Open Access
How to translate text using browser tools
4 April 2002 Oxidative Stress and Acclimation Mechanisms in Plants
Ruth Grene
Author Affiliations +

INTRODUCTION

One of the most crucial functions of plant cells is their ability to respond to fluctuations in their environment. Understanding the connections between a plant's initial responses and the downstream events that constitute successful adjustment to its altered environment is one of the next grand challenges of plant biology. Oxidative stress from environmental sources and developmental transitions such as seed maturation involves the formation of reactive oxygen species (ROS) in plant cells. The redox-modulated changes that follow are central events in cellular responses. Thiol redox regulation (Figure 1) partially mediated through the redox state of the glutathione pool (GSH/GSSG), regulation of the glutathione biosynthetic pathway, and ROS themselves are each thought to have important roles as environmental sensors and/or modulators of global patterns of gene expression in development and defense. Exposure of green tissue to potentially damaging light intensities involves redox sensing molecular events throughout the plant, originating at the plastoquinone (PQ) pool in the thylakoid membrane. Major defense genes whose expression is affected by the redox state of the PQ pool include both cytosolic and chloroplast ascorbate peroxidases (APX) ( Karpinska et al. 2000 ). The superoxide dismutase (SOD) gene families appear to be specialized in function with respect to subcellular location and other as yet unknown factors (Alscher et al. 2002). In the case of peroxisomes, the imposition of oxidative stress gives rise to organelle proliferation, thus adding another layer of complexity to stress responses (Lopez-Huertas et al. 2000). Defense mechanisms involving molecular chaperones and methionine sulfoxide reductase are becoming recognized as important players in resistance to oxidative stress throughout the cell.

Intracellular origins of ROS and their multiple damaging effects

Any circumstance in which cellular redox homeostasis is disrupted can lead to oxidative stress or the generation of ROS (Asada 1994). Production of ROS during environmental stress is one of the main causes for decreases in productivity, injury, and death that accompany these stresses in plants. ROS are produced in both unstressed and stressed cells, and in various locations (Halliwell and Gutteridge 1989) (Figure 2). They are generated endogenously during certain developmental transitions such as seed maturation and as a result of normal, unstressed, photosynthetic and respiratory metabolism. An initial oxyradical product, the superoxide radical (O2-.), upon further reaction within the cell, can form more ROS such as hydroxyl radicals and singlet oxygen. Superoxide is a charged molecule and cannot cross biological membranes. Subcellular compartmentation of defense mechanisms is, therefore, crucial for efficient removal of superoxide anions at their sites of generation throughout the cell. Hydrogen peroxide, on the other hand, which is formed as a result of SOD action, is capable of diffusing across membranes and is thought to fulfil a signaling function in defense responses (Mullineaux et al. 2000).

ROS play an important role in endonuclease activation and consequent DNA damage (Hagar et al. 1996). In the presence of metal ions such as Fe or Cu(II), hydroxyl radicals are formed very rapidly. Hydroxyl radicals can cause damage to all classes of biologically important macromolecules, especially nucleic acids. Hydroxyl radicals can also modify proteins so as to make them more susceptible to proteolytic attack. There is evidently considerable specificity associated with this degradative process since proteins have widely differing susceptibilities to attack by ROS (Davies 1987). Once damaged, proteins can be broken down further by specific endopeptidases such as the one found bound to the thylakoid membrane (Casano et al. 1994). A multicatalytic proteinase complex has been demonstrated in plant systems, with the capacity to selectively break down oxidatively damaged proteins (Van Nocker et al. 1996)

Metabolic defense mechanisms: limiting ROS-mediated damage

Plant cells respond defensively to oxidative stress by removing the ROS and maintaining antioxidant defense compounds at levels that reflect ambient environmental conditions (Scandalios 1997). Metabolic containment mechanisms for ROS involving antioxidant genes and associated processes are likely to have predated or co-evolved with the appearance of aerobiosis and represent fundamental adaptations of aerobic systems to an oxygen dependent metabolism. The mechanisms that act to adjust antioxidant levels to afford protection include changes in antioxidant gene expression (Cushman and Bohnert 2000).

ROS themselves play a role in intracellular redox sensing, activating antioxidant resistance mechanisms, among other adaptive processes (Toledano and Leonard 1991; Karpinski et al. 1997; May, et al. 1998a). A number of redox sensitive transcription factors have been identified in animal, bacterial, and plant cells (Pastori and Foyer 2001).

Functional roles of these responses include the protection of redox-sensitive enzymatic processes, the preservation of membrane integrity, and the protection of DNA and proteins (Scandalios 1997). Redox-sensitive regulatory enzymes such as fructose-1,6-bisphosphatase (FbPase) can be protected from oxidation/inactivation by the action of antioxidants such as glutathione. Under unstressed conditions, the formation and removal of O2 are in balance. The defense system, when presented with increased ROS formation under stress conditions, can be overwhelmed when it is unable to remove the toxic molecular species with increased enzymatic or non-enzymatic antioxidant processes.

Organelles such as the peroxisome and the chloroplast, where ROS are being produced at a relatively high rate, are especially at risk. In the case of the chloroplast, changes in light intensity and temperature or limitations in the substrates of photosynthesis occur frequently, resulting in increased production of ROS (Alscher et al. 1997; Karpinska and Karpinski 2000). ROS are produced at high levels in peroxisomes. Hydrogen peroxide is produced in the peroxisomal respiratory pathway by flavin oxidase. Fatty acid beta oxidation and glycolate oxidase action are other sources of hydrogen peroxide production in the peroxisome. Developmental transitions such as seed maturation, in which peroxisomes play an important role, also involve oxidative stress (Leprince et al. 1990; Walters 1998).

Antioxidant defense molecules have several roles

Ascorbic acid, glutathione, and α-tocopherol have each been shown to act as antioxidants in the detoxification of ROS. These compounds have central and interrelated roles, acting both non-enzymatically and as substrates in enzyme-catalyzed detoxification reactions (Foyer 1993; Hess 1993; Hausladen and Alscher 1994; Winkler et al. 1994; Chaudiere and Ferrari-Iliou 1999). An anti-ROS response includes the induction of genes that belong to ROS scavenging mechanisms.

Metabolic cycles located within the aqueous phase of the peroxisome, chloroplast, cytosol, and the mitochondrion successively oxidize and re-reduce glutathione and ascorbate, using NAD (P) H as the ultimate electron donor. Ascorbate, reduced glutathione (GSH), ascorbate peroxidase (APX), glutathione reductase (GR), superoxide dismutase (SOD), and monodehydroascorbate reductase (MDHAR) are involved in several contexts in antioxidant regeneration throughout the plant cell. The enzymes involved are hydrophilic in nature, although in some instances they are known to be loosely associated with the membranes where the ROS are generated. The version that is found in the chloroplast is shown in Figure 3 including a depiction of the role of the hydrophobic antioxidant α-tocopherol. Different isoforms of the antioxidant enzymes are located in different subcellular compartments (see below). Evidence to date suggests a coordinated response to ROS among different members of the different SOD gene families.

SOD and antioxidant defenses

Within a cell, the superoxide dismutases (SODs) constitute the first line of defense against ROS. O2 is produced at any location where an electron transport chain is present, and hence O2 activation may occur in different compartments of the cell (Elstner 1991), including mitochondria, chloroplasts, microsomes, glyoxysomes, peroxisomes, apoplasts, and the cytosol. This being the case, it is not surprising to find that SODs are present in all these subcellular locations (Figure 4). While all compartments of the cell are possible sites for O2- formation, chloroplasts, mitochondria, and peroxisomes are thought to be the most important generators of ROS (Fridovich 1986).

Takahashi and Asada (1983) showed that phospholipid membranes are impermeable to charged O2 molecules. Therefore, it is crucial that SODs are present for the removal of O2- in the compartments where O2- radicals are formed (Takahashi and Asada 1983). Based on the metal co-factor used by the enzyme, SODs are classified into three groups: iron SOD (Fe SOD), manganese SOD (Mn SOD), and copper-zinc SOD (Cu-Zn SOD). Fe SODs are located in the chloroplast, Mn SODs in the mitochondrion and the peroxisome, and Cu-Zn SODs in the chloroplast, the cytosol, and possibly the extracellular space (Figure 4). Comparison of deduced amino acid sequences from these three different types of SODs suggest that Mn and Fe SODs are more ancient types of SODs (Figure 5), and these enzymes most probably have arisen from the same ancestral enzyme, whereas Cu-Zn SODs have no sequence similarity to Mn and Fe SODs and probably have evolved separately in eukaryotes (Kanematsu and Asada 1990; Smith and Doolittle 1992). The evolutionary reason for the separation of SODs with different metal requirements is probably related to the different availability of soluble transition metal compounds in the biosphere in relation to the O2 content of the atmosphere in different geological eras (Bannister et al. 1991). We now visit each SOD group in turn.

Iron SODs.

The group of Fe SODs probably constitutes the most ancient SOD group. Bannister et al. (1991) suggest that iron was probably the first metal used as a metal cofactor at the active site of the first SOD because of an abundance of iron in soluble Fe (II) form at the time. As the levels of O2 in the environment increased, the mineral components of the environment were oxidized. The decrease in available Fe (II) in the environment caused a shift to the use of a more available metal, Mn (III).

Fe SOD is found both in prokaryotes and in eukaryotes. In eukaryotes it has been isolated from Euglena gracilis (Kanematsu and Asada 1979) and higher plants. Fe SOD is inactivated by H2O2 and is resistant to KCN inhibition. In all plant species examined to date, it is inferred to be located in the chloroplast.

Kliebenstein et al. (1998) report three Fe SODs in Arabidopsis. The absence of Fe SOD in animals has given rise to the proposal that the Fe SOD gene originated in the plastid and moved to the nuclear genome during evolution. Support of this theory comes from the existence of several conserved regions that are present in plant and cyanobacterial Fe SOD sequences, but absent in non-photosynthetic bacteria (Bowler et al. 1994). All three Fe SOD plant sequences encode a unique tripeptide (SRL for N. plumbaginifolia and G. max and ARL for A. thaliana) close to the carboxyl terminus of the enzyme. Although this sequence has shown to direct the proteins to peroxisomes in other proteins, it has yet to be determined whether this is a functional sequence or not. The conserved SRL/ARL sequence is not present in the prokaryotic Fe SOD proteins showing that it is not obligatory for enzyme function (Van Camp et al. 1994).

Manganese SODs.

As the levels of O2 in the environment increased, the amount of available Fe (II) in the environment decreased, causing a shift to the more available metal, Mn (III). As a consequence, Mn SODs are likely to be second only to Fe SODs in antiquity and certainly evolved from the ancestral Fe SODs, perhaps by way of the cambialistic SODs. Mn SODs occur in mitochondria and peroxisomes. Mn SODs carry only one metal atom per subunit. These enzymes cannot function without the Mn atom present at the active site. Even though Mn and Fe SODs have a high similarity in their primary, secondary, and tertiary structure, these enzymes have diverged sufficiently that Fe (II) could not restore the activity of Mn SOD and vice versa (Fridovich 1986). Catalysis by Mn SODs is through the attraction of negatively charged O2- molecules to a site formed from positively charged amino acids present at the active site of the enzyme. The metal present in the active site then donates an electron directly to the O2-, reducing one O2- molecule, which in turn forms H2O2 by reacting with a proton (Asada 1994; Bowler et al. 1994). Plant Mn SODs have approximately 65% sequence similarity to one another, and these enzymes also have high similarities to bacterial Mn SODs (Bowler et al. 1994).

Although Mn SOD is known as the mitochondrial enzyme of eukaryotes, an Mn-containing SOD has also been located in the peroxisomes. del Rio et al. (1992) showed the presence of one peroxisomal and one mitochondrial Mn SOD by using immunolocalization assays in watermelon. Four genes that encode Mn SOD were reported in maize (Zea mays) (Zhu et al. 1999 ). Deduced amino acid sequences from these four isoenzymes have a mitochondrial targeting sequence, indicating that all are located in the mitochondria. In Nicotiana plumbaginifolis, two nuclear-encoded Mn SOD genes were isolated and the tissue-specific expression Mn SOD was shown by analyzing promoter fusion with β-glucuronidase (GUS) in transgenic plants (Van Camp et al. 1996).

Copper-Zinc SODs.

When the atmosphere was completely replenished with oxygen, Fe (II) was almost completely unavailable in the atmosphere and insoluble Cu (I) was converted into soluble Cu (II). At this stage, Cu (II) began to be used as the metal cofactor at the active sites of SODs. Since Fe and Mn have similar electrical properties, the transition from the use of iron to the use of manganese required little change in SOD protein structure. Thus, Mn and Fe SODs are structurally very similar. The existence of archaic Mn/Fe1-containing SODs supports this theory. However the electrical properties of Cu-Zn2 SODs differ greatly from those of Fe and Mn SODs. Therefore, a major change in the structure of the protein occurred after Cu became a metal cofactor (Bannister et al. 1991).

There are at least two forms of Cu-Zn SOD in plant cells, chloroplastic and cytosolic. Deduced amino acid sequences of these two isoforms show approximately 68% similarity, whereas there is approximately 90% similarity among the chloroplastic Cu-Zn SODs (Cu-Zn SODchl) and 80-90% similarity among the cytosolic Cu-Zn SODs (Cu-Zn SODcyt) (Figure 6). Cu-Zn SODchl is a soluble enzyme and is localized in the stroma (Asada and Kiso 1973; Asada et al. 1973). Localization studies performed with an immunogold-labeled antibody raised against Cu-Zn SODchl from spinach leaves showed that this soluble enzyme is not uniformly distributed in the chloroplast but rather is localized mainly on the stromal face of thylakoid membranes (Ogawa et al. 1995) where photosystem I (PSI) is located. The two other Cu-Zn SODs are considered “cytoplasmic” because they have not been detected in intact chloroplasts (Ogawa et al. 1996). However, when immunogold-labeled antibodies raised against “cytosolic” Cu-Zn SOD were used in localization experiments, it was shown that these enzymes were located in the nucleus and apoplast (Ogawa et al. 1996). More than 40% of the immunogold particles were found in the apoplast and approximately 25% was found in the nucleus (Ogawa et al. 1996). Researchers proposed that Cu-Zn SOD in the apoplast functions in lignification and that in the nucleus it protects the cell against fatal mutations caused by O2- molecules (Ogawa et al. 1996; Ogawa et al. 1997).

SOD expression and activity under oxidative stress

Kliebenstein et al. (1998) subjected Arabidopsis to a series of oxidative stresses and observed changes in the seven SODs -- three Fe SODs denoted FSD1, FSD2, and FSD3; three Cu-Zn SODs denoted CSD1, CSD2, and CSD3; and one Mn SOD denoted MSD1 -- that are present in that plant, both at the mRNA and the protein level. They reported increases in FSD2 in response to UV irradiation, and to high light at the mRNA level, but no response of FSD2 mRNA to ozone exposure. They found that FSD1 is under the control of a circadian clock at the mRNA level. Kliebenstein et al. (1998) did not report enzyme activities. However, it is of interest to note that they report FSD2 responding to specific stresses. Thus, it appears that members of the Fe SOD family are specialized for specific responses to stress.

Kliebenstein et al. (1998) report no effect of their series of oxidative stress treatments on Mn SOD in Arabidopsis. In contrast, Mn SOD was reported to respond positively to salt stress (Hernandez et al. 1993; Hernandez et al. 1995; Gomez et al. 1999), manganese toxicity (Gonzalez et al. 1998), chilling stress (Lee et al. 1999; Lee and Lee 2000), and drought (Wu et al. 1999). An explanation for the differences in responses between the Mn and Fe SODs in plants can be correlated with their disparate subcellular locations and the sites of action of the various oxidative stresses that were used. The stresses that did not affect Mn SOD may all have their site of action in the chloroplast.

Overexpression of SODs can lead to protection against specific stresses. The importance of subcellular location. Plants overexpressing SODs have been engineered with the goal of increasing stress tolerance. Both successful and unsuccessful results have been obtained from attempts to create resistant plants (Tepperman and Dunsmuir 1990; Perl et al. 1993; Slooten et al. 1995; Van Camp et al. 1996).

The site of action for ozone has been suggested to be in the apoplast, but when Mn SOD was overexpressed in the chloroplasts of tobacco plants, less damage was observed in the leaves. When the overproduction of the enzyme was in the mitochondria, less protection was observed. Also, it was shown that high levels of chloroplastic Mn SOD activity protected the plant from visible injury caused by ozone, suggesting that ozone may cause oxidative stress in the chloroplast, as well as the apoplast (Van Camp et al. 1994). When an A. thaliana Fe SOD gene was targeted into N. tabacum cv Petit Havana SR1 chloroplasts, an increased protection against O2 generated both in the plasmalemma and photosystem II (PSII) was observed (Van Camp et al. 1996). Arisi et al. (1998) overexpressed an Arabidopsis Fe SOD in poplar and observed a protective effect on PSII under limiting carbon dioxide conditions. Van Camp et al. (1994) overexpressed Fe SOD in the chloroplast and found that induction of Cu-Zn SODs was suppressed, lending some credence to the hypothesis that a cross-family signaling pathway for SOD induction exists in the plant cell. In contrast, when mitochondrial Mn SOD was targeted to the chloroplasts of tobacco plants, protection was only observed against stress generated in the plasmalemma. An increase in the activity of APX, DHAR, and MDHAR, other scavenging pathway enzymes, was also observed. It was concluded that the protection provided by overproduction of Mn SOD was dependent on whether the other enzymes — DHAR and MDAR — were or were not present in elevated levels

Bowler et al. (1994) targeted N. plumbaginiafolis mitochondrial Mn SOD sequence into N. tabacum mitochondria and chloroplasts. When the transformed plants were treated with methyl viologen, it was observed that the plants that had Mn SOD expressed in their chloroplast had remarkable protection, accompanied by an increased SOD activity against methyl viologen stress compared to control plants in light. This protective effect was less observable in plants that were kept in the dark, since the H2O2 scavenging system is not activated without light (Foyer and Halliwell 1976; Nakano and Asada 1980). Even though Mn SOD is not inactivated by H2O2, the balance between generated O2- and H2O2 is disrupted in these transgenic plants, which may increase the formation of OH.. Although increased protection was observed in plants that had foreign Mn SOD targeted into the mitochondria, this effect was not as remarkable compared to plants that had Mn SOD targeted into the chloroplasts. This may be due to the majority of the O2 being generated in the chloroplast in plants exposed to methyl viologen in light (Bowler et al. 1991).

Taken together, these results suggest that regulation of expression of plant Fe and Mn SODs may differ. The subcellular location of the protein also appears to play a role. In some instances, both chloroplastic and cytosolic Cu-Zn SODs afford protection against the same stresses as Fe SOD. There appears, also, to be specialization within the Fe SOD family. We have compared the upstream sequences of all Arabidopsis SOD genes in order to determine similarities or differences among their respective regulatory regions.

Comparing the information in SOD genomic sequences.

With the complete sequencing of the Arabidopsis genome, it is a propitious time to ask what the appropriate application of computation to that genome might reveal about the evolution and regulation of SODs in plants. The upstream regions of the seven Arabidopsis SOD open reading frames were extracted with the intention of mining them for promoter sequences related to the regulation of gene expression under different stresses. We developed bioinformatics tools for searching an upstream region for particular promoter sequences that bind to known transcription factors (Table 1). The area of computational tools for analyzing promoter sequences is in its infancy (Prestridge 2000), and existing tools are limited in their abilities to address our questions and to analyze an entire genome. The ABA responsive element (ABRE) appears to be associated with genes responding to osmotic stress (high osmoticum, salt, desiccation, and cold) and binds to several similar sequences of eight nucleotides (Guan and Scandalios 1998; Choi et al. 2000); we used the consensus sequence YACGTGGC. NF-κ?β is a transcription factor that activates immunoglobulin-k genes; we utilized the consensus sequence GGRNNYYCC (Smith et al. 2000). The heat shock protein gene promoter consensus sequence is the palindromic sequence TTCNNGAA (Santos et al. 1996). Finally, the Y-box motif has consensus sequence GATTGG and mediates redox-dependent transcription activation (Guan and Scandalios 1998). The seven promoter regions were examined for these four consensus sequences. Diamonds (♦) in Table 1 summarize the exact or close matches found at plausible upstream locations (within 1000 nucleotides of the ATG where transcription begins). The different patterns of consensus sequences found suggest that the phenomenon of differential expression of the SOD genes under different stresses can be explained, at least in part, through promoter sequence analysis. Further work will include extracting the upstream sequences for all known or putative Arabidopsis genes and comparing these sequences on a genome-wide basis with new computational tools we are developing. Employing also information concerning co-regulated genes it will become possible to identify what promoter sequence combinations correspond to what kind of regulation.

Multiple defense roles for glutathione as antioxidant and redox sensor.

Upon the imposition of oxidative stress, the existing pool of reduced glutathione (GSH) is converted to oxidized glutathione (GSSG) and glutathione biosynthesis is stimulated (May and Leaver 1993; Madamanchi et al. 1994). The rate-limiting step for glutathione synthesis is thought to be gamma-glutamylcysteine synthetase, which is feedback regulated by GSH and is controlled primarily by the level of available L-cysteine (May et al.1998b). Increasing glutathione biosynthetic capacity has been shown to enhance resistance to oxidative stress (Arisi et al. 1998; Zhu et al. 1999). Glutathione reductase (GR) activities increase as the glutathione pool increases through a multi-level control mechanism, which includes coordinate activation of genes encoding glutathione biosynthetic enzymes and GR (Xiang and Oliver 1998).

Glutathione acts as a redox sensor of environmental cues and forms part of the multiple regulatory circuitry coordinating defense gene expression. There are two known genes encoding GR (Madamanchi et al. 1992; Creissen et al. 1995a; Creissen et al. 1995b). The redox state of the GSH/GSSG couple may act as a direct link between environmental cues and crucial molecular adaptive responses of plant cells (Hausladen and Alscher 1993; Broadbent et al. 1995; Roxas et al. 1997; May et al. 1998a; May et al. 1998b; Baginsky et al. 1999). Glutathione has been reported to regulate rates of cell division (Sanchez-Fernandez et al. 1997) and the induction of antioxidant defenses, as exemplified by the induction of Cu-Zn SOD (Herouart et al. 1993). Glutathione has been suggested as an intermediary in a redox sensing signaling pathway in plants involving the ROS-mediated oxidation of membrane lipids to oxylipins as the initial step (Ball et al. 2001). It is also thought to act specifically as a regulator of the transcription of chloroplast genes (Baginsky et al. 1999). The antioxidant genes known to date are all nuclear-encoded, however.

Previous biochemical data demonstrated the existence of multiple isoforms of GR, which have been assigned to various organelles, including the peroxisome, the chloroplast, the mitochondrion and the peroxisome (Creissen et al. 1991). Only two Arabidopsis GR genes have been characterized, however (Stevens et al. 2000). A peroxisomal GR gene has not been characterized to date, although a peroxisomal GR protein has been described (Corpas et al. 2001). Consequently, it is very likely that other members of the GR gene family exist.

Oxidative stress through impaired mitochondrial electron transport in developing seeds. Role of glutathione.

Seeds develop through a series of stages that may be tracked at morphological, physiological, and molecular levels. Arabidopsis seeds are orthodox (desiccation tolerant) and their ontogeny consists of some 20 distinct developmental stages. Embryogenesis in Arabidopsis is completed in approximately 9 days at 25°C. During the final stage of embryogenesis, seeds must acquire desiccation tolerance to survive. As both the fruit (silique) and seeds dry, metabolism declines rapidly as water is lost. As respiratory activity slows during desiccation, electrons leak and react with oxygen to generate reduced oxygen species, so antioxidants are necessary to protect mitochondria from damage (Vertucci and Farrant 1995). Desiccation tolerance is one of the unusual properties of seeds that, despite intensive investigation, remains rather poorly understood. Free radical scavengers may provide additional protection during desiccation because the development of desiccation tolerance, which involves a period of water stress, coincides with an increase in free radical scavengers in seeds (Vertucci and Farrant 1995; Haslekas et al. 1998). Water stress involves the production of ROS and their containment by antioxidants (Bohnert and Jensen 1996; Bohnert and Shevelava 1998). Thus, the availability of antioxidants such as GSH may be essential for seed maturation. Since GR is a major determinant in the maintenance of reduced glutathione levels, it is not surprising to find that seed aging has been linked to decreased GR activity (De Vos et al. 1994; Bailly et al. 1996; et al. 1996). Lipid peroxidation and autoxidation are oxidative reactions associated with seed deterioration and death (Walters 1998) that may be mitigated by radical scavenging systems.

Transgenic Arabidopsis, antisense with respect to either one of the two known GR genes (GR1 and GR2), have been created to examine the protective role of these antioxidant gene products against ROS throughout development. We have characterized transgenic antisense Arabidopsis thaliana (L) Heynh. plants that are depleted in the expression of one or other of the two known Arabidopsis glutathione reductase genes (GR1, organellar and GR2, cytosolic), (Grene, unpublished). Our working hypothesis is that activity levels of GR play a pivotal role in redox sensing and adjustment processes as well as a direct role in the maintenance of reduced glutathione. Abnormal seed morphologies and/or embryo abortions first appeared in antisense GR2 plants (ANGR2) 8 days after flowering in at least four transformant lines. Seeds of antisense GR1(ANGR1) lines showed 33% abnormalities at maturity. An altered phenotype expressed as a change in seedling growth habit was observed in ANGR2, but not in antisense ANGR1 lines. ANGR2 lines showed an increased time to flowering. GR genes may play important individual roles in seed development, and subsequent seedling development related to the levels of glutathione and effects of ROS. Since the peroxisome contains GR, its ROS sensing mechanism(s) of peroxisomes are likely to respond to alterations in glutathione metabolism. GR1 is present in the chloroplast where it protects photosynthesis, while GR2 is thought to be cytosolic. However, GR1 may reside in the mitochondrion as well (Creissen et al. 1995b). Our data to date suggest the importance of both genes for seed viability and for seedling development.

Increasing glutathione biosynthetic capacity has been shown to enhance resistance to oxidative stress (Arisi et al. 1998; Zhu et al. 1999). The ability of the plant cell to maintain and increase reduced glutathione levels is an important factor in protecting photosynthesis against sulfur dioxide (Alscher et al. 1987), which depends at least in part on GR. Transgenic tobacco overexpressing (3x) an E. coli gene (gor) encoding GR targeted to the plastid was more resistant to paraquat and sulfur dioxide than were nontransformed plants (Aono et al. 1993). Over-expression of gor in the chloroplasts of poplar led to increases in both total glutathione pool sizes and in the ratio of GSH to GSSG. Transformed plants showed enhanced protection against chloroplast-localized oxidative stress (Foyer et al. 1995). Transgenic tobacco plants with increased levels of GSSG were found to grow better under salinity and chilling stress than their nontransformed counterparts, suggesting that resistance/adaptive pathways were stimulated by GSSG (Roxas et al. 1997).

Ascorbate and APX : Multiple roles in antioxidant defenses

APX exists as a multigene family in Arabidopsis. APX 1 and 2 are both cytosolic enzymes. A membrane associated APX has been described in the peroxisome, and also in the chloroplast (Mullen and Trelease 2000). The chloroplast contains two distinct APX enzymes as well, one which is free in the stroma, and one which is associated with the thylakoids. The expression of all of the known members of the APX gene family are affected by events originating in the chloroplast (Karpinski et al. 1997) . When photosynthetic tissue is shifted from low to high light (EEE), ROS are produced, in particular hydrogen peroxide. Reduced glutathione is oxidized to GSSG, and cytosolic APX2 mRNA becomes detectable. APX2 mRNA is induced within seven minutes of exposure to EEE. The induction of both APX1 and APX2 is causally related to the redox state of the plastoquinone pool (Pfannschmidt et al. 1999).

Sen Gupta et al. (1993) reported that 3-fold increases in cytosolic APX activity and APX mRNA occur in transgenic tobacco plants that overexpress chloroplast Cu-Zn SOD. This induction could be mimicked in non-transformed tobacco leaf discs with the addition of H2O2. Van Camp et al. (1996) demonstrated the existence of three separate APX isozymes in Arabiodopsis, two plastidic and one cytosolic, on activity gels. The data of Van Camp et al. (1996) showed that chloroplast APX1 activity was higher in transgenic plants overexpressing Fe SOD. These results suggest that increasing the level of gene product in one part of the pathway can affect other enzymes in the same pathway. This type of co-regulation may be critical in interpreting results using antisense plants where the cytosolic Cu-Zn SOD is suppressed. These data also suggest that H2O2 plays a role in a stress-responsive signal transduction pathway, since the presence of active Fe SOD enzyme in the overexpressors could lead to increased levels of cellular H2O2. Prasad et al. (1994) present evidence that is compatible with a signaling role for hydrogen peroxide in the acclimation of maize seedlings to chilling stress. Recently a gene which acts to regulate APX gene expression has been found, in which glutathione levels are lower, but the glutathione biosynthetic pathway itself is unaffected (Mullineaux et al. 2000).

Stress-mediated changes in the abundance of a particular transcript do not always correlate with corresponding changes in antioxidant protein level and/or enzyme activities (Williamson and Scandalios 1992; Mittler and Zilinskas 1994). Paraquat-mediated increases in APX mRNA levels in pea leaves were not reflected in correspondingly large increases in APX enzyme activities (Donahue et al. 1997). The data of Mittler and Zilinskas (1994) for APX responses to drought also suggest the existence of stress-mediated post-transcriptional processes.

In addition to its role in the scavenging cycle, ascorbate acts as a reductant in the regeneration of α-tocopherol and in the zeaxanthin cycle (Foyer 1993) (Figure 3). A third role of ascorbate is at the thylakoid surface within the chloroplast where it acts as reductant in the APX-mediated scavenging of H2O2. Ascorbate is reduced to the monodehydroascorbate radical as a result of this thylakoid-associated process (Grace et al. 1995). Ascorbic acid is regenerated by a light-dependent process at the thylakoid that can utilize ferredoxin as the source of reductant. In cases where fully oxidized dehydroascorbic acid (DHA) is produced, reduced glutathione (GSH) is the source of reductant for the regeneration of ascorbic acid. Since APX appears to play such an essential role in the scavenging process, the processes within the chloroplast associated with oxygen uptake and APX function and the reduction of molecular oxygen have been named “Mehler-ascorbate peroxidase photorespiration” (Mullineaux et al. 2000) (Figure 7).

Molecular chaperones

Four distinct functions have been assigned to molecular chaperones. They can act as repair proteins, they can remove proteins that are irretrievably damaged, and they can facilitate the import of newly synthesized proteins into the interior of organelles such as the peroxisome. The fourth function is as antioxidant molecules themselves in conjunction with protein methionine-sulfoxide reductase.

Molecular chaperones interact to protect against heat and water stress through the repair of denatured proteins. Evidence is accumulating to suggest an important role for heat shock proteins/molecular chaperones in stress resistance in plant and animal systems (Gustavsson et al. 1999; Wehmeyer and Vierling 2000; Harndahl et al. 2001). Increased expression of HSPs of the 70, 101 and sHSP classes were observed in drought acclimated rooted cuttings of loblolly pine. HSP70 is known to occur in glyoxysomes; in fact, the glyoxysomal protein is encoded by the same gene as the chloroplast form of the protein (Wimmer et al. 1997). The small heat shock proteins that are localized to the cytosol appear to respond to specific developmental signals associated with the acquisition of desiccation tolerance that occur during seed development (Wehmeyer and Vierling 2000). A small HSP (HSP18.1) has been shown to interact with HSP70 to reactivate heat-denatured luciferase (Lee and Vierling 2000). HSP90 was ineffective in their reactivation system, and was also found to be unresponsive to drought stress in loblolly pine (Heath et al. 2002). The cytosolic HSPs prevent heat-mediated and water-stress-mediated aggregation of proteins. These HSPs may prevent loss of conformation in low-water conditions and may be important in peroxisomes. Denatured substrate proteins are bound to the sHSP oligomers in vitro presumably by hydrophobic regions (Lee and Vierling 1998; Harndahl et al. 2001). It is thought that the sHSPs act to bind denatured proteins and to maintain them in a state that allows for ATP-dependent refolding by larger HSPs/molecular chaperones (Lee and Vierling 2000). The small heat shock proteins that are localized to the cytosol appear to respond to specific developmental signals associated with the acquisition of desiccation tolerance that occur during seed development (Wehmeyer and Vierling 2000). No small heat shock proteins have been specifically associated with peroxisomes or glyoxysomes to date. However, since the HSP70 present in those organelles is so similar to the corresponding chloroplast HSP70, it is possible that just such an interaction between sHSPs and the larger molecules may in fact exist. A protein of the DnaK (J) class was found to be essential for the HSP70-mediated refolding/repair mechanism (Lee and Vierling 2000). A DnaJ protein has been found in association with glyoxysomes (Diefenbach and Kindl 2000), albeit at the membrane surface. This protein reacts specifically with a particular cytosolic isoform of HSP70, and not with other forms of HSP70. Another DnaJ protein has been identified as a peroxin in a yeast system (Hettema et al. 1998).

Molecular chaperones and methionine sulfoxidation. Surface methionine residues are preferentially oxidized in proteins. Methionine residues act as an antioxidant protein reservoir [Figure 8 adapted from Hoshi and Heinemann (2001)]. Amino acid residues in proteins are one of the major targets of ROS attack. The side chains of methionine and cysteine are more sensitive to oxidation than the side chains of other amino acids. This differential sensitivity has been exploited by nature to create a protective mechanism against ROS-mediated protein damage. Cysteine residues are maintained in a reduced state through the action of glutathione. Surface-exposed methionine residues are available for oxidation by molecules such as hydrogen peroxide, thus effectively lowering the degree of the threat. In the case of glutamine synthetase from E. coli, it was found that 8 of the 16 methionine residues present in the protein could be oxidized with little effect in overall enzyme activity (Levine et al. 1999). They are re-reduced through the action of protein methionine sulfoxide reductase (PMSR), using thioredoxin as a source of reductant (Lowther et al. 2000; Hoshi and Heinemann 2001). PMSR is a highly conserved protein from E.coli to human (Brot and Weissbach 2000). Combined with the reductive action of PMSR, which restores the methionines to their original state, a larger mechanism within the cell is now able to maintain a stable configuration of protein state. It is thought that the degree to which methionine oxidation occurs may be underestimated due to imprecision of assay methods (Squier and Bigelow 2000). A proposed overall scheme for redox regulation of cellular defense processes involving thiol redox control is represented in Figure 1. sHSPs contain a unique methionine rich domain at its N terminus, which exists as an amphipathic alpha helix. Some of these met residues exist at the surface of the protein, and are oxidized very readily, with resultant loss of chaperone activity (Harndahl et al. 2001). The chloroplast HSP21 is easily and reversibly oxidized by the same concentration of hydrogen peroxide as that which brings about HSP21 oxidation in vivo in Arabidopsis leaves (Harndahl et al. 1999). The oxidation of surface methionine residues, which is mediated by the hydroxyl radical, ozone, and poeroxynitrile, as well as hydrogen peroxide, has been proposed to function in an antioxidant capacity in animal and yeast systems (Moskovitz et al. 1997; Preville et al. 1999).

Molecular chaperones can interact with glutathione to protect against oxidative stress. Mehlen et al. (1996) and Garrido et al. (1998) present evidence of an interaction between sHSPs and glutathione in mammalian cells that resulted in increased resistance to cell death induced by tumor necrosis factor or by hydrogen peroxide. Resistance was dependent both on increases in reduced glutathione and on increases in expression of sHSPs, and was shown to decrease the levels of cellular ROS. Preville et al. (1999) demonstrated that sHSPS can protect against oxidative stress in L929 cells. This interaction has not yet been investigated in plant systems. However, much evidence points to the importance of glutathione biosynthesis in protection against ROS damage (Alscher 1989; Noctor et al. 1998; Noctor and Foyer 1998; May et al. 1998a).

Redox sensing and photosynthetic function

Studies of the responses of foliar tissue to oxidative stress have focused primarily on the photosynthetic machinery. Within the chloroplast, the rapid loss of oxygen evolution activity in the presence of Cu(II) was found to be mainly due to the formation of OH. radicals from superoxide ion via a Cu(II)-catalyzed Haber-Weiss mechanism (Yruela et al. 1996). Hydrogen peroxide and lipid hydroperoxides, other potentially toxic ROS, are also generated. Hydrogen peroxide can cause DNA breakage, as described above, and also can inactivate thiol-containing enzymes such as the thioredoxin-modulated enzymes of the chloroplast stroma (Hagar et al. 1996). An imbalance in which the redox steady state of the cell is altered in the direction of pro-oxidants can result in the potentially dangerous univalent reduction of molecular oxygen to the superoxide anion radical described above. In the case of photosynthetic electron transport, oxygen uptake associated with the photoreduction of oxygen to superoxide is called the Mehler reaction, in honor of its initial discoverer (Figure 7). Changing environmental conditions such as vicissitudes of temperature, humidity, water availability, salt stress, or light intensity can lead to increased production of ROS within the chloroplast. Damage of leaves due to air pollutants such as sulfur dioxide and ozone or photodynamic herbicides such as paraquat is also mediated through the production of ROS (Mehlhorn et al. 1990; Foyer and Mullineaux 1994; Okpodu et al. 1996). The site of action of sulfur dioxide, photodynamic herbicides, and high light is the chloroplast, whereas ozone and most pathogens act in the extracellular space. Species, biotypes, or cultivars often show cross-resistance to two or more oxidative stresses (e.g. paraquat, high light, sulfur dioxide). These observations have been used to propose a common basis for resistance against ROS (Gressel and Galun 1994). However, ozone tolerance was not correlated with resistance to paraquat in tobacco. A first hypothesis might be that cross-resistance occurs between stresses which originate in the same subcellular compartment (e.g., the chloroplast in the case of paraquat and sulfur dioxide) but not between stresses which have disparate sites of action within the plant leaf (e.g., ozone and paraquat).

Redox sensing plays a central role in the interaction of oxidative stress, in particular, excess excitation energy (EEE), with the photosynthetic machinery (Figure 9) (Karpinski et al. 1997; Karpinski et al. 1999; Pfannschmidt et al. 1999; Karpinska et al. 2000; Mullineaux et al. 2000). Under light saturating conditions, only 10% of absorbed light energy is used to fix carbon (Mullineaux et al. 2000). The remaining energy must be dissipated in non-destructive ways. Under these conditions, or under conditions where carbon dioxide or water are limiting, redox changes associated with the plastoquinones are thought to act as environmental sensors, inducing signal transduction pathways that result in the activation of defense genes such as cytosolic APX, among others. Redox sensing via the plastoquinone pool affects both chloroplast and cytosolic gene expression.

Redox sensing: the central role of hydrogen peroxide

Exposure to hydrogen peroxide has been reported to result in the induction of at least 100 genes in Arabidopsis (Neill et al. 2001). Hydrogen peroxide accumulation in barley that had been inoculated with powdery mildew was shown to induce glutathione biosynthesis during the hypersensitive response (Vanacker et al. 2000). Molecular chaperones, glutathione-S-transferases, various protein kinases, and redox-sensitive transcription factors (Pastori and Foyer 2001) are known ROS-responsive genes. Some of these genes, e.g. the molecular chaperones, are also induced by long term exposure to drought (Heath et al. 2002). Hydrogen peroxide is thought to exert its inductive effect through oxidation of cysteine and methionine residues, oxidation of particular membrane lipids which act as oxylipin receptors in a signaling cascade that involves glutathione (Hamberg 1999; Ball et al. 2001), and direct influences on protein kinase cascades (Kovtun et al. 2000). Kovtun et al. (2000) demonstrated the existence of a hydrogen peroxide mediated protein kinase cascade, which signals the activation of defense genes such as HSP18 (small heat shock protein) and glutathione S-transferase.

Hydrogen peroxide has been shown to participate in the systemic acclimation response demonstrated by Karpinski et al. (1999) in which partial exposure of low light adapted Arabidopsis plants to EEE results in acclimation to high light in unexposed leaves (Figure 10). Since more hydrogen peroxide is produced in the peroxisome as photorespiration increases, the peroxisome may contain a redox sensing mechanism in addition to the PQ redox- sensing mechanism in the chloroplast. If EEE-exposed leaves are pre-treated with glutathione or ascorbate, APX1 and 2 induction is less, and the leaves become more susceptible to oxidative stress. The application of catalase, but not SOD, leads to the disappearance of the systemic acclimation phenomenon, confirming the importance of hydrogen peroxide in the signaling pathway. Treatment of leaves with hydrogen peroxide prior to their exposure to EEE results in greater resistance to photooxidative damage, presumably because of the induction of defense responses prior to exposure to stress (Karpinska et al. 2000).

Hydrogen peroxide mediated peroxisomal biogenesis. Lopez-Huertas et al. (2000) reported a stimulatory effect of hydrogen peroxide on peroxisomal biogenesis. Hydrogen peroxide is both a means of induction of stress resistance, and a product of stress imposition in the plant cell. Consequently, it has been widely used as a tool to investigate the mounting of defense mechanisms and the response of defense pathways. In the case of peroxisomes, their response to the presence of hydrogen peroxide consists in an increase in peroxisomal biogenesis. A major component of peroxisomal biogenesis is the import of cytosolic-synthesized proteins into the interior of the organelle. Import of proteins into the peroxisome is known to involve at least one molecular chaperone — an HSP70 (Corpas et al. 2001). The HSP70 interacts with a DNA J-like protein at the peroxisomal surface. Since hydrogen peroxide was used as the only representative ROS in the peroxisomal proliferation experiments, there is no information as yet on effects of other ROS on peroxisomal biogenesis. Charged species such as the superoxide anion cannot cross membranes, and thus are not so likely as candidates for mediating signals across bounding membranes. The inhibition of catalase by the superoxide anion would result in increased levels of hydrogen peroxide, however, which could constitute the initiation of a peroxisome-specific signaling cascade. Peroxisomal proliferation also occurs during senescence (Corpas et al. 2001) and may result from the well-documented increases in ROS levels that occur with aging.

Prospects for improving stress resistance

The emerging picture of ROS-mediated cellular events and redox control of gene expression is complex (Figure 11). Glutathione remains a prime candidate for engineering increased stress resistance in plant cells, especially in organelles such as the peroxisome where ROS are produced at high levels. Since the ascorbate/glutathione scavenging pathway is present in peroxisomes, chloroplasts, and mitochondria, it is likely that glutathione levels are high in all these subcellular compartments, and that a mechanism for providing GSH to the organelle exists. Transport of glutathione across the peroxisomal and mitochondrial bounding membranes is one possible focus for engineering increased stress resistance. Engineering increased flux through the cytosolic glutathione biosynthetic pathway is another possibility; a strategy that has proved to provide increased protection to photosynthesis in poplar (Foyer et al. 1995). HSP70s are known to function in transport of proteins into the interior of the peroxisome, and a member of the HSP70 gene family has been identified in the interior of the organelle. It is not yet known if the peroxisome contains sHSPs as well as HSP70s. Exposure to hydrogen peroxide results in the induction of peroxisomal biogenesis. Kovtun et al. (2000) demonstrated the existence of a hydrogen peroxide mediated protein kinase cascade, which signals the activation of defense genes such as HSP18 (small heat shock protein) and glutathione S-transferase. The ROS-sensing mechanism that elicits peroxisomal biogenesis is unknown, but it could also involve sHSPs and glutathione in a manner analogous to the mammalian system.

Although a cytosolic and a plastidic form of PMSR have been described in plants (Sadanandom et al. 2000), little information concerning their respective roles in antioxidant defense is yet available. A search for the in vivo substrates of PMSR in the peroxisome and the cytosol could yield valuable information. The PMSR mechanism may constitute an important additional antioxidant mechanism. Peroxisomal HSPs containing the methionine rich region present in HSP21 found in plastids are good candidates for PMSR substrates. This mechanism may also form part of the ROS sensing signaling process that gives rise to peroxisomal biogenesis. Taken together, the overall goal would be to improve the interactions shown in Figure 11 so as to increase the speed and efficiency of the signaling pathways that give rise to the mobilization of antioxidant defense mechanisms. Thiol redox control is proposed to play an essential and central role in mediating plant cell antioxidant responses to the imposition of oxidative stress.

REFERENCES

1.

R. Alscher , J. Bower , and W. Zipfel . 1987. “The basis for different sensitivities of photosynthesis to SO2 in two cultivars of pea”. J. Exp. Bot 38:1199–108. Google Scholar

2.

R. G. Alscher 1989. “Biosynthesis and antioxidant function of glutathione in plants”. Physiol. Plant 77:11457–464. Google Scholar

3.

R. G. Alscher , J. L. Donahue , and C. L. Cramer . 1997. “Reactive oxygen species and antioxidants: Relationships in green cells”. Physiol. Plant 100:11224–233. Google Scholar

4.

R. G. Alscher , N. Erturk , and L. S. Heath . 2002. “Role of superoxide dismutases (SODs) in controlling oxidative stress in plants”. J. Exp. Biol 53:111331–1341. Google Scholar

5.

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 34:11129–135. Google Scholar

6.

A. M. Arisi , G. Cornic , L. Jouanin , and C. H. Foyer . 1998. “Overexpression of iron superoxide dismutase in transformed poplar modifies the regulation of photosynthesis at low CO2 partial pressures or following exposure to the prooxidant herbicide methyl viologen”. Plant Physiol 117/2:11565–74. Google Scholar

7.

K. Asada 1994. “Production and action of active oxygen species in photosynthetic tissue. ” In Causes of Photooxidative Stress and Amelioration of Defense Systems in Plants. (C.H. Foyer and P.M. Mullineaux, eds), pp. 1177–104. Boca Raton, CRC Press. Google Scholar

8.

K. Asada and K. Kiso . 1973. “Initiation of aerobic oxidation of sulfite by illuminated spinach chloroplasts”. Eur. J. Biochem 33:11253–257. Google Scholar

9.

K. Asada , M. Urano , and M. Takahashi . 1973. “Subcellular location of superoxide dismutase in spinach leaves and preparation and properties of crystalline spinach superoxide dismutase”. Eur. J. Biochem 36:11257–266. Google Scholar

10.

S. Baginsky , K. Tiller , T. Pfannschmidt , and G. Link . 1999. “PTK, the chloroplast RNA polymerase-associated protein kinase from mustard (Sinapis alba), mediates redox control of plastid in vitro transcription”. Plant Mol. Biol 39/5:111013–23. Google Scholar

11.

C. Bailly , A. Benamaur , F. Corbineau , and D. Come . 1996. “Changes in malondialdehyde content and in superoxide dismutase, catalase and glutathione reductase activities in sunflower seeds as related to deterioration during accelerated aging”. Physiol. Plant 97:11104–110. Google Scholar

12.

L. Ball , O. Richard , U. Bechtold , C. Penkett , H. Reynolds , B. Kular , G. Creissen , S. Karpinski , W. Schuch , and P. Mullineaux . 2001. “Changes in Global Gene Expression in Response to Excess Excitation Energy in Arabidopsis thaliana”. J. Exp. Bot 52.(Special Issue). Google Scholar

13.

W. H. Bannister , J. V. Bannister , D. Barra , J. Bond , and F. Bossa . 1991. “Evolutionary aspects of superoxide dismutase: The copper/zinc enzyme”. Free Radic. Res. Commun 12-13:(Pt 1). 11349–61. Google Scholar

14.

H. Bohnert and E. Shevelava . 1998. “Plant Stress Adaptations- making metabolism move”. Curr. Opin. Plant Biol 3:11267–274. Google Scholar

15.

H. J. Bohnert and R. G. Jensen . 1996. “Strategies for engineering water-stress tolerance in plants”. Trends Biotech 14:1189–97. Google Scholar

16.

C. Bowler , L. Slooten , S. Vandenbranden , R. De Rycke , J. Botterman , C. Sybesma , M. Van Montagu , and D. Inze . 1991. “Manganese superoxide dismutase can reduce cellular damage mediated by oxygen radicals in transgenic plants”. EMBO J 10/7:111723–32. Google Scholar

17.

C. Bowler , W. Van Camp , M. Van Montagu , and D. Inze . 1994. “Superoxide dismutases in plants”. Crit. Rev. Plant Sci 13:11199–218. Google Scholar

18.

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 8:11247–255. Google Scholar

19.

N. Brot and H. Weissbach . 2000. “Peptide methionine sulfoxide reductase: biochemistry and physiological role”. Biopolymers 55/4:11288–96. Google Scholar

20.

L. M. Casano , H. R. Lascano , and V. S. Trippi . 1994. “Hydroxyl radicals and a thylakoid-bound endopeptidase are involved in light and oxygen-induced proteolysis in oat chloroplasts”. Plant Cell Physiol 35:11145–152. Google Scholar

21.

J. Chaudiere and R. Ferrari-Iliou . 1999. “Intracellular antioxidants: from chemical to biochemical mechanisms”. Food Chem. Toxicol 37:9-10. 11949–62. Google Scholar

22.

H. Choi , J. Hong , J. Ha , J. Kang , and S. Y. Kim . 2000. “ABFs, a family of ABA-responsive element binding factors”. J. Biol. Chem 275/3:111723–30. Google Scholar

23.

F. J. Corpas , J. B. Barroso , and L. A. del Rio . 2001. “Peroxisomes as a source of reactive oxygen species and nitric oxide signal molecules in plant cells”. Trends Plant Sci 6/4:11145–50. Google Scholar

24.

G. Creissen , P. Broadbent , R. Stevens , A. R. Wellburn , and P. Mullineaux . 1995a. “Manipulation of glutathione metabolism in transgenic plants”. Biochem. Soc. Transac 24:11465–469. Google Scholar

25.

G. Creissen , E. A. Edwards , C. Enard , A. Wellburn , and P. Mullineaux . 1991. “Molecular characterization of glutathione reductase cDNAs from pea (Pisum sativum L).”. Plant J 2:11129–131. Google Scholar

26.

G. Creissen , H. Reynolds , Y. Xue , and P. Mullineaux . 1995b. “Simultaneous targeting of pea glutathione reductase and of a bacterial fusion protein to chloroplasts and mitochondria in transgenic tobacco”. Plant J 8:11167–175. Google Scholar

27.

J. C. Cushman and H. J. Bohnert . 2000. “Genomic approaches to plant stress tolerance”. Curr. Opin. Plant Biol 3:11117–24. Google Scholar

28.

K. J. A. Davies 1987. “Protein damage and degradation by oxygen radicals”. J. Biol. Chem 262:119895–9901. Google Scholar

29.

M. De Paula , M. Perez-Otaola , M. Darder , M. Torres , G. Frutos , and C. J. Martinez-Honduvilla . 1996. “Function of the ascorbate-glutathione cycle in aged sunflower seeds”. Physiol. Plant 96:11543–555. Google Scholar

30.

C. H. R. De Vos , H. L. Kraak , and R. L. Bino . 1994. “Aging of tomato seeds involves glutathione oxidation”. Physiol. Plant 92:11131–139. Google Scholar

31.

L. A. del Rio , L. M. Sandalio , J. M. Palma , P. Bueno , and F. J. Corpas . 1992. “Metabolism of oxygen radicals in peroxisomed and cellular implications”. Free Rad. Biol. Med 13:11557–80. Google Scholar

32.

J. Diefenbach and H. Kindl . 2000. “The membrane-bound DnaJ protein located at the cytosolic site of glyoxysomes specifically binds the cytosolic isoform 1 of Hsp70 but not other Hsp70 species”. Eur. J. Biochem 267/3:11746–54. Google Scholar

33.

J. L. Donahue , C. M. Okpodu , C. L. Cramer , E. A. Grabau , and R. G. Alscher . 1997. “Responses of antioxidants to paraquat in pea leaves”. Plant Physiol 113:11249–57. Google Scholar

34.

E. F. Elstner 1991. “Mechanisms of oxygen activation in different compartments of plant cells. ” In Active Oxygen/Oxidative Stress and Plant Metabolism. (E.J. Pell and K.L. Steffen, eds), pp. 1113–25. Rockville, MD, American Society of Plant Physiologists Press. Google Scholar

35.

C. H. Foyer 1993. “Ascorbic acid”. In antioxidants in higher plants (R.G. Alscher and J.L. Hess, eds), pp. 1131–58. CRC Press, Boca Raton, Fl. Google Scholar

36.

C. H. Foyer and B. Halliwell . 1976. “The presence of glutathione and glutathione reductase in chloroplasts: A proposed role in ascorbic acid metabolism”. Planta 133:1121–25. Google Scholar

37.

C. H. Foyer and P. Mullineaux . Eds. 1994. Causes of Photooxidative Stress and Amelioration of Defense Systems in Plants. Boca Raton, CRC Press. Google Scholar

38.

C. H. Foyer , N. Sourian , 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 109:111047–1057. Google Scholar

39.

I. Fridovich 1986. “Superoxide dismutases”. Adv. Enzymol. Relat. Areas Mol. Biol 58:1161–97. Google Scholar

40.

C. Garrido , A. Fromentin , B. Bonnotte , N. Favre , M. Moutet , A. P. Arrigo , P. Mehlen , and E. Solary . 1998. “Heat shock protein 27 enhances the tumorigenicity of immunogenic rat colon carcinoma cell clones”. Cancer Res 58/23:115495–9. Google Scholar

41.

J. M. Gomez , J. A. Hernandez , A. Jimenez , L. A. del Rio , and F. Sevilla . 1999. “Differential response of antioxidative enzymes of chloroplasts and mitochondria to long-term NaCl stress of pea plants”. Free Radic. Res 31:Suppl. 11S11–8. Google Scholar

42.

A. Gonzalez , K. L. Steffen , and J. P. Lynch . 1998. “Light and excess manganese. Implications for oxidative stress in common bean.”. Plant Physiol 118/2:11493–504. Google Scholar

43.

S. Grace , R. Pace , and T. Wydrzynski . 1995. “Formation and decay of monodehydroascorbate radicals in illuminated thylakoids as determined by EPR spectroscopy”. Biochim. Biophys. Acta 1229:11155–165. Google Scholar

44.

J. Gressel and E. Galun . 1994. Genetic control of photooxidant tolerance. ” In causes of photooxidative stress and amelioration of defense systems in plants. (C. Foyer and P. Mullineaux, eds), pp. 11237–273. Boca Raton, CRC Press. Google Scholar

45.

L. Guan and J. G. Scandalios . 1998. “Two structurally similar maize cytosolic superoxide dismutase genes, Sod4 and Sod4A, respond differentially to abscisic acid and high osmoticum”. Plant Physiol 117/1:11217–24. Google Scholar

46.

N. Gustavsson , U. Harndahl , A. Emanuelsson , P. Roepstorff , and C. Sundby . 1999. “Methionine sulfoxidation of the chloroplast small heat shock protein and conformational changes in the oligomer”. Protein Sci 8/11:112506–12. Google Scholar

47.

H. Hagar , N. Ueda , and S. V. Shah . 1996. “Role of reactive oxygen metabolites in DNA damage and cell death in chemical hypoxic injury to LLC-PK1 cells”. Am. J. Physiol 271:11209–215. Google Scholar

48.

B. Halliwell and J. M. C. Gutteridge . 1989. Free Radicals in Biology and Medicine. Oxford, Clarendon Press. Google Scholar

49.

M. Hamberg 1999. “An epoxy alcohol synthase pathway in higher plants: biosynthesis of antifungal trihydroxy oxylipins in leaves of potato”. Lipids 34/11:111131–42. Google Scholar

50.

U. Harndahl , R. B. Hall , K. W. Osteryoung , E. Vierling , J. F. Bornman , and C. Sundby . 1999. “The chloroplast small heat shock protein undergoes oxidation-dependent conformational changes and may protect plants from oxidative stress”. Cell Stress Chaperones 4/2:11129–38. Google Scholar

51.

U. Harndahl , B. P. Kokke , N. Gustavsson , S. Linse , K. Berggren , F. Tjerneld , W. C. Boelens , and C. Sundby . 2001. “The chaperone-like activity of a small heat shock protein is lost after sulfoxidation of conserved methionines in a surface-exposed amphipathic alpha-helix”. Biochim. Biophys. Acta 1545:1-2. 11227–37. Google Scholar

52.

C. Haslekas , R. Stacy , V. Nygaard , F. Culianez-Macia , and R. Aalen . 1998. “The expression of a periredoxin antioxidant gene, AtPer1, in Arabidopsis thaliana is seed specific and related to dormancy”. Plant Mol. Biol 36:11833–845. Google Scholar

53.

A. Hausladen and R. Alscher . 1994. “Purification and characterization of glutathione reductase isozymes specific for the state of cold hardiness of red spruce”. Plant Physiol 105:11205–213. Google Scholar

54.

A. Hausladen and R. G. Alscher . 1993. “Glutathione. ” In antioxidants in higher plants. (R.G. Alscher and J.L. Hess, eds) pp. 111–30. CRC Press. Google Scholar

55.

L. S. Heath , N. Ramakrishnan , R. R. Sederoff , R. W. Whetten , B. I. Chevone , C. A. Struble , V. Y. Jouenne , D. Chen , L. van Zyl , and R. Grene . 2002. “ Studying the functional genomics of stress responses in loblolly pine with the Expresso microarray experiment management system”. Comp. Funct. Genom 3:11226–243. Google Scholar

56.

J. A. Hernandez , F. J. Corpas , M. Gomez , L. A. del Rio , and F. Sevilla . 1993. “Salt-induced oxidative stress mediated by activated oxygen species in pea leaf mitochondria”. Physiol. Plant 89:11103–10. Google Scholar

57.

J. A. Hernandez , E. Olmos , F. J. Corpas , F. Sevilla , and L. A. del Rio . 1995. “Salt-induced oxidative stress in chloroplasts of pea plants”. Plant Sci 105:11151–67. Google Scholar

58.

D. Herouart , M. Van Montagu , and D. Inze . 1993. “Redox-activated expression of the cytosolic copper/zinc superoxide dismutase gene in Nicotiana”. Proc. Natl. Acad. Sci. USA 90:113108–3112. Google Scholar

59.

J. L. Hess 1993. “Vitamin E, alpha-tocopherol. ” In antioxidants in Higher Plants. (R.G. Alscher and J.L. Hess, eds) pp. 11111–134. Boca Raton, FL, CRC. Google Scholar

60.

E. H. Hettema , C. C. Ruigrok , M. G. Koerkamp , M. van den Berg , H. F. Tabak , B. Distel , and I. Braakman . 1998. “The cytosolic DnaJ-like protein djp1p is involved specifically in peroxisomal protein import”. J. Cell Biol 142/2:11421–34. Google Scholar

61.

T. Hoshi and S. Heinemann . 2001. “Regulation of cell function by methionine oxidation and reduction”. J. Physiol 531:(Pt 1). 111–11. Google Scholar

62.

S. Kanematsu and K. Asada . 1979. “Ferric and manganic superoxide dismutases in Euglena gracilis”. Arch. Biochem. Biophys 195/2:11535–45. Google Scholar

63.

S. Kanematsu and K. Asada . 1990. “Characteristic amino acid sequences of chloroplast and cytosol isozymes of CuZn-superoxide dismutase in spinach, rice and horsetail”. Plant Cell Physiol 31:1199–112. Google Scholar

64.

B. Karpinska , G. Wingsle , and S. Karpinski . 2000. “Antagonistic effects of hydrogen peroxide and glutathione on acclimation to excess excitation energy in Arabidopsis”. IUBMB Life 50/1:1121–6. Google Scholar

65.

S. Karpinski , C. Escobar , B. Karpinska , G. Creissen , and P. M. Mullineaux . 1997. “Photosynthetic electron transport regulates the expression of cytosolic ascorbate peroxidase genes in Arabidopsis during excess light stress”. Plant Cell 9:11627–640. Google Scholar

66.

S. Karpinski , H. Reynolds , B. Karpinska , G. Wingsle , G. Creissen , and P. Mullineaux . 1999. “Systemic signaling and acclimation in response to excess excitation energy in Arabidopsis”. Science 284/5414:11654–7. Google Scholar

67.

D. J. Kliebenstein , R. Monde , and R. L. Last . 1998. “Superoxide dismutase in Arabidopsis: An eclectic enzyme family with disparate regulation and protein localization”. Plant Physiol 118/2:11637–50. Google Scholar

68.

Y. Kovtun , W. L. Chiu , G. Tena , and J. Sheen . 2000. “Functional analysis of oxidative stress-activated mitogen-activated protein kinase cascade in plants”. Proc. Natl. Acad. Sci. USA 97/6:112940–5. Google Scholar

69.

G. J. Lee and E. Vierling . 1998. “Expression, purification, and molecular chaperone activity of plant recombinant small heat shock proteins”. Methods Enzymol 290:11350–65. Google Scholar

70.

G. J. Lee and E. Vierling . 2000. “A small heat shock protein cooperates with heat shock protein 70 systems to reactivate a heat-denatured protein”. Plant Physiol 122/1:11189–98. Google Scholar

71.

H. Lee , L. Xiong , M. Ishitani , B. Stevenson , and J. K. Zhu . 1999. “Cold-regulated gene expression and freezing tolerance in an Arabidopsis thaliana mutant”. Plant J 17/3:11301–8. Google Scholar

72.

Y. B. Lee and S. M. Lee . 2000. “Effect of S-adenosylmethionine on hepatic injury from sequential cold and warm ischemia”. Arch. Pharm. Res 23/5:11495–500. Google Scholar

73.

O. Leprince , P. C. Thorpe , R. Deltour , N. M. Atherton , and G. A. F. Hendry . 1990. “The role of free radicals and radical processing systems in loss of desiccation tolerance in germinating maize”. New Phytol 116:11573–580. Google Scholar

74.

R. L. Levine , B. S. Berlett , J. Moskovitz , L. Mosoni , and E. R. Stadtman . 1999. “Methionine residues may protect proteins from critical oxidative damage”. Mech. Ageing Dev 107/3:11323–32. Google Scholar

75.

E. Lopez-Huertas , W. L. Charlton , B. Johnson , I. A. Graham , and A. Baker . 2000. “Stress induces peroxisome biogenesis genes”. EMBO J 19/24:116770–7. Google Scholar

76.

W. T. Lowther , N. Brot , H. Weissbach , and B. W. Matthews . 2000. “Structure and mechanism of peptide methionine sulfoxide reductase, an “anti-oxidation” enzyme”. Biochemistry 39/44:1113307–12. Google Scholar

77.

N. Madamanchi , R. Alscher , K. Hatzios , and C. Cramer . 1994. “Acquired resistance to herbicides in pea cultivars by exposure to sulfur dioxide”. Pest. Biochem. Physiol 48:1131–40. Google Scholar

78.

N. Madamanchi , J. Anderson , R. Alscher , C. Cramer , and J. Hess . 1992. “Purification of multiple forms of glutathione reductase from pea (Pisum sativum L) seedlings and enzyme levels in ozone fumigated pea leaves.”. Plant Physiol 100:11138–145. Google Scholar

79.

M. J. May and C. J. Leaver . 1993. “Oxidative stimulation of glutathione synthesis in Arabidopsis thaliana suspension cultures”. Plant Physiol 103:11621–627. Google Scholar

80.

M. J. May , T. Vernoux , C. Leaver , M. Van Montagu , and D. Inze . 1998a. “Glutathione homeostasis in plants: implications for environmental sensing and plant development”. J. Exp. Bot 49:11649–667. Google Scholar

81.

M. J. May , T. Vernoux , R. Sanchez-Fernandez , M. Van Montagu , and D. Inze . 1998b. “Evidence for posttranscriptional activation of gamma-glutamylcysteine synthetase during plant stress responses”. Proc. Natl. Acad. Sci. USA 95/20:1112049–54. Google Scholar

82.

P. Mehlen , C. Kretz-Remy , X. Preville , and A. P. Arrigo . 1996. “Human HSP27, Drosophila HSP27 and human alphaB-crystallin expression-mediated increase in glutathione is essential for the protective activity of these proteins against TNFalpha-induced cell death”. EMBO J 15/11:112695–706. Google Scholar

83.

H. Mehlhorn , B. J. Tabner , and A. R. Wellburn . 1990. “Electron spin resonance evidence for the formation of free radicals in plants exposed to ozone”. Physiol. Plant 79:11377–383. Google Scholar

84.

R. Mittler and B. Zilinskas . 1994. “Regulation of pea cytosolic ascorbate peroxidase and other antioxidant enzymes during the progression of drought stress and following recovery from drought”. Plant J 5:11397–405. Google Scholar

85.

J. Moskovitz , B. S. Berlett , J. M. Poston , and E. R. Stadtman . 1997. “The yeast peptide-methionine sulfoxide reductase functions as an antioxidant in vivo”. Proc. Natl. Acad. Sci. USA 94/18:119585–9. Google Scholar

86.

R. T. Mullen and R. N. Trelease . 2000. “The sorting signals for peroxisomal membrane-bound ascorbate peroxidase are within its C-terminal tail”. J. Biol. Chem 275/21:1116337–44. Google Scholar

87.

P. Mullineaux , L. Ball , C. Escobar , B. Karpinska , G. Creissen , and S. Karpinski . 2000. “Are diverse signalling pathways integrated in the regulation of arabidopsis antioxidant defence gene expression in response to excess excitation energy?” Philos. Trans. R. Soc. Lond. B. Biol. Sci 355/1402:111531–40. Google Scholar

88.

Y. Nakano and K. Asada . 1980. “Spinach chloroplasts scavenge hydrogen peroxide on illumination”. Plant Cell Physiol 21:111295–307. Google Scholar

89.

S. J. Neill , R. Desikan , A. Clarke , R. Hurst , and J. T. Hancock . 2001. “Hydrogen Peroxide and Nitric Oxide as Signalling Molecules in Plants”. J. Exp. Bot 52.(Special Issue). Google Scholar

90.

G. Noctor , A-C. Arisi , L. Joanin , K. Kunert , H. Rennenberg , and C. Foyer . 1998. “Glutathione: biosynthesis, metabolism and relationship to stress tolerance explored in transformed plants”. J. Exp. Bot 49:11623–647. Google Scholar

91.

G. Noctor and C. H. Foyer . 1998. “Ascorbate and Glutathione: Keeping active oxygen under control”. Ann. Rev. Plant Physiol. Plant Mol. Biol 49:11249–279. Google Scholar

92.

K. Ogawa , S. Kanematsu , and K. Asada . 1996. “Intra- and extra-cellular localization of “cytosolic” CuZn superoxide dismutase in spinach leaf and hypocotyl”. Plant Cell Physiol 37:11790–9. Google Scholar

93.

K. Ogawa , S. Kanematsu , and K. Asada . 1997. “Generation of superoxide anion and localization of CuZn superoxide dismutase in the vascular tissue of spinach hypocotyls: Their association with lignification”. Plant Cell Physiol 38:111118–26. Google Scholar

94.

K. Ogawa , S. Kanematsu , K. Takebe , and K. Asada . 1995. “Attachment of Cu, Zn SOD to thylakoid membranes at the site of superoxide generation (PS1) in spinach chloroplasts: Detection by immuno-gold labeling after rapid freezing”. Plant Cell Physiol 36:11565–73. Google Scholar

95.

C. M. Okpodu , R. G. Alscher , E. A. Grabau , and C. L. Cramer . 1996. “Physiological, Biochemical and Molecular Effects of Sulfur Dioxide”. J. Plant Physiol 148:11309–316. Google Scholar

96.

G. M. Pastori and C. H. Foyer . 2001. “Identifying Oxidative Stress Responsive Genes by Transposon Tagging”. J. Exp. Bot 52.(Special Issue). Google Scholar

97.

A. Perl , R. Perl-Treves , S. Galili , D. Aviv , E. Shalgi , S. Malkin , and E. Galun . 1993. “Enhanced oxidative stress defence in transgenic potato expressing tomato Cu,Zn superoxide dismutases”. Theor. Appl. Genet 85:11568–76. Google Scholar

98.

T. Pfannschmidt , A. Nilsson , A. Tullberg , G. Link , and J. F. Allen . 1999. “Direct transcriptional control of the chloroplast genes psbA and psaAB adjusts photosynthesis to light energy distribution in plants”. IUBMB Life 48/3:11271–6. Google Scholar

99.

T. K. Prasad , M. D. Anderson , and C. R. Stewart . 1994. “Acclimation, hydrogen peroxide, and abscisic acid protect mitochondria against irreversible chilling injury in maize seedlings”. Plant Physiol 105:11619–627. Google Scholar

100.

D. S. Prestridge 2000. “Computer software for eukaryotic promoter analysis”. Methods Mol. Biol 130:11265–95. Google Scholar

101.

X. Preville , F. Salvemini , S. Giraud , S. Chaufour , C. Paul , G. Stepien , M. V. Ursini , and A. P. Arrigo . 1999. “Mammalian small stress proteins protect against oxidative stress through their ability to increase glucose-6-phosphate dehydrogenase activity and by maintaining optimal cellular detoxifying machinery”. Exp. Cell Res 247/1:1161–78. Google Scholar

102.

V. P. Roxas , J. Smith , K. Roger , E. R. Allen , and R. D. Allen . 1997. “Overexpression of glutathione S-transferase/glutathione peroxidase enhances the growth of transgenic tobacco seedlings during stress”. Nature Biotech 15:11988–991. Google Scholar

103.

A. Sadanandom , Z. Poghosyan , D. J. Fairbairn , and D. J. Murphy . 2000. “Differential regulation of plastidial and cytosolic isoforms of peptide methionine sulfoxide reductase in Arabidopsis”. Plant Physiol 123/1:11255–64. Google Scholar

104.

R. Sanchez-Fernandez , M. Fricker , L. B. Corben , N. S. White , N. Sheard , C. J. Leaver , and M. Van Montagu . 1997. “Cell proliferation and hair tip growth in the Arabidopsis root are under mechanistically different forms of redox control”. Proc. Natl. Acad. Sci. USA 94:112745–2750. Google Scholar

105.

M. Santos , H. Gousseau , C. Lister , C. Foyer , G. Creissen , and P. Mullineaux . 1996. “Cytosolic ascorbate peroxidase from Arabidopsis thaliana L is encoded by a small multigene family.”. Planta 198/1:1164–9. Google Scholar

106.

J. G. Scandalios 1997. “Molecular Genetics of Superoxide Dismutases in Plants. ” In Oxidative Stress and the Molecular Biology of Antioxidative Defenses. (J.G. Scandalios, eds) pp. 11527–568. Plainview, Cold Spring Harbor. Google Scholar

107.

A. Sen Gupta , J. L. Heinen , A. S. Holaday , J. J. Burke , and R. D. Allen . 1993. “Increased resistance to oxidative stress in transgenic plants that overexpress chloroplastic Cu/Zn superoxide dismutase”. Proc. Natl. Acad. Sci. USA 90:111629–1633. Google Scholar

108.

L. Slooten , K. Capiau , W. Van Camp , M. Van Montagu , C. Subesma , and D. Inze . 1995. “Factors affecting the enhancements of oxidative stress tolerance in transgenic tobacco overexpressing manganese superoxide dismutase in the chloroplasts”. Plant Physiol 107:11737–50. Google Scholar

109.

A. D. Smith , S. P. Datta , G. H. Smith , P. N. Campbell , R. Bentley , H. A. McKenzie , D. A. Bender , A. J. Harris , T. W. Goodwin , J. H. Parish , and C. Stanford . Eds. 2000. Oxford Dictionary of Biochemistry and Molecular Biology (Revised Edition). Oxford, Oxford University Press. Google Scholar

110.

M. W. Smith and R. F. Doolittle . 1992. “A comparison of evolutionary rates of the two major kinds of superoxide dismutases”. J. Mol. Evol 34:11175–84. Google Scholar

111.

T. C. Squier and D. J. Bigelow . 2000. “Protein oxidation and age-dependent alterations in calcium homeostasis”. Front. Biosci 5:11D504–26. Google Scholar

112.

R. G. Stevens , G. P. Creissen , and P. M. Mullineaux . 2000. “Characterisation of pea cytosolic glutathione reductase expressed in transgenic tobacco”. Planta 211/4:11537–45. Google Scholar

113.

M. A. Takahashi and K. Asada . 1983. “Superoxide anion permeability of phospholipid membranes and chloroplast thylakoids”. Arch. Biochem. Biophys 226/2:11558–66. Google Scholar

114.

J. M. Tepperman and P. Dunsmuir . 1990. “Transformed plants with elevated levels of chloroplastic SOD are not more resistant to superoxide toxicity”. Plant Mol. Biol 14/4:11501–11. Google Scholar

115.

M. B. Toledano and W. J. Leonard . 1991. “Modulation of transcription factor NF-kappa B binding activity by oxidation-reduction in vitro”. Proc. Natl. Acad. Sci. USA 88:114328–4332. Google Scholar

116.

W. Van Camp , K. Capiau , M. Van Montagu , D. Inze , and L. Slooten . 1996. “Enhancement of oxidative stress tolerance in transgenic tobacco plants overproducing Fe-superoxide dismutase in chloroplasts”. Plant Physiol 112:111703–14. Google Scholar

117.

W. Van Camp , H. Willekens , C. Bowler , M. Van Montagu , D. Inze , C. Langebartels , and H. Sandermann . 1994. “Elevated levels of superoxide dismutase protect transgenic plants against ozone damage”. Bio/Technology 12:11165–8. Google Scholar

118.

S. Van Nocker , Q. Deveraux , M. Rechsteiner , and R. D. Vierstra . 1996. “Arabidopsis MBP1 gene encodes a conserved ubiquitin recognition component of the 26S proteasome”. Proc. Natl. Acad. Sci. USA 93:11856–860. Google Scholar

119.

H. Vanacker , T. L. Carver , and C. H. Foyer . 2000. “Early H(2) O(2) accumulation in mesophyll cells leads to induction of glutathione during the hyper-sensitive response in the barley-powdery mildew interaction”. Plant Physiol 123/4:111289–300. Google Scholar

120.

C. W. Vertucci and J. M. Farrant . 1995. “Acquisition and loss of desiccation tolerance. ” In Seed Development and Germination. (J. Kigal and G.Galili, eds) pp. 11237–271. New York, Marcel Dekker. Google Scholar

121.

C. Walters 1998. “Understanding the mechanisms and kinetics of seed aging”. Seed Sci. Res 8:11223–244. Google Scholar

122.

N. Wehmeyer and E. Vierling . 2000. “The expression of small heat shock proteins in seeds responds to discrete developmental signals and suggests a general protective role in desiccation tolerance”. Plant Physiol 122/4:111099–108. Google Scholar

123.

J. D. Williamson and J. G. Scandalios . 1992. “Differential response of maize catalases and superoxide dismutases to the photoactivated fungal toxin cercosporin”. Plant J 2:11351–358. Google Scholar

124.

B. Wimmer , F. Lottspeich , I. van der Klei , M. Veenhuis , and C. Gietl . 1997. “The glyoxysomal and plastid molecular chaperones (70-kDa heat shock protein) of watermelon cotyledons are encoded by a single gene”. Proc Natl Acad Sci USA 94/25:1113624–9. Google Scholar

125.

B. S. Winkler , S. M. Orselli , and T. S. Rex . 1994. “The Redox Couple between Glutathione and Ascorbic Acid: A Chemical and Physiological Perspective”. Free Radic. Biol. Med 17:11333–349. Google Scholar

126.

G. Wu , R. W. Wilen , A. J. Robertson , and L. V. Gusta . 1999. “Isolation, chromosomal localization, and differential expression of mitochondrial manganese superoxide dismutase and chloroplastic copper/zinc superoxide dismutase genes in wheat”. Plant Physiol 120/2:11513–20. Google Scholar

127.

C. Xiang and D. Oliver . 1998. “Glutathione metabolic genes coordinately respond to heavy metals and jasmonic acid in Arabidopsis”. Plant Cell 10:111539–1550. Google Scholar

128.

I. Yruela , J. J. Pueyo , P. J. Alonso , and R. Picorel . 1996. “Photoinhibition of photosystem II from higher plants. Effect of copper inhibition.”. J. Biol. Chem 271:1127408–27415. Google Scholar

129.

Y. L. Zhu , E. A. H. Philon-Smits , A. S. Tarun , S. U. Weber , L. Jouanin , and N. Terry . 1999. “Cadmium tolerance and accumulation in Indian mustard is enhanced by overexpressing gamma-glutamylcysteine synthetase”. Plant Physiol 121:111169–1178. Google Scholar

Figure 1.

Thiol redox control and stress defense

i1543-8120-49-1-1-f01.gif

Figure 2.

Reactive oxygen species (ROS) arise throughout the cell.

i1543-8120-49-1-1-f02.jpg

Figure 3.

The scavenging of active oxygen species in the chloroplast in both the lipid membrane phase and the stroma, linked to redox cycles for ascorbate and glutathione and the oxidation of α-tocopherol (Vitamin E). Abbreviations are as follows: P-LIPID -OO, phospholipid peroxy radical; P-LIPID-OOH, phospholipid peroxide; P-LIPID-OH, phospholipid alcohol; VIT-E(OH), α-tocopherol (vitamin E); VIT-E (O*), α-chromanoxyl radical; PHGPX, phospholipid hydroperoxide-dependent glutathione peroxidase; GSH, reduced glutathione; GSSG, glutathione disulfide; GR, glutathione reductase; DHAR, dehydroascorbate reductase; ASC, ascorbic acid; DHA, dehydroascorbate; MDA, monodehydroascorbate free radical; MDAR, monodehydroascorbate free radical reductase; APX, ascorbate reductase; SOD, superoxide dismutase; O2-, superoxide ion. Reaction 1 is the non-enzyme-catalyzed spontaneous dismutation of two MDA molecules to one ASC and one DHA, respectively. From Mullineaux et al. (2000).

i1543-8120-49-1-1-f03.gif

Figure 4.

Cellular locations of superoxide dismutase (SOD).

i1543-8120-49-1-1-f04.jpg

Figure 5.

Relatedness of SOD protein sequences in some plant cells. Tree was obtained using MegAlign in DNA Star.

i1543-8120-49-1-1-f05.gif

Figure 6.

A comparison of cytosolic, chloroplast, and peroxisomal Cu-Zn SODs from higher plants. Abbreviations are as follows: cyt, cytosolic; cp, chloroplast; prx, peroxisomal; ec, extracellular. From Kliebenstein et al. (1998).

i1543-8120-49-1-1-f06.jpg

Figure 7.

The Mehler peroxidase reaction (from Mullineaux et al. 2000). The APX-catalyzed reaction is the reduction of H2O2 to water using ascorbate as the electron donor. The oxidized ascorbate, monodehydroascorbate free radical or its disproportionation product dehydroascorbate, are reduced back to ascorbate using two electrons ultimately derived from PS II.

i1543-8120-49-1-1-f07.gif

Figure 8.

Oxidation and reduction of methionine residues as an antioxidant reservoir. (Adapted from Hoshi and Heinemann, 2001.) *Cytosolic and chloroplast forms of PMSR are known in plants. The thioredoxin activation system in the chloroplast is well known as a source of reductant for redox regulated stromal enzymes.

i1543-8120-49-1-1-f08.gif

Figure 9.

Redox control of gene expression in green cells. The Mehler peroxidase reaction, which generates superoxide from O2 at PS I, is one possible source of ROS which, along with the PQ-associated redox sensor, regulates antioxidant defense gene expression. (From Mullineaux et al. 2000)

i1543-8120-49-1-1-f09.gif

Figure 10.

Systemic acquired acclimation. Luciferase expression in a detached wounded leaf from a five-week-old short-day-grown Arabidopsis rosette transgenic for an APX2-promoter-LUC fusion. Leaves were detached from the plant at the petiole and were subjected to a series of parallel slashes in the lowest quarter of the leaf. EL was applied for 40 min to the top one-quarter of the leaf, the remainder being shaded with aluminum foil. At the end of the light stress period, the leaves were sprayed with luciferin and 30 min later an image was recorded using a charged couple device camera. Abbreviations are as follows: CW, control wounded leaf; EL, excess light; LL, low light; LUC, luciferase. (From Karpinski et al. 1999)

i1543-8120-49-1-1-f10.gif

Figure 11.

Redox regulation of gene expression. (1), (2) and (3) can function as redox sensors. (Adapted from Hoshi and Heinemann 2001)

i1543-8120-49-1-1-f11.gif

Table 1.

A comparison of the upstream regions of the seven known Arabidopsis SODs using the identification of sequences known to bind to four transcription factors. (From Alscher et al. 2002)

i1543-8120-49-1-1-t01.gif

[1] Every molecule of Mn and Fe SOD contains either an atom of manganese or an atom of iron depending on the species or the availability of the metal in archaic SODs. The potential use of either Fe or Mn is denoted as such by a slash, Mn/Fe.

[2] Every molecule of the Cu-Zn SOD enzyme contains both an atom of copper and an atom of zinc, as denoted by the hyphen.

The American Society of Plant Biologists
Ruth Grene "Oxidative Stress and Acclimation Mechanisms in Plants," The Arabidopsis Book 2002(1), (4 April 2002). https://doi.org/10.1199/tab.0036.1
Published: 4 April 2002
Back to Top