathione to the nucleus. Proc. Natl. Acad. Sci. USA 95: 2956–2960. Google Scholar

341.

A. Wachter , S. Wolf , H. Steiniger , J. Bogs , and T. Rausch (2005). Differential targeting of GSH1 and GSH2 is achieved by multiple transcription initiation: implications for the compartmentation of glutathione biosynthesis in the Brassicaceae. Plant J. 41: 15–30. Google Scholar

342.

U. Wagner , R. Edwards , D.P. Dixon , and F. Mauch (2002). Probing the diversity of the Arabidopsis glutathione S-transferase gene family. Plant Mol. Biol. 49: 515–532. Google Scholar

343.

W. Wang , and N. Ballatori (1998). Endogenous glutathione conjugates: occurrence and biological functions. Pharmacol. Rev. 50: 335–355. Google Scholar

344.

Z. Wang , S. Zing , R. Birkenbihl , and S. Zachgo (2009) Conserved functions of rice and Arabidopsis CC-type glutaredoxins in flower development and pathogen response. Mol. Plant. 2: 323–325. Google Scholar

345.

M. Watanabe , K. Mochida , T. Kato , S. Tabata , N. Yoshimoto , M. Noji , and K. Saito (2008). Comparative genomics and reverse genetics analysis reveal indispensable functions of the serine acetyltransferase gene family in Arabidopsis. Plant Cell 20: 2484–2496. Google Scholar

346.

W.W. Wells , D.P. Xu , Y. Yang , and P.A. Rocque (1990) Mammalianthi-oltransferase (glutaredoxin) and protein disulfide isomerase have dehydroascorbate reductase activity. J. Biol. Chem. 265: 15361–15364. Google Scholar

347.

H. Willekens , S. Chamnongpol , M. Davey , M. Schraudner , C. Langebartels , M. Van Montagu , D. Inzé , and W. Van Camp (1997). Catalase is a sink for H2O2 and is indispensable for stress defense in C3 plants. EMBO J. 16: 4806–4816. Google Scholar

348.

V.P. Wingate , M.A. Lawton , and C.J. Lamb (1988). Glutathione causes a massive and selective induction of plant defense genes. Plant Physiol. 87, 206–210. Google Scholar

349.

M. Wirtz , and R. Hell (2007). Dominant-negative modification reveals the regulatory function of the multimeric cysteine synthase protein complex in transgenic tobacco. Plant Cell 19, 625–639. Google Scholar

350.

A.E. Wolf , K.J. Dietz , and P. Schroder (1996). Degradation of glutathione S-conjugates by a carboxypeptidase in the plant vacuole. FEBS Lett. 384: 31–34. Google Scholar

351.

Wolosiuk R.A. , and B.B. Buchanan (1977). Thioredoxin and glutathione regulate photosynthesis in choroplasts. Nature 266: 565–567. Google Scholar

352.

J. Wünschmann , A. Beck. , L. Meyer , T. Letzel , E. Grill , and K.J. Lendzian (2007). Phytochelatins are synthesised by two vacuolar serine carboxypeptidases in Saccharomyces cerevisiae. FEBS lett. 581: 1681– 1687. Google Scholar

353.

C. Xiang , D.J. Oliver (1998). Glutathione metabolic genes coordinately respond to heavy metals and jasmonic acid in Arabidopsis. Plant Cell 10: 1539–1550. Google Scholar

354.

C. Xiang , and D. Bertrand (2000). Glutathione synthesis in Arabidopsis: multilevel controls coordinate responses to stress. In Sulfur Nutrition and Sulphur Assimilation in Higher Plants , (eds C. Brunold , H. Rennenberg , L.J. De Kok , I. Stulen & J.C. Davidian ), pp. 409–412. Paul Haupt, Bern, Switzerland. Google Scholar

355.

C. Xiang , B.L. Werner , E.M. Christensen , and D.J. Oliver (2001). The biological functions of glutathione revisited in Arabidopsis transgenic plants with altered glutathione levels. Plant Physiol. 126: 564–574. Google Scholar

356.

Xing S. , and S. Zachgo (2008). ROXY1 and ROXY2, two Arabidopsis glutaredoxin genes, are required for anther development. Plant J. 53: 790–801. Google Scholar

357.

S. Xing , M.G. Rosso , and S. Zachgo (2005). ROXY1, a member of the plant glutaredoxin family, is required for petal development in Arabidopsis thaliana. Development 132: 1555–1565. Google Scholar

358.

Y. Yan , S. Stolz , A. Chétalat , P. Reymond , M. Pagni , L. Dubugnon , and E.E. Farmer (2007). A downstream mediator in the growth repression limb of the jasmonate pathway. Plant Cell 19: 2470–2483. Google Scholar

359.

M. Zaffagnini , L. Michelet , C. Marchand , F. Sparla , P. Decottignies , P. Le Maréchal , M. Miginiac-Maslow , G. Noctor , P. Trost , and S.D. Lemaire (2007). The thioredoxin independent isoform of chloroplastic glyceraldehyde-3-phosphate dehydrogenase is selectively regulated by glutathionylation. FEBS J. 274: 212–226. Google Scholar

360.

M. Zaffagnini , L. Michelet , V. Massot , P. Trost , and S.D. Lemaire (2008). Biochemical characterization of glutaredoxins from Chlamydomonas reinhardtii reveals the unique properties of a chloroplastic CGFStype glutaredoxin. J. Biol. Chem. 283: 8868–8876. Google Scholar

361.

B. Zechmann , and M. Müller (2010). Subcellular compartmentation of glutathione in dicotyledonous plants. Protoplasma 246: 15–24. Google Scholar

362.

B. Zechmann , G. Zellnig , A. Urbanek-Krajnc , and M. Müller (2007). Artifical elevation of glutathione affects symptom development in ZYMV-infected Cucurbita pepo L. plant. Arch. Virol. 152: 747–762. Google Scholar

363.

B. Zechmannn , F. Mauch , and M. Müller (2008). Subcellular immunocytochemical analysis detects the highest concentrations of glutathione in mitochondria and not in plastids. J. Exp. Bot. 59: 4017–4027. Google Scholar

364.

M.Y. Zhang , A. Bourbouloux , O Cagnac , C.V. Srikanth, D. Rentsch , A.K. Bachhawat , and S. Delrot (2004). A novel family of transporters mediating the transport of glutathione derivatives in plants. Plant Physiol. 134: 482–491. Google Scholar

365.

Y.L. Zhu , E.A.H. Pilon-Smits , A.S. Tarun , S.U. Weber , L. Jouanin , and N. Terry (1999a) Cadmium tolerance and accumulation in Indian mustard is enhanced by overexpressing γ-glutamylcysteine synthetase. Plant Physiol. 121: 1169–1177. Google Scholar

366.

Y.L. Zhu , E.A.H. Pilon-Smits , L. Jouanin , and N. Terry (1999b) Overexpression of glutathione synthetase in Indian mustard enhances cadmium accumulation and tolerance. Plant Physiol. 119: 73–79. Google Scholar

367.

M. Ziemann , M. Bhave , and S. Zachgo (2009). Origin and diversification of land plant CC-type glutaredoxins. Genome Biol. Evol. 1: 265–277. Google Scholar

Notes

[1] GSH is used here to indicate the thiol (reduced) form of glutathione while GSSG denotes the disulfide form. The term ‘glutathione’ is used where no distinction is drawn or both forms may be concerned.

© 2011 American Society of Plant Biologists
Graham Noctor, Guillaume Queval, Amna Mhamdi, Sejir Chaouch, and Christine H. Foyer "Glutathione," The Arabidopsis Book 2011(9), 1-32, (18 February 2011). https://doi.org/10.1199/tab.0142
Published: 18 February 2011
Back to Top