Skip to main content
EMBO Reports logoLink to EMBO Reports
. 2008 May;9(5):435–439. doi: 10.1038/embor.2008.57

No single way to understand singlet oxygen signalling in plants

Chanhong Kim 1,2, Rasa Meskauskiene 1, Klaus Apel 2,a,1, Christophe Laloi 1
PMCID: PMC2373375  PMID: 18451767

Abstract

When plant cells are under environmental stress, several chemically distinct reactive oxygen species (ROS) are generated simultaneously in various intracellular compartments and these can cause oxidative damage or act as signals. The conditional flu mutant of Arabidopsis, which generates singlet oxygen in plastids during a dark-to-light transition, has allowed the biological activity of singlet oxygen to be determined, and the criteria to distinguish between cytotoxicity and signalling of this particular ROS to be defined. The genetic basis of singlet-oxygen-mediated signalling has been revealed by the mutation of two nuclear genes encoding the plastid proteins EXECUTER (EX)1 and EX2, which are sufficient to abrogate singlet-oxygen-dependent stress responses. Conversely, responses due to higher cytotoxic levels of singlet oxygen are not suppressed in the ex1/ex2 background. Whether singlet oxygen levels lower than those that trigger genetically controlled cell death activate acclimation is now under investigation.

Keywords: Arabidopsis, oxidative stress, singlet oxygen, flu mutant, executer

Introduction

Reactive oxygen species (ROS) are produced continuously as unavoidable by-products of aerobic metabolism (Halliwell, 2006). Oxygen in its triplet ground state is chemically inert but it can be converted to a ROS either by electron transfer—leading to, by stepwise reduction, the formation of the superoxide radical (O2•−), hydrogen peroxide (H2O2) and hydroxyl radical (HO)—or energy transfer—which leads to the formation of singlet oxygen (1O2; Foote, 1968; Gollnick, 1968). During evolution, the biological activities of ROS seem to have undergone several modifications. The continuous release of highly reactive ROS has necessitated the development of ROS scavengers to minimize the cytotoxic effects of ROS in the cell (Apel & Hirt, 2004). At the same time, the detection of rapid changes in ROS concentrations that result from metabolic disturbances or external factors has been used by cells to activate stress-related responses and to readjust their homeostasis (Gechev et al, 2006). Finally, the discovery of genetically controlled ROS production by enzymes such as NADPH oxidases has assigned a surprisingly large diversity of biological activities to ROS. These activities affect a range of processes, including the defence reactions of plants and vertebrates against pathogens (Sagi & Fluhr, 2006), the regulation of cell expansion and development (Foreman et al, 2003; Gapper & Dolan, 2006), and plant–fungus mutualistic interactions (Takemoto et al, 2006; Tanaka et al, 2006).

In plants, chloroplasts and peroxisomes are the main sites of ROS production (Asada, 2006; Foyer & Noctor, 2003). The enhanced generation of ROS in these cellular compartments has been attributed to the disturbance of light-driven photosynthetic electron transport by various environmental factors that trigger stress responses (Niyogi, 1999). Under these environmental conditions, plants are exposed to light intensities that exceed their capacity to assimilate CO2 and lead to the over-reduction of the electron-transport chain, and ultimately can result in the inhibition of photosynthesis. Plants use two strategies to protect their photosynthetic apparatus: first, the thermal dissipation of excess excitation energy in the photosystem II (PSII) antennae—known as non-photochemical quenching (Muller et al, 2001); and second, the transfer of electrons from PSII to alternative sinks, thereby sustaining partial oxidation of PSII acceptors and preventing photoinactivation of PSII—known as photochemical quenching. Photochemical quenching can be achieved in chloroplasts through, for example, the direct reduction of oxygen to O2•− by reduced electron-transport components associated with PSI (Asada, 1999; Rizhsky et al, 2003) and by reactions linked to the photorespiratory cycle that result in the enhanced production of H2O2 in peroxisomes (Kozaki & Takeba, 1996). 1O2 is continuously produced by PSII through energy transfer from excited chlorophyll to oxygen (Gollnick, 1968; Krieger-Liszkay, 2005). The quenching of 1O2 has been linked primarily to the turnover of the D1 protein of the PSII reaction centre (Telfer et al, 1994; Trebst, 2003), although tocopherols and carotenoids also participate in preserving PSII from photoinactivation (Havaux et al, 2005; Krieger-Liszkay & Trebst, 2006; Telfer, 2005).

Plant responses to singlet oxygen

One of the difficulties in studying the biological activity of a particular ROS originates from the fact that several chemically distinct ROS are generated simultaneously in cells under stress, therefore making it almost impossible to link a particular stress response to a specific ROS. However, to study the specific activity of 1O2, this problem has been overcome by using the conditional flu mutant of Arabidopsis, which allows the induction of only 1O2 within plastids in a non-invasive and controlled manner.

FLU is a nuclear-encoded protein that is tightly associated with plastid membranes. It regulates the Mg2+ branch of tetrapyrrole biosynthesis by interacting directly with glutamyl tRNA reductase (Meskauskiene et al, 2001; Meskauskiene & Apel, 2002). In contrast to the wild type, etiolated flu seedlings are no longer able to restrict the accumulation of protochlorophyllide (Pchlide)—the immediate precursor of chlorophyllide (Chlide)—which is a potent photosensitizer that generates 1O2 in the light (Fig 1, D; Meskauskiene et al, 2001; op den Camp et al, 2003). As a result, when these seedlings are transferred from the dark to the light they rapidly bleach and die (Fig 1, D→L; Meskauskiene et al, 2001). However, the flu mutant remains viable when it is grown from germination under continuous light (Fig 1, LL). Under these conditions, Pchlide is immediately photoreduced to Chlide and therefore does not accumulate. Provided the flu plants are kept under continuous light, no obvious differences are observed between them and the wild type. This conditional phenotype of the flu mutant has been exploited to study the physiological role of 1O2 by growing flu plants initially under continuous light—until they reach the developmental stage of interest—then transferring them to the dark and re-exposing them to light. The release of 1O2 in the flu mutant has been determined directly by using the 1O2-specific probe DanePy (Kalai et al, 1998; op den Camp et al, 2003). Seedlings of flu bleach and die when they are kept under repeated 16 h:8 h light:dark cycles (Fig 1, L/D), whereas mature plants grown initially under continuous light develop necrotic lesions on their leaves and stop their growth immediately after being moved to daily dark:light cycles (op den Camp et al, 2003).

Figure 1.

Figure 1

Cytotoxicity versus signalling of 1O2 in Arabidopsis seedlings. In etiolated flu and executer1/flu (ex1/flu) seedlings, similar excess amounts of free protochlorophyllide (Pchlide) accumulate that, on excitation with blue light, emit a strong red fluorescence (D). In pre-illuminated flu seedlings exposed to 16 h: 8 h light:dark cycles (L/D) inactivation of the EX1 protein abrogates bleaching. However, in etiolated flu seedlings, which accumulate four- to fivefold the amounts of free Pchlide in seedlings grown in continuous light and transferred to the dark for 8 h, the 1O2-mediated collapse during re-illumination is not suppressed by the ex1 mutation (D→L). All three plant lines grew equally well under continuous light (LL). Adapted from Przybyla et al (2008) and reprinted with permission from Wiley–Blackwell Publishing Ltd. Wt, wild type.

These stress responses could result either from physicochemical damage caused by 1O2 or by the activation of a genetically determined stress-response programme. Initially, we addressed this question by using experiments in which plants were grown under continuous light and then moved to one 8 h dark period before re-exposure to light.

Singlet-oxygen-responsive gene network

Under these experimental conditions, several global gene-expression studies have been carried out with flu plants using Affymetrix GeneChip arrays (Laloi et al, 2006, 2007; Lee et al, 2007; op den Camp et al, 2003). Within 30 min of the release of 1O2 in plastids, approximately 300 nuclear genes are rapidly upregulated at least threefold in flu plants relative to the wild type. On the basis of comparative transcriptome analyses, we identified groups of genes that have been shown previously to also be under the control of phytohormones, in particular, abscisic acid, ethylene, methyl jasmonate and salicylic acid (Danon et al, 2005; Ochsenbein et al, 2006). Previous attempts to modulate the triggering of stress responses in flu plants by using phytohormone-related mutants revealed that the ethylene, methyl jasmonate and salicylic acid pathways do have a role in mediating the cell-death response of flu (Danon et al, 2005; Ochsenbein et al, 2006). Furthermore, we identified genes that seem to be more strictly under the control of 1O2 (Gadjev et al, 2006; Laloi et al, 2006; op den Camp et al, 2003). In addition, approximately 50 genes encoding putative transcription factors showed a threefold increase in expression within 30 min of the release of 1O2. These include ethylene responsive factors, WRKY transcription factors, zinc-finger proteins and several DNA-binding proteins. Many genes involved in putative signal-transduction pathways and calcium regulation, such as protein kinases, calcium and calmodulin-binding proteins, were also identified (Danon et al, 2005; Laloi et al, 2006; Lee et al, 2007; op den Camp et al, 2003). We are now using genetic screens to identify modulators and constituents of the 1O2 signalling network. We have searched for 1O2-responsive and 1O2-specific promoters that, in combination with the luciferase (LUC) reporter gene, can be used in the flu background to monitor 1O2-dependent activation of nuclear genes. A LUC reporter gene line has been mutated with ethyl methane sulphonate (EMS), and second-site mutants with a constitutively upregulated LUC expression have been isolated. The preliminary characterization of these mutants supports our view that 1O2 signalling does not operate as an isolated linear pathway, but rather merges in a complex signalling network that integrates various cues and relays these to the nucleus (A. Baruah, K. Simkova, C.L. & K.A., unpublished data).

The finding that 1O2 can selectively activate nuclear genes, which are not—or only poorly—responsive to O2•− or H2O2, has also been reported for mammals (Klotz et al, 2003), the unicellular algae Chlamydomonas (Leisinger et al, 2001) and the phototrophic bacterium Rhodobacter sphaeroides (Anthony et al, 2005). In Chlamydomonas, a glutathione peroxidase homologous gene (GPXH) was shown to be strongly induced by 1O2-generating photosensitizers, but only weakly induced by O2•− or H2O2 (Fischer et al, 2005; Leisinger et al, 2001). A second gene, encoding heat-shock protein (HSP)70A, was shown to be induced by both 1O2 and H2O2. The study of its promoter has revealed that activation by 1O2 might require promoter elements that are different from those used by H2O2 (Shao et al, 2007). In Rhodobacter sphaeroides, the group IV sigma factor sigma(E) is required to activate a specific transcriptional response to 1O2, thereby protecting cells against this ROS (Anthony et al, 2005; Campbell et al, 2007; Glaeser & Klug, 2005). Protein synthesis patterns were also distinct from those after H2O2 treatment, but similar to those after high light exposure (Glaeser et al, 2007). Although the response to 1O2 is regulated primarily by sigma(E), mutant strain analysis revealed the existence of sigma(E)-independent pathways (Glaeser et al, 2007). In addition, the phrA gene, which encodes a photolyase, is regulated by both 1O2 and H2O2 in a sigma(E)-dependent manner, but not by treatment with paraquat, which generates O2•− that subsequently dismutates to H2O2 (Hendrischk et al, 2007).

Interactions of 1O2 signalling with other ROS

These results strongly suggest that, in various photosynthetic organisms, 1O2 and other ROS affect nuclear gene expression through distinct signalling pathways. These signalling pathways do not always operate independently and might interact with each other. In Arabidopsis, evidence supporting such interactions has been provided by non-invasively modulating the concentration of H2O2 in transgenic mutant plants that overexpress the plastid-specific thylakoid-bound ascorbate peroxidase (Murgia et al, 2004). The overexpression of this H2O2 scavenger enhanced the intensity of 1O2-mediated stress responses in the flu mutant, suggesting that H2O2 either directly or indirectly antagonizes 1O2-mediated signalling (Laloi et al, 2007). Similarly, pre-treatment of Chlamydomonas reinhardtii with low concentrations of the 1O2-generating photosensitizer rose bengal induced a rapid and transient acclimation to subsequently more intense 1O2-producing stress but, conversely, increased its sensitivity to paraquat (Ledford et al, 2007). Therefore, 1O2-mediated signalling might interact with H2O2-derived signals, suggesting that cells not only discriminate between different ROS but also that their responses to a given ROS can be modified by the concomitant perception of a second ROS. This discovery might explain previous observations suggesting that H2O2 can protect plants from high light stress. It seems that sensing the relative quantities of different ROS help to fine-tune the response to a specific stress.

Signalling versus cytotoxicity

An extensive second-site mutant screen of flu plants was undertaken to identify suppressor mutants that, during the 8 h dark period, accumulate similar excess amounts of free Pchlide as the original flu line but, on re-illumination, suppress 1O2-mediated growth inhibition and cell death of mature flu plants and/or the bleaching of flu seedlings. So far, we have focused our studies on one particular group of suppressor mutants, known as executer (ex). Allelism tests and mapping of the mutated genes revealed that they represent only a single nuclear locus, EX1, which encodes a plastid protein unrelated to known proteins (Wagner et al, 2004). Under non-permissive dark:light conditions the ex1/flu double mutant generates similar amounts of 1O2 as the parental flu line, but behaves similarly to wild-type seedlings as they remain viable and do not bleach (Fig 1, L/D), whereas mature plants continue to grow. It seems that because of the ex1 mutation, the flu plants have lost the ability to perceive the presence of 1O2 in their chloroplasts and to activate 1O2-mediated response programmes. These results show that bleaching of flu seedlings, growth inhibition and cell death of mature flu plants cannot be ascribed to photo-oxidative damage by 1O2, but instead reflect the 1O2-mediated activation of genetically controlled response programmes that require the activity of the EX1 gene. Enhanced levels of 1O2 within plastids of the flu mutant not only trigger marked phenotypic changes but also modulate nuclear gene expression. Indeed, as previously mentioned, within the first 30 min of the release of 1O2 rapid changes in nuclear gene expression occur. Inactivation of EX1 attenuated the upregulation of 1O2-responsive nuclear genes but did not fully eliminate these changes. A second, related nuclear-encoded protein, known as EX2, has been identified that is also implicated in the changes in signalling of 1O2-dependent nuclear gene expression (Lee et al, 2007). On inactivation of EX2 in the flu mutant, additional 1O2-responsive genes emerge and genes already upregulated in flu plants are either stimulated further or downregulated. The primary function of EX2 seems to be that of a modulator attenuating and controlling the activity of EX1. When both EXECUTER proteins are inactive in the ex1/ex2/flu triple mutant, most of the 1O2-responsive gene transcripts are close to wild-type level (Lee et al, 2007).

Both EX1 and EX2 are plastid-localized proteins that associate with thylakoid membranes (Lee et al, 2007). Therefore, the sensing of 1O2 seems to take place within the plastid, in close vicinity to its production sites within plastid membranes of the Arabidopsis flu mutant (Przybyla et al, 2008; op den Camp et al, 2003). In wild-type plants, 1O2 can be produced in the reaction centre of PSII and in the antenna system (Krieger-Liszkay, 2005). Whether these different production sites will affect differentially the sensing and signalling of 1O2 is not yet clear. Indeed, a diffusion distance significantly larger than what was previously thought has recently been reported for 1O2 in rat nerve cells (Skovsen et al, 2005). This indicates that sensing of 1O2 might be distant from the production site. In Chlamydomonas cells under high light-stress (3,500 μmol m−2 s−1), a small fraction of 1O2 produced from the PSII reaction centre has been reported to leave the chloroplast and activate the nuclear gene GPXH (Fischer et al, 2007). Whether EXECUTER proteins of Chlamydomonas are involved in mediating the 1O2 effect has not yet been addressed.

In our initial experiments, we compared flu and ex1/flu mutants treated with single or repetitive dark periods of 8 h at the rosette leaf or seedling stages. Under these conditions, as described above, introduction of the ex1 mutation suppresses growth inhibition and seedling lethality of the flu mutation. However, etiolated ex1/flu seedlings, which accumulate four- to fivefold the amounts of free Pchlide in flu and ex1/flu seedlings grown in continuous light and transferred to the dark for 8 h, bleach and die in a manner similar to the flu single mutant seedlings when transferred to light (Przybyla et al, 2008; Fig 1, D→L). This indicates that the amount of 1O2 produced in flu plants after an 8 h dark:light shift does not reach a cytotoxic level and induces genetically controlled stress programmes, whereas etiolated flu seedlings transferred to light generate cytotoxic amounts of 1O2 (Fig 2). In agreement with this idea, measurements of oxidation products of polyunsaturated fatty acids show that flu plants re-illuminated after being kept in the dark for 8 h accumulate enzymatically produced oxylipins, such as 13-hydroxy octadecatrienoic acid (13-HOT), 13-hydroxy octadecadienoic acid (13-HOD), oxophytodienoic acid and jasmonic acid, but not the products of non-enzymatic lipid peroxidation, such as 10-HOD, 12-HOD, 10-HOT and 15-HOT (Przybyla et al, 2008). By contrast, etiolated flu seedlings moved into the light accumulate large amounts of the non-enzymatically formed oxylipins 10-HOD, 12-HOD, 10-HOT and 15-HOT, which indicates that, under these more harsh conditions, toxic effects of 1O2 prevail (Fig 2; Przybyla et al, 2008).

Figure 2.

Figure 2

Proposed model of plant responses to various levels of 1O2. Constantly changing environmental factors can lead to the generation of variable levels of 1O2 that can be mimicked in the flu mutant by modulating the length of exposure to darkness. Increasing dark incubation of flu plants leads to a proportional increase of free protochlorophyllide in the dark and of 1O2 generation on illumination. When the production of 1O2 is low, an acclimatory response might be activated that allows a plant to survive at subsequent higher levels of 1O2. Greater amounts of 1O2, for example in flu plants treated with 8 h of darkness and re-illuminated, trigger genetically controlled cell death through EXECUTER-dependent pathways. Most enzymatically oxidized lipid derivatives are produced under these conditions. When the level of 1O2 increases further, plants die owing to chemical damage caused by 1O2 and a massive production of non-enzymatically oxidized lipid is observed. EX, EXECUTER protein; PCD, programmed cell death.

Perspectives

An interesting remaining question is whether 1O2 levels lower than those produced in flu mutants after 8 h of darkness induce qualitatively different changes in gene expression that might lead to acclimation (Fig 2). This question could be addressed by first pre-exposing flu mutants to short dark periods that give rise to low concentrations of 1O2 during re-illumination and then applying a much harsher stress treatment. Recent studies of Chlamydomonas indicate that sub-lethal stress leads to a transient or moderate elevation of 1O2, and results in the activation of a subset of 1O2-responsive genes and protection against a subsequent more severe stress (Ledford et al, 2007). The identification of differentially affected genes under conditions that induce genetically controlled cell death or acclimation will be important for modifying genetic constraints that determine the adaptability of plants to environmental changes.

graphic file with name embor200857-i1.jpg

Chanhong Kim

graphic file with name embor200857-i2.jpg

Rasa Meskauskiene

graphic file with name embor200857-i3.jpg

Klaus Apel

graphic file with name embor200857-i4.jpg

Christophe Laloi

Acknowledgments

Work in our group has been supported by grants from the Swiss National Science Foundation (NSF), the Functional Genomic Center Zurich (FGCZ) and ETH-Zurich.

References

  1. Anthony JR, Warczak KL, Donohue TJ (2005) A transcriptional response to singlet oxygen, a toxic byproduct of photosynthesis. Proc Natl Acad Sci USA 102: 6502–6507 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Apel K, Hirt H (2004) Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Annu Rev Plant Biol 55: 373–399 [DOI] [PubMed] [Google Scholar]
  3. Asada K (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 [DOI] [PubMed] [Google Scholar]
  4. Asada K (2006) Production and scavenging of reactive oxygen species in chloroplasts and their functions. Plant Physiol 141: 391–396 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Campbell EA, Greenwell R, Anthony JR, Wang S, Lim L, Das K, Sofia HJ, Donohue TJ, Darst SA (2007) A conserved structural module regulates transcriptional responses to diverse stress signals in bacteria. Mol Cell 27: 793–805 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Danon A, Miersch O, Felix G, Camp RG, Apel K (2005) Concurrent activation of cell death-regulating signalling pathways by singlet oxygen in Arabidopsis thaliana. Plant J 41: 68–80 [DOI] [PubMed] [Google Scholar]
  7. Fischer BB, Krieger-Liszkay A, Eggen RIL (2005) Oxidative stress induced by the photosensitizers neutral red (type I) or rose bengal (type II) in the light causes different molecular responses in Chlamydomonas reinhardtii. Plant Sci 168: 747–759 [Google Scholar]
  8. Fischer BB, Krieger-Liszkay A, Hideg E, Snyrychova I, Wiesendanger M, Eggen RI (2007) Role of singlet oxygen in chloroplast to nucleus retrograde signaling in Chlamydomonas reinhardtii. FEBS Lett 581: 5555–5560 [DOI] [PubMed] [Google Scholar]
  9. Foote CS (1968) Mechanisms of photosensitized oxidation. There are several different types of photosensitized oxidation which may be important in biological systems. Science 162: 963–970 [DOI] [PubMed] [Google Scholar]
  10. Foreman J et al. (2003) Reactive oxygen species produced by NADPH oxidase regulate plant cell growth. Nature 422: 442–446 [DOI] [PubMed] [Google Scholar]
  11. Foyer CH, Noctor G (2003) Redox sensing and signalling associated with reactive oxygen in chloroplasts, peroxisomes and mitochondria. Plant Physiol 119: 355–364 [Google Scholar]
  12. Gadjev I, Vanderauwera S, Gechev TS, Laloi C, Minkov IN, Shulaev V, Apel K, Inze D, Mittler R, Van Breusegem F (2006) Transcriptomic footprints disclose specificity of reactive oxygen species signalling in Arabidopsis. Plant Physiol 141: 436–445 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Gapper C, Dolan L (2006) Control of plant development by reactive oxygen species. Plant Physiol 141: 341–345 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Gechev TS, Van Breusegem F, Stone JM, Denev I, Laloi C (2006) Reactive oxygen species as signals that modulate plant stress responses and programmed cell death. Bioessays 28: 1091–1101 [DOI] [PubMed] [Google Scholar]
  15. Glaeser J, Klug G (2005) Photo-oxidative stress in Rhodobacter sphaeroides: protective role of carotenoids and expression of selected genes. Microbiology 151: 1927–1938 [DOI] [PubMed] [Google Scholar]
  16. Glaeser J, Zobawa M, Lottspeich F, Klug G (2007) Protein synthesis patterns reveal a complex regulatory response to singlet oxygen in Rhodobacter. J Proteome Res 6: 2460–2471 [DOI] [PubMed] [Google Scholar]
  17. Gollnick K (1968) Type 2 photosensitized oxygenation reactions. Adv Chem Ser 77: 78–101 [Google Scholar]
  18. Halliwell B (2006) Reactive species and antioxidants. Redox biology is a fundamental theme of aerobic life. Plant Physiol 141: 312–322 [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Havaux M, Eymery F, Porfirova S, Rey P, Dormann P (2005) Vitamin E protects against photoinhibition and photooxidative stress in Arabidopsis thaliana. Plant Cell 17: 3451–3469 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Hendrischk AK, Braatsch S, Glaeser J, Klug G (2007) The phrA gene of Rhodobacter sphaeroides encodes a photolyase and is regulated by singlet oxygen and peroxide in a sigma(E)-dependent manner. Microbiology 153: 1842–1851 [DOI] [PubMed] [Google Scholar]
  21. Kalai T, Hideg E, Vass I, Hideg K (1998) Double (fluorescent and spin) sensors for detection of reactive oxygen species in the thylakoid membrane. Free Radic Biol Med 24: 649–652 [DOI] [PubMed] [Google Scholar]
  22. Klotz LO, Kroncke KD, Sies H (2003) Singlet oxygen-induced signalling effects in mammalian cells. Photochem Photobiol Sci 2: 88–94 [DOI] [PubMed] [Google Scholar]
  23. Kozaki A, Takeba G (1996) Photorespiration protects C3 plants from photooxidation. Nature 384: 557–560 [Google Scholar]
  24. Krieger-Liszkay A (2005) Singlet oxygen production in photosynthesis. J Exp Bot 56: 337–346 [DOI] [PubMed] [Google Scholar]
  25. Krieger-Liszkay A, Trebst A (2006) Tocopherol is the scavenger of singlet oxygen produced by the triplet states of chlorophyll in the PSII reaction centre. J Exp Bot 57: 1677–1684 [DOI] [PubMed] [Google Scholar]
  26. Laloi C, Przybyla D, Apel K (2006) A genetic approach towards elucidating the biological activity of different reactive oxygen species in Arabidopsis thaliana. J Exp Bot 57: 1719–1724 [DOI] [PubMed] [Google Scholar]
  27. Laloi C, Stachowiak M, Pers-Kamczyc E, Warzych E, Murgia I, Apel K (2007) Cross-talk between singlet oxygen- and hydrogen peroxide-dependent signalling of stress responses in Arabidopsis thaliana. Proc Natl Acad Sci USA 104: 672–677 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Ledford HK, Chin BL, Niyogi KK (2007) Acclimation to singlet oxygen stress in Chlamydomonas reinhardtii. Eukaryot Cell 6: 919–930 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Lee KP, Kim C, Landgraf F, Apel K (2007) EXECUTER1- and EXECUTER2-dependent transfer of stress-related signals from the plastid to the nucleus of Arabidopsis thaliana. Proc Natl Acad Sci USA 104: 10270–10275 [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Leisinger U, Rufenacht K, Fischer B, Pesaro M, Spengler A, Zehnder AJ, Eggen RI (2001) The glutathione peroxidase homologous gene from Chlamydomonas reinhardtii is transcriptionally up-regulated by singlet oxygen. Plant Mol Biol 46: 395–408 [DOI] [PubMed] [Google Scholar]
  31. Meskauskiene R, Apel K (2002) Interaction of FLU, a negative regulator of tetrapyrrole biosynthesis, with the glutamyl-tRNA reductase requires the tetratricopeptide repeat domain of FLU. FEBS Lett 532: 27–30 [DOI] [PubMed] [Google Scholar]
  32. Meskauskiene R, Nater M, Goslings D, Kessler F, op den Camp R, Apel K (2001) FLU: a negative regulator of chlorophyll biosynthesis in Arabidopsis thaliana. Proc Natl Acad Sci USA 98: 12826–12831 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Muller P, Li XP, Niyogi KK (2001) Non-photochemical quenching. A response to excess light energy. Plant Physiol 125: 1558–1566 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Murgia I, Tarantino D, Vannini C, Bracale M, Carravieri S, Soave C (2004) Arabidopsis thaliana plants overexpressing thylakoidal ascorbate peroxidase show increased resistance to paraquat-induced photooxidative stress and to nitric oxide-induced cell death. Plant J 38: 940–953 [DOI] [PubMed] [Google Scholar]
  35. Niyogi KK (1999) Photoprotection revisited: genetic and molecular approaches. Annu Rev Plant Physiol Plant Mol Biol 50: 333–359 [DOI] [PubMed] [Google Scholar]
  36. Ochsenbein C, Przybyla D, Danon A, Landgraf F, Gobel C, Imboden A, Feussner I, Apel K (2006) The role of EDS1 (enhanced disease susceptibility) during singlet oxygen-mediated stress responses of Arabidopsis. Plant J 47: 445–456 [DOI] [PubMed] [Google Scholar]
  37. op den Camp RG et al. (2003) Rapid induction of distinct stress responses after the release of singlet oxygen in Arabidopsis. Plant Cell 15: 2320–2332 [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Przybyla D, Gobel C, Imboden A, Feussner I, Hamberg M, Apel K (2008) Enzymatic, but not non-enzymatic 1O2-mediated peroxidation of polyunsaturated fatty acids forms part of the EXECUTER1-dependent stress response program in the flu mutant of Arabidopsis thaliana. Plant J 54: 236–248 [DOI] [PubMed] [Google Scholar]
  39. Rizhsky L, Liang H, Mittler R (2003) The water–water cycle is essential for chloroplast protection in the absence of stress. J Biol Chem 278: 38921–38925 [DOI] [PubMed] [Google Scholar]
  40. Sagi M, Fluhr R (2006) Production of reactive oxygen species by plant NADPH oxidases. Plant Physiol 141: 336–340 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Shao N, Krieger-Liszkay A, Schroda M, Beck CF (2007) A reporter system for the individual detection of hydrogen peroxide and singlet oxygen: its use for the assay of reactive oxygen species produced in vivo. Plant J 50: 475–487 [DOI] [PubMed] [Google Scholar]
  42. Skovsen E, Snyder JW, Lambert JD, Ogilby PR (2005) Lifetime and diffusion of singlet oxygen in a cell. J Phys Chem B 109: 8570–8573 [DOI] [PubMed] [Google Scholar]
  43. Takemoto D, Tanaka A, Scott B (2006) A p67Phox-like regulator is recruited to control hyphal branching in a fungal–grass mutualistic symbiosis. Plant Cell 18: 2807–2821 [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Tanaka A, Christensen MJ, Takemoto D, Park P, Scott B (2006) Reactive oxygen species play a role in regulating a fungus–perennial ryegrass mutualistic interaction. Plant Cell 18: 1052–1066 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Telfer A (2005) Too much light? How β-carotene protects the photosystem II reaction centre. Photochem Photobiol Sci 4: 950–956 [DOI] [PubMed] [Google Scholar]
  46. Telfer A, Bishop SM, Phillips D, Barber J (1994) Isolated photosynthetic reaction-center of photosystem-II as a sensitizer for the formation of singlet oxygen—detection and quantum yield determination using a chemical trapping technique. J Biol Chem 269: 13244–13253 [PubMed] [Google Scholar]
  47. Trebst A (2003) Function of β-carotene and tocopherol in photosystem II. Z Naturforsch 58c: 609–620 [DOI] [PubMed] [Google Scholar]
  48. Wagner D et al. (2004) The genetic basis of singlet oxygen-induced stress responses of Arabidopsis thaliana. Science 306: 1183–1185 [DOI] [PubMed] [Google Scholar]

Articles from EMBO Reports are provided here courtesy of Nature Publishing Group

RESOURCES