Skip to main content
Plant Signaling & Behavior logoLink to Plant Signaling & Behavior
. 2012 Dec 1;7(12):1621–1633. doi: 10.4161/psb.22455

Reactive oxygen species generation and signaling in plants

Baishnab Charan Tripathy 1,*, Ralf Oelmüller 2
PMCID: PMC3578903  PMID: 23072988

Abstract

The introduction of molecular oxygen into the atmosphere was accompanied by the generation of reactive oxygen species (ROS) as side products of many biochemical reactions. ROS are permanently generated in plastids, peroxisomes, mitochiondria, the cytosol and the apoplast. Imbalance between ROS generation and safe detoxification generates oxidative stress and the accumulating ROS are harmful for the plants. On the other hand, specific ROS function as signaling molecules and activate signal transduction processes in response to various stresses. Here, we summarize the generation of ROS in the different cellular compartments and the signaling processes which are induced by ROS.

Keyword: reactive oxygen species, signal transduction, plastids

Introduction

With introduction of molecular oxygen (O2) into our atmosphere by O2-evolving photosynthetic organisms early in the evolution of aerobic life, reactive oxygen species (ROS) have become an integral part of life. The activation or reduction of oxygen gives rise to reactive ROS that includes the singlet oxygen (1O2), superoxide (O2), hydrogen peroxide (H2O2) and hydroxyl radical (HO.). Plants and other living organisms in the oxidizing environment constantly produce ROS in chloroplasts, mitochondria, peroxisomes and other sites of the cell because of their metabolic processes such as photosynthesis and respiration.

The generation of ROS in plants is triggered by different kinds of environmental stresses, such as high light, high or low temperature, salinity, drought, nutrient deficiency and pathogen attack. Plants and other living organisms have evolved a host of anti-oxidants and anti-oxidative enzymes and other small molecules to harmlessly dissipate ROS. Imbalance between ROS production and their detoxification by enzymatic and non-enzymatic reactions causes oxidative stress. As a result of higher net ROS formation, there is photooxidative damage to DNA, proteins and lipids and ultimately cell death. ROS also act as signaling molecules involved in growth and developmental processes, pathogen defense responses such as hypersensitive reaction and systemic acquired resistance, stress hormone production, acclimation and programmed cell death.1

Imbalance between ROS generation and safe detoxification represents metabolic states that frequently are referred to as oxidative stress. In biological systems, oxidative stress results from the presence of elevated levels of oxidizing agents that are able to abstract electrons from essential organic molecules and disturb cellular function. When ROS production exceeds antioxidizing capacity, this can lead to cell damage and ultimately cell death by ROS toxicity and/or specific ROS-activated cell-death-inducing signaling events.

Although dioxygen in its ground state is relatively unreactive, partial activation or reduction gives rise to ROS including 1O2, O2, H2O2 and HO.. The evolution of aerobic metabolic processes such as respiration and photosynthesis unavoidably led to the production of such ROS in mitochondria, chloroplast and peroxisomes along with the other cellular and extracellular compartments. The common feature among the different ROS types is their capacity to cause oxidative damage to proteins, DNA and lipids.

Chemistry of Oxygen in Biological Systems

Although molecular oxygen contains an even number of electron, it has two unpaired electron in its molecular orbital. These two electrons have the same spin quantum number, and if O2 attempts to oxidize another atom or molecule by accepting a pair of electron from it, both new electrons must be of parallel spin so as to fit into the vacant orbital. Usually, a pair of electrons in an atomic or molecular orbital would have anti-parallel spins. This imposes an important restriction on oxidation by O2. Molecular O2 is usually constrained to one electron transfer reaction at a time.

A way to overcome spin restriction, O2 molecule interacts with another paramagnetic center. Transition metals like Fe or Cu frequently have unpaired electrons.

FeII + O2 → FeIII + O2– .

In aqueous solution O2– . decays to H2O2

O2– . + O2– . + 2H+ → H2O2 + O2

The protonation of O2– . produces the hydroperoxyl radical, a far more reactive species.

O2– . + H+ → HOO .

The decomposition of H2O2 can be a source of hydroxyl radical, HO .

FeIII + O2– . → FeII + O2

FeII + H2O2 → FeIII + HO + HO .

Production of Reactive Oxygen Intermediates in Plants

Organelles with a highly oxidizing metabolic activity or with an intense rate of electron flow, such as chloroplast, mitochondrium and peroxisome, are a major source of ROS production in plants. Along with these organelles, peroxidases and amine oxidases present in cell walls and NADPH oxidase located in the plasma membrane produce ROS, often in response to stress signals. Oxygen is continuously produced inside the chloroplast due to photosynthetic electron transport and simultaneously removed by reduction and assimilation.

Photoreduction of oxygen to the superoxide radical occurred due to reduced electron transport components associated with photosystem (PS)I and a reaction linked to the photorespiratory cycle in the peroxisome. When the availability of CO2 is restricted inside the leaves due to various environmental stresses, ribulose-1,5-bisphosphate carboxylase/oxygenase catalyzes a competitive reaction in which oxygen is favored over CO2 as a substrate. This oxygenation reaction leads to the formation of glycolate, which is transported to peroxisomes and produce H2O2 after subsequent oxidation by glycolate oxidase.

Plasma membrane NADPH-dependent oxidases contain a multimeric flavocytochrome that forms an electron transport chain capable of reducing O2 to O2 . Inhibitors of this oxidase impair ROS production.2 In addition to NADPH oxidase, pH dependent cell wall peroxidase, oxalate oxidase and amine oxidases have been proposed to produce ROS in the apoplast.3,4

Though mitochondria are the major source of ROS production in mammalian cells, in green tissues its contribution for ROS is not effective.5 The mitochondrial electron transport chain can produce ROS, which can transfer a single electron to oxygen and produce O2-. The mitochondrial alternative oxidase (AOX) catalyses the O2-dependent oxidation of ROS. Lack of AOX induction caused increased ROS production6 and tobacco plants overexpressing AOX have small hypersensitive response lesions than wild type in response to virus infection.7 H2O2 treatment of Arabidopsis cells and H2O2 accumulation in catalase–deficient tobacco lead to induction of antioxidant defenses and increased AOX levels in mitochondria.8

Formation of Singlet Oxygen in Plants

Oxygen in its ground state is not very reactive and does not have any deleterious effect. The ground state molecular oxygen is a triplet state (3O2) and in fact a biradical, as it has two unpaired electrons. Its two unpaired electrons have parallel spins () that do not allow them to react with most molecules. However, if the triplet oxygen absorbs sufficient energy, the spin restriction is removed and the spin of one of its unpaired electrons is reversed. As a result there is generation of singlet oxygen (1O2), whose outermost pair of electrons has antiparallel spins (). Singlet oxygen molecules are also formed when superoxide radicals interact with hydroxyl radicals. The excitation energy required to produce 1O2 from the triplet oxygen is 94 kJ mol-1.9 It has a short half-life of about 200 ns in cells with a possible diffusion distance of about 270 nm and could even diffuse out of the chloroplast into the cytosol.9

In plants, 1O2 is mainly produced by the chlorophyll (Chl) and its tetrapyrrole metabolic intermediates in the presence of light. Chl, the most abundant pigment in land plants, is the main light absorbing pigment and is present both in the light harvesting complex (LHC) and the photosynthetic reaction centers. The excited state of these is long lived and allow conversion of the excitation energy to an electrochemical potential via charge separation. Inefficient transfer of energy results in the generation of triplet state Chl that reacts with triplet oxygen to produce the highly reactive 1O2. In the light harvesting complex, the 1O2 is quenched by the carotenoids. (for further information on the photochemistry of carotenoids, see ref. 10.)

1O2 is produced near the reaction centers of the PSs. With increase in light intensity i.e., from the early morning to noon, light absorption by leaves increases almost linearly. However, the rate of photosynthesis reaches its maximum value much before the linear increase in light absorption ceases. Therefore plants end up absorbing more light than they could utilize in photosynthesis. This results in the over excitation of the photosynthetic apparatus. In the presence of excess light energy, the QA and QB (the first and second plastoquinone electron acceptors of PSII) in the electron transport chain is over-reduced11 and because of that, charge separation cannot be completed between P680 and pheophytin. As a result the triplet state of the reaction center Chl P680 (3P680) is favored12,13 leading to the formation of 1O2 .14 Normally when excess light is absorbed, an alternative dissipating pathway is activated that safely returns 1Chl* to its ground state before it is converted to 3Chl*. The excitation energy of excess 1Chl* is dissipated by zeaxanthin or other carotenoids as heat in Chl and/or carotenoid binding protein complexes.15-19 The carotenoids, which quench the excited state of Chl, must be in close proximity with triplet Chl i.e., within a maximum distance of 3.6 Å. In this spin exchange reaction, the triplet state of carotenoids is formed that can dissipate the excess energy as heat. In the reaction center, the distance between Chl and carotenoid is too large to allow triplet quenching. 1O2, produced in the reaction center, directly reacts with carotenoids. The release of 1O2 is also detected in isolated PS II particles20 and in thylakoids.21-241O2 is also generated from the cytochrome-b6/f-complex.25

Generation of Singlet Oxygen from Chlorophyll Biosynthesis Intermediates

Upon illumination, Chl biosynthesis intermediates i.e., protochlorophyllide (Pchlide) or protoporphyrin IX (Proto IX) produce 1O2 in plants and cause oxidative damage.21,22,26,27 Formation of reactive oxygen species, from Chl biosynthesis intermediates was also proposed by others.28-35 The site of generation of 1O2 is mostly in the thylakoids. This is because Chl biosynthesis intermediates are partially hydrophobic, and consequently are loosely attached to the thylakoid membranes.36-38 Although they are associated with the thylakoid membranes, these tetrapyrroles do not form pigment protein complexes and hence are not connected to the reaction center. Although some of the carotenoids are present in the lipid bilayer, a lot more are located in the pigment-protein complexes and they are spatially too far from Chl biosynthesis intermediates to quench their triplet states.39,40 Synthesis of Chl biosynthetic intermediates is highly regulated so that intermediates are not overproduced in plants. However, Chl biosynthesis intermediates that are normally present in plants are capable of producing 1O2 that cause oxidative damage in high light and several other stress conditions.21

Porphyrin-Generating Compounds

There are two important types of porphyrin-generating compounds. One consists of 5-aminolevulinic acid (ALA), the substrate of tetrapyrroles and the other is a group of diphenyl ethers that inhibit protoporphyrinogen oxidase activity, thereby deregulating the tetrapyrrole metabolism. Cercosporin, rose bengal and several other compounds could generate 1O2 and O2-.41,42

5-Aminolevulinic acid

5-Aminolevulinic acid (ALA) is the sole precursor of all tetrapyrroles i.e., Chl, hemes, sirohemes, and phytochromobilins. Tetrapyrrole intermediates are photosensitizers and generate radicals and ROS, especially 1O2 in the presence of light. So plants regulate their own tetrapyrrole biosynthesis and degradation pathway to avoid the consequence of the excess generation of ROS. The major regulatory point is at the production of the initial precursor ALA. Therefore, ALA synthesis is the rate-limiting step of the tetrapyrrole biosynthetic pathway. ALA is formed from glutamyl-tRNA by the enzyme glutamyl-tRNA reductase (GluTR). In Arabidopsis this enzyme has three isoforms (HEMA1, HEMA2 and HEMA3) and their expression levels are different in different plant tissues. Pchlide, which accumulates in the dark, repress the ALA synthesis by feed-back inhibition.43,44 However, when ALA is applied externally, green plants bypass the regulatory feedback inhibition of Pchlide pool and induce excess accumulation of Mg-tetrapyrroles in the dark.31,32,45 When only ALA is applied externally, Pchlide is the major porphyrin that accumulates. But after ALA + Modulator (= a compound structurally related to tetrapyrrole molecule) treatments, several other types of porphyrins accumulate, depending upon the target site of the modulator.33 ALA is added along with the modulator for providing the carbon skeleton to accumulate porphyrins. Eight molecules of ALA are required to form one molecule of tetrapyrrole. The mode of action of some of the modulators has been attributed to their metal chelating properties. Enzymes in the porphyrin synthesizing pathway essentially require certain metal ions for their activity. Another way by which some of these modulating chemicals may be acting is by stimulating enzyme activity, i.e. by behaving as cofactor analogs.

Diphenyl ethers

The diphenyl ethers (DPEs) are inhibitors of the enzyme protoporphyrinogen oxidase (protox),46 an enzyme that converts protoporphyrinogen IX to Proto IX. The latter is an intermediate in Chl and heme biosynthetic pathway. DPEs cause plants to overaccumulate a large quantity of Proto IX.47 Pchlide, Mg-Proto IX and Mg-Proto IX monomethylester (MPE) are also found to be elevated, but to a significantly less extent as compared to Proto IX, in tissues with inhibited protox activity. Though DPE entry into cells is light-independent, light and pigments are mandatory for their herbicidal action. Initial symptom of DPE damage is seen as water soaked spots on leaf tissue, followed by loss of leaf turgidity, bleaching and necrosis. More lipophilic DPEs like oxyfluorfen and acifluorfen-methyl have greater potency as herbicides than the more polar acifluorfen. This correlation may also be partly due to the greater ease of penetration through the cuticle by the more lipophilic molecules. DPEs can inhibit carotenoid synthesis, ATP formation, photosynthetic electron transport, and induce membrane peroxidation by causing massive accumulation of Proto IX.48,49 Superoxide radical is not of primary importance in the development of DPE toxicity.50-52 Oxyfluorfen added to isolated thylakoid membranes, in vitro, generates 1O2 during illumination.42 Treatment with the protox inhibitor acifluorfen-sodium (AF-Na) in the light induced the overaccumulation of protoporphyrinogen IX that migrates out of the chloroplast to the cytoplasm where it is oxidized to Proto IX by plasma membrane bound AF-Na-insensitive protoporphyrinogen oxidase.47,53 A part of Proto IX so generated in the plasma membrane migrates back to the chloroplast and partitions between the cytoplasm and the chloroplast.53,54 The 1O2 generated by the photosensitization reaction of Proto IX creates necrotic spots and cell death by destroying the plasma membrane.27 The DPE herbicide lactofen induces cell death and expression of PR1, PR5 and PR10 protein in soybean plants. The anthocyanin biosynthesis pathway genes i.e., chalcone synthase and chalcone reductase are also induced in lactofen treated samples.55

Type II and Type I Photosensitization Reactions of Tetrapyrroles

In type II photosensitization reaction, a sensitizer can transfer its excitation energy to a ground state oxygen molecule, resulting in 1O2. ALA or DPE compounds induce over-accumulation of non-phototransformable Pchlide or Proto IX, respectively. In the presence of light, these tetrapyrroles generate 1O2 through type II photosensitization reaction, which ultimately destroys the plant.21,22,27 1O2 is quite selective and fails to react with molecules that are not enough electron-rich, and simply returns to the ground state.

The type I photosensitization involves hydrogen atom or electron transfer from the sensitizer to the substrate.56 The resulting free radicals can subsequently react with O2 to produce oxidized products or other reactive species. The products are often peroxides, which can in turn breakdown to induce free radical chain auto-oxidation, leading to further oxidation in a non-photochemical step.

LOO۰ + LH ---> LOOH + L,

where L and LOO۰ are lipid free radicals; LH is an unsaturated lipid and LOOH is a lipid hydroperoxide.

Sensitizers (Sens) can produce superoxide radical (O2–.) by undergoing electron transfer processes with the substrate or O2.

3Sens + Subs ----> Sens- + Subsox

Sens- + O2 ----> Sens + O2– .

or

3Sens + O2 ----> Sens+ + O2– .

These reactions produce O2– ., which can subsequently give rise to the highly reactive hydroxyl radical (OH). The hydroxyl radicals thus produced can react with organic molecules in a variety of ways or can initiate radical chain auto-oxidation.

Intracellular Destruction of Singlet Oxygen.

The most efficient mechanism of detoxification of 1O2 in plants involves carotenoids. The carotenoids reach the triplet excited state by absorbing the excess energy of 1O2 which returns it to its triplet ground state (3O2). They finally dissipate the excess acquired energy as heat.57 The physical quencher has to be lipid soluble and needs to be very close to the photosensitizer. Carotenoids, because of their conjugated double bonds, are the most abundant quenchers of 1O2 in the pigment bed of the photosynthetic apparatus. Photosynthetic antenna systems have several xanthophylls i.e., lutein, violaxanthin, neoxanthin, and zeaxanthin. Out of these, lutein is the most abundant as it is needed for efficient quenching of 3Chl*. Zeaxanthin is synthesized from violaxanthin under high-light stress by the violaxanthin deepoxidase enzyme and is involved in energy dependent quenching of Chl a fluorescence (see e.g., 40. Tocopherol is lipid soluble and is a minor but significant component of 1O2 quenchers in the thylakoid membranes. The suppression of both zeaxanthin and tocopherol in the npq1/ vte1 double mutant results in 1O2-mediated lipid peroxidation in high light.58,59 In the PSII reaction center, especially under high light regime, 1O2 is quenched by β-carotene and α-tocopherol.60

Scavengers of 1O2 are usually water soluble and are themselves oxidized or destroyed while detoxifying the ROS. The oxidized scavenger is re-reduced by a set of biochemical reduction reactions at a great cost to the cell. The cell has only limited capability to resynthesize the destroyed scavengers. Therefore, the cells become extremely prone to 1O2 attack. Ascorbate is an example of 1O2 scavenger that is oxidized after detoxification. It is predominantly present in the plastids. 1O2 reacts with ascorbate to produce dehydroascorbate.61 The latter is converted back to ascorbate by the dehydroascorbate reductase, and the glutathione reductase enzymes involved in the Halliwell-Asada pathway. Vitamin B6 (pyridoxine, pyridoxal, pyridoxamine) can efficiently scavenge 1O2.62,63 The fungus Cercospora secretes cercosporin, a 1O2-generating photosensitizer into the extracellular matrix during plant infection.63 The cercosporin secreted to the host cell by the fungus absorbs solar energy and transfers its energy to oxygen to generate 1O2 that kills the host cell. However, the fungus itself is protected against 1O2-mediated damage by the 1O2- scavenger vitamin B6. In the same vein, the pyridoxine synthase is involved in tolerance to oxidative stress.64 Similarly, exogenous vitamin-B6 protects protoplasts of the flu (fluorescence) mutants of Arabidopsis thaliana that generated 1O2.65 Flavonoids that are present in plants in high concentrations in the cytoplasm and isoprene that is mostly synthesized in the chloroplasts could also function as 1O2 quenchers.66-69 The water soluble chlorophyll binding protein (WSCP) binds to free Chl molecules as well as to its biosynthetic intermediates and does not allow them to get photoactivated to produce1O2. It acts as a physical barrier between free Chl molecules and molecular oxygen.70

Singlet Oxygen-induced Oxidative Damage in Mutants

Chlorophyll anabolic mutants

Tigrina mutant of barley accumulates 2-10 times more Pchlide in darkness than the wild type.71 Homozygous tigrina-d mutants are fully green and viable if grown in continuous weak light, but show a green-white banded phenotype, when grown under light/dark cycles. They have normal level of protochlorophyllide oxidoreductase (POR) and when illuminated, the excess Pchlide causes photodynamic damage resulting in the formation of necrotic patches.72 Runge et al.73 isolated xantha mutants of Arabidopsis and classified them in two groups, mutants that are blocked in various steps of the Chl biosynthetic pathway prior to POR, and mutants that accumulate Pchlide in the dark. The etiolated PORA and PORB mutant seedlings accumulate significant amounts of non-phototransformable Pchlide in darkness and upon light exposure they show bleaching effect and germination defect.74 However, overexpression of PORA and PORB results in the efficient transformation of non-phototransformable Pchlide to chlorophyllide (Chlide). Transgenic seedlings grown under far red light when transferred to white light are more resistant to photobleaching because of high POR proteins.75

The isolation and studies on Arabidopsis flu mutant by Klaus Apel’s group confirm the role of Pchlide in 1O2 generation; it leads to oxidative damage. In flu mutant there is a massive accumulation of Pchilde if those plants are grown under constant dark/light cycle and there is growth arrest and cell death because of generation of 1O2. However if the plants are grown under continuous light, there is no accumulation of Pchlide, no generation of 1O2 and plants behave like wild type plants. The FLU protein interacts with GulTR (HEMA1) and regulates the accumulation of Pchlide in darkness.76 In the flu mutant there is no regulatory point at the GluTR level and there is a massive accumulation of Pchlide. Lee et al.77 have revealed that the TIGRINA d gene of barley is an ortholog of the FLU gene of Arabidopsis thaliana. Pchlide-mediated 1O2 formation leads to the induction of the early stress-responsive gene.26 There is no change in amounts of other photosensitizers i.e, Proto IX, Mg -proto IX and MPE in the flu mutant. Two major stress reactions were observed when dark-grown flu plants were returned to light: a cell death response and a rapid inhibition of growth. Oxygenation derivatives of linolenic acid, by far the most prominent polyunsaturated fatty acid of chloroplast membrane lipids, start to accumulate rapidly in the flu mutant after the dark/light shift. The oxidation of linolenic acid is not caused by direct interaction with 1O2 but instead occurs enzymatically. Thus, the development of stress symptoms in the flu mutant seems not to be attributable to cell damage caused by 1O2 but rather appears to result from the more indirect role of this ROS. Vitamin B6 that quenches 1O2 in fungi was able to protect flu protoplasts from cell death65; further, protoplasts of flu mutant depleted of both ethylene and salicylic acid had reduced cell death. However, when the gene Executer1 (exe1) was mutated in the flu background, the exe1/flu double mutant accumulated free Pchlide in the dark like the flu mutant, but unlike the wild type plants. After transfer to light, exe1/flu generated 1O2 in amounts similar to those of flu but grew like wild type when kept under non-permissive light-dark cycles.78 In flu plants, the growth rate was reduced immediately after the beginning of re-illumination. The exe1/flu plants, however, grew like wild-type plants. Growth inhibition of flu plants was particularly striking when plants were transferred to repeated light-dark cycles, whereas the exe1/flu mutant continued to grow like wild-type plants.78 Both assays demonstrate that the rapid bleaching of flu seedlings and the inhibition of growth after the release of 1O2 are not by the toxicity of this ROS; further, these do not reflect photooxidative damage and injury, but instead result from the activation of genetically controlled responses that require the activity of the Executer1 gene. The isolation of the Executer 2 protein also shows a similar kind of response in flu back ground.79 Inactivation of executer proteins blocks the 1O2-mediated signaling from the chloroplast to the nucleus that affects the normal plastid development in germinating seeds.80 Coll et al.81 have isolated another 1O2-linked death activator (soldat8) that encodes the SIGMA6 factor of the plastid RNA polymerase, specifically abrogate 1O2-mediated stress responses in young flu seedlings without grossly affecting 1O2-mediated stress responses of mature flu plants. The other protein named PRL1 (Pleiotropic response locus 1) also affect the expression of 1O2- responsive genes in Arabidopsis.82

Apart from Pchlide, early intermediates i.e., coproporphyrin also acts as a photosensitizer.83-85 The antisense coproporphyrinogen oxidase (that converts coproporphyrinogen III to protoporphyrinogen IX) in tobacco plants, have an excessive amount of coproporphyrin. This oxidized porphyrin gives rise to photodynamic reactions, which affect cellular processes resulting in retarded growth and necrotic leaves.84,85 The Arabidopsis coproporphyrinogen oxidase mutants (lin2; lesion initiation 2) had pale leaves and developed lesions on the young leaves.83 3,3-Diamino benzidine and trypan blue staining of the mutant leaves shows H2O2 accumulation and cell death. Seedlings homozygous for a null mutation in the cpx1 gene of maize completely lack chlorophyll and develop necrotic lesions in the light.86 The accumulation of uroporphyrin I in the uroporphyrinogen III cosynthase antisense barley plants results in necrotic leaves and ultimately cell death because of accumulation of ROS.87 Like uroporphyrin I, uroporphyrin III, an oxidized derivative of uroporphyrinogen III, an intermediate of the chlorophyll biosynthesis pathway, also acts as a photosensitizer. Accumulation of uroporphyrin III leads to light-dependent necrosis in tobacco30,88 and in maize.89 Antisense tobacco plants of uroporphyrinogen decarboxylase have stunted growth with necrotic leaves and high PR1 gene expression. The maize lesion mimics a mutant, coding for uroporphyrinogen decarboxylase that has necrotic spots in the leaves. Inhibition of protox in Arabidopsis leads to production of lesion-mimic phenotype, high endogenous level of salicylic acid and PR1 gene expression.90 Overexpression of plastidic protox leads to resistance to the DPE herbicide acifluorfen. The overexpressed plants did not show any necrotic leaves.91 Tobacco plants having reduced ferrochelatase activity also show necrotic leaves in a light intensity dependent manner.92

Chlorophyll catabolic mutants

Intermediates involved in the Chl degradation pathway also produce ROS. Squash plants expressing the mature (lacking the N-terminal 21 amino acids) citrus chlorophyllase protein display a lesion-mimic phenotype when grown under natural light. The phenotype is caused by the accumulation of chlorophyllide, which is a photodynamic porphyrin molecule.93 The Arabidopsis pheophorbide a oxygenase (PAO, also called acd1, accelerated cell death 1) mutant shows a cell death phenotype because of the accumulation of the Chl degradation intermediate pheophorbide a. The latter gets photoactivated in the presence of light and generates ROS that form lesions in the mutant plants. The lesions that form in acd1 mutant leaves start mostly at the tip of the leaf and subsequently run down the leaf blade.94 Hirashima et al.95 also observed that the accumulation of pheophorbide a in dark-grown acd1 antisense plants caused cell death. LLS1 (lethal leaf spot 1), the homologue of ACD1 (Accelerated cell death 1/ Pheophorbide a oxygenase) in maize, is responsible for Chl catabolism. The maize lls1 mutant formed lesions when grown in the light.96 Similarly the Arabidopsis red chlorophyll catabolite reductase (RCCR, also called acd2, Accelearated cell deah 2) mutant showed lesion formation in leaves and spontaneous cell death phenotype.97 It is observed that the accumulation of H2O2 in the acd2 mitochondria is causal for its cell death phenotype.98 The lesion formation in acd2 is caused by the accumulation of red chlorophyll catabolite (RCC) in darkness that generates 1O2 in the presence of light.99 Further work should be done to check whether the generation of H2O2 and 1O2 are independent events or whether one leads to the other.

Abiotic stress-mediated ROS generation

One of the mechanisms that may lead to apoptosis upon extreme desiccation due to water-stress, salt-stress etc. is via ROS. In desiccated photosynthetic tissues, particular problems arise if excited Chl molecules are present but carbon fixation is limited by water deficiency. Under these conditions, electron flow continues, and excitation energy can be passed from photo-excited chlorophyll pigments to ground state oxygen (3O2) forming singlet oxygen (1O2). In addition, superoxide (O2-), hydrogen peroxides (H2O2) and the highly toxic hydroxyl radical (OH·) can be produced by photosystem II.100 ROS are also produced during normal metabolism in the electron transport chains of respiration and photosynthesis,101 but desiccation greatly enhances their production.102 To avoid 1O2 formation, photosynthetic organisms can dissipate excess energy via non-photochemical quenching (NPQ). This is most likely to involve the xanthophyll cycle in which violaxanthin is stepwise de-epoxidized to antheraxanthin and then to zeaxanthin, while solar radiation is dissipated as heat.103,104 In addition to their function as accessory pigments in photosynthesis, other carotenoids can also contribute to energy dissipation.105 Moreover, antioxidants such as glutathione, ascorbate, tocopherol, and related enzymes such as superoxide dismutase, catalase, peroxidase and others scavenge ROS.102

NADPH Oxidases

The respiratory burst oxidase homologues (Rbohs) oxidize cytosolic NADPH and transfer the electron to O2, thereby generating superoxide which is subsequently converted to H2O2. Ten RBOH genes are present in the Arabidopsis genome, they encode plasmamembrane-localized proteins, where the apoplastic oxidase domain is responsible for superoxide generation in the apoplast and the N-terminal extension in the cytoplasm contains the regulatory regions calcium-binding EF-hands and phosphorylation domains.106,107 Thus, there is a crosstalk between calcium and phosphorylation in controlling apoplastic ROS production.108 Furthermore, NADPH oxidase undergoes S-nitrosylation and this is required for enzyme activity.109,110 Rbohs are involved in defense against pathogen attacks, e.g. by activating hypersensitive responses and regulating innate immunity in plants111-114 and extracellular lignin production (115, and ref. therein), but also in nodule formation in Medicago truncatula,116 abiotic stress responses such as heat, cold, drought, salinity and high light.117,118 The ROS signal can be delivered from cell to cell through diffusion; and this long distance signals may participate in systemic acquired resistance.118,119 NADPH oxidase-mediated ROS production regulates polarized cell expansion in root hairs and pollen tip growth,120,121 or promotes seed ripening.122 RbohC and –D are highly expressed in shoot and roots and control most of the inducible ROS production; RbohC is particularly involved in root hair formation120 and mechanosensing,123 and RbohH and –J are specifically expressed in pollen.121

Oxidative Burst, ROS in Plant/Microbe Interaction

Most work has been performed for the role of NADPH oxidases in plant immunity. While ROS production increases rapidly after pathogen attack to establish local and systemic resistance Pogány et al.124 described a dual role of ROS and RbohD in the Arabidopsis-Alternaria pathosystem. RbohD triggers death in cells that are damaged by fungal infection but simultaneously inhibits death in neighboring cells through the suppression of free salicylic acid and ethylene levels. Comparably, Sclerotinia sclerotiorum, a necrotrophic fungus with a broad host range, produces the key pathogenicity factor oxalic acid, which induces apoptotic-like programmed cell death in plant hosts. This induction requires ROS generation in the host, a process triggered by fungal secreted oxalic acid. Conversely, during the initial stages of infection, oxalic acid also dampens the plant oxidative burst. This scenario presents a challenge regarding the mechanistic details of oxalic acid function; as oxalic acid both suppresses and induces host ROS during the compatible interaction. Initially, S. sclerotiorum via oxalic acid generates a reducing environment in host cells that suppress host defense responses including the oxidative burst and callose deposition, akin to compatible biotrophic pathogens. Once infection is established, however, this necrotroph induces the generation of plant ROS leading to programmed cell death of host tissue, the result of which is of direct benefit to the pathogen.125 These and other results indicate that reducing the cellular environment directly or indirectly suppresses the host-plant oxidative burst and defense mechanisms against pathogen infections.125 A survey of the literature also suggest that ROS levels (mainly or exclusively produced by NADPH oxidases) are higher after infections with necrotrophic compared to biotrophic pathogens, although this requires a more detailed analysis. In general, low antioxidant activity has been reported for plants exposed to necrotrophic pathogens while biotrophic pathogens activate the antioxidant system to propagate in the living plant cells. This is consistent with the observations that ROS production by plants infected with beneficial microbes is low and antioxidant enzyme activities high.126,127 Although ROS have been mainly associated with pathogen attack, ROS are also detected in other biotic interactions including beneficial symbiotic interactions with bacteria or mycorrhiza, suggesting that ROS production is a common feature of different biotic interactions.128 ROS might have different functions during various steps of the beneficial symbiosis, however it is generally excepted that ROS accumulates during early phases of the interaction when the beneficial symbiosis is not yet stable and declines thereafter.129 Whether this is caused by a general reduction in host defense processes once the two symbionts have been recognized as friends, or caused by an active repression from the microbe, is not known yet. Beneficial fungi or microbe are also known to protect plants against pathogens, e.g. by scavenging ROS, in particular when the plants are attacked by necrotrophic fungi. Finally, the important role of fungus-derived ROS for mycorrhiza formation has been nicely demonstrated for the Epichloë festuca/Lolium perenne symbiosis. Inactivation of the fungal NADPH oxidase-encoding noxA gene shifts the symbiosis from mutualistic to antagonistic, because fungal growth is no longer restricted.130

ROS and Signal Transduction

ROS sensors/receptors induce signaling cascades that lead to differential gene expression. This can occur directly through activating receptors, components of signaling pathways and/or even transcription factors in the nucleus. As ROS are short-lived, sensing must be quite efficient. Possible scenarios are: ROS are sensed at the apoplastic site of the plasma membrane (e.g. after the activation of NADPH oxidases), activate intercellular signaling and ultimately changes in gene expression pattern, or ROS activate signaling molecules within the cell or even organelles (e.g. by 1O2 generated in plastids). There might be a “linear” signal transduction pathway, but ROS may also influence the signaling events along such a proposed pathway at different levels. It is likely that different signaling pathways merge and/or interact with each other. In contrast, there might be a general effect of ROS on the cellular redox homeostasis, in particular if the ROS concentration in the cell is quite high. Since changes in organellar ROS levels, such as 1O2 in the plastids, lead to altered gene expression in the nucleus, ROS clearly initiate retrograde signaling from the plastids to the nucleus. However, there is also intensive cross-talk between the organelles themselves: for instance, under high light stress, mitochrondria re-oxidize excess reduction equivalents generated in the plastids to protect the plastids and its electron transfer chain from over-reduction.131 The utilization of mitochrondria to reduce photosynthesis-derived reduction equivalents is one strategy to reduce the production of ROS under stress. Mitochrondrial enzymes also form part of the respiratory pathway to salvage glycolate from photorespiration.

Plant cells sense ROS via at least 3 different mechanisms:

1. Unidentified receptor proteins

2. Redox sensitive compounds including transcription factors, such as NPR1 or HSFs

3. Direct inhibition of phosphatases by ROS

ROS are sensed via yet unknown mechanisms or receptors. All ROS with signaling functions, i.e. H2O2, 1O2, OH·, and O2·–132 must somehow interact with a cellular target to activate signaling events. Whether different ROS species are perceived by different perception mechanisms is unknown, however the specificity in the responses and the different locations, where ROS are generated, suggest different targets or perception mechanisms and therefore specificity. Downstream of the activation of putative receptors, it is likely that ROS perception is transduced via changes in the redox status of the signaling components. The information from the different ROS species can be either integrated or the signaling events are specific for individual ROS species.

To study apoplastic ROS sensing, the air pollutant ozone which generates ROS in the apoplast is a powerful tool. Different Arabidopsis ecotypes also show different sensitivities to ozone treatments, ranging from tolerant to extreme sensitive. Brosché et al.133 perform quantitative traits loci (QTL) mapping to identify genes that regulated ozone sensitivity in natural populations. The responses induced by ozone are quite similar to those induced by pathogens or pathogen-associated molecular patterns suggesting that ozone in the apoplast triggers hypersensitive-like programmed cell death responses. Several mutants altered in hormone biosynthesis or signalling showed also changes in ozone-induced transcriptional responses.134 Gauthier et al.135 have identified two members of the cysteine-rich receptor-like kinase gene family and a leucine-rich repeat (LRR-RLK) protein in Arabidopsis which are transcriptionally induced after ozone treatment, and thus they could sense ROS through redox modifications of their extracellular domain. Corresponding knock-out lines lead to sensitivity to extracellular ROS resulting in subtle lesion formation upon ozone treatment indicating their involvement in ROS signaling. Some of them are specifically upregulated by extracellular ROS but not by high light leading to the production of 1O2, suggesting specificity in the perception of different ROS species. Another candidate for ROS perception is the STIG1-like protein GRIM REAPER.136

Downstream signaling events associated with ROS sensing involve Ca2+ and Ca2+ binding protein such as calmodulin,137-139 the activation of G-protein140 and the activation of phospholipid signaling.141-143 A serine/threonine protein kinase (Oxidative Signal Inducible 1; OXI1) has been shown to play a central role in ROS sensing by activating mitogen–activated-protein kinases (MAPKs) 3 and -6 through Ca2+.142 The expression of OXI1 is induced in response to a wide range of H2O2-generating stimuli. OXI1 kinase activity is itself also induced by H2O2 in vivo.142 The OXI1 protein kinase is required for plant immunity against Pseudomonas syringae in Arabidopsis.144 OXI1, which belongs to the AGC kinases well characterized in mammalian systems,143 is important for two oxidative burst-mediated processes: basal resistance to microbial pathogens and root hair growth. To identify possible components of the OXI1 signalling pathway, phosphoproteomic techniques were used to detect alterations in the abundance of phosphorylated proteins and peptides in an oxi1 null mutant of Arabidopsis.145 Five proteins, including a putative F-box and 3-phosphoinositide-dependent kinase 1 (PDK1), show reduced phosphorylation in the oxi1 mutant, and may be direct or indirect targets of OXI1. Four proteins, including ethylene insensitive 2 and phospholipase Dγ, show increased phosphorylation in the oxi1 mutant. Therefore, these are putative early signaling candidate proteins in the OXI1 pathway. The diverse activities of these proteins, including protein degradation and hormone signalling, may suggest crosstalk between OXI1 and other signal transduction cascades.145 A MAPK cascade involving MAPK3/6 acts downstream of OXI1 and controls the activation of different defense mechanisms in response to ROS stresses.1,146 MAPKs are also activated by PDK1 through the phospholipase-C/D-phosphatidic-acid pathway.141 H2O2 activates several MAPKs.147 In Arabidopsis, H2O2 activates the MAPKs, MPK3, and MPK6 via MAPKKK ANP1.146 Overexpression of ANP1 in transgenic plants resulted in increased tolerance to heat shock, freezing and salt stress.146 H2O2 also increases expression of the Arabidopsis nucleotide diphosphate (NDP) kinase 2.148 Overexpression of AtNDPK2 reduced accumulation of H2O2 and enhanced tolerance to multiple stresses including cold, salt, and oxidative stress. The effect of NDPK2 might be mediated by the MAPKs, MPK3 and MPK6, because NDPK2 can interact and activate them.

The expression of different transcription factors is enhanced by ROS and includes members of the WRKY, Zat, RAV, GRAS and Myb families.149-154 In E. coli, the transcription factor OxyR,155 and in budding yeast, Yap1 play major roles156 in oxidative stress signaling. OxyR and Yap1 are redox-sensitive transcription factors and modulate gene expression in response to oxidative stress. Different types of ROS react with different cysteinyl residues and can give rise to differently modified products, possibly explaining how ROS species can induce different sets of genes via the same transcription factor.156,157 Microarray analysis of H2O2-induced gene expression indicates potential H2O2-responsive cis-elements in genes regulated by H2O2. Recent studies using knockout plants have revealed that the zinc-finger protein Zat12 is required for Apx1 expression and plant protection during oxidative stress, and that the highly conserved zinc finger paralogs LOL1 and LSD1 have antagonistic effects on SOD and O2 accumulation.152 The transmembrane sensory kinase functions through its capacity to autophosphorylate a histidine residue in response to the presence or absence of an external stimulus. The phosphoryl group is subsequently transferred from the histidine to an aspartate residue in the response regulator. The induced conformational change in the response regulator alters its DNA binding affinity and thus promotes gene regulation of certain promoters. Recent work has identified human protein tyrosine phosphatase PTP1B to be modified by H2O2 at the active site cysteine.158 A similar regulation likely occurs in plants because PTP1, an Arabidopsis PTP that can inactivate the Arabidopsis MPK6, can be inactivated by H2O2.159 Also, phosphatases involved in abscisic acid (ABA) signaling within guard cells have been identified whose in vitro activity was modulated reversibly by H2O2.160,161 Peleg-Grossman et al.162 have shown that cytoplasmic H2O2 prevents translocation of NPR1 to the nucleus and inhibits the induction of PR genes in Arabidopsis. NPR1 is a transcriptional coactivator and regulates salicylic acid-dependent gene expression.163 NPR1 protein exists in the cytosol in oligomeric form. During pathogen attack the NPR1 protein is reduced to its monomeric state and translocated to the nucleus, where it binds a bZIP transcription factor of the TGA/OBF family.164-166 The oligomeric state of the NPR1 protein is sustained by disulfide S-S bonds that involve Cys82 and Cys216.164 The formation and dissipation of disulfide bonds is regulated by redox changes, including ROS production and their degradation by antioxidants.167 Cytoplasmic ROS function maintains the NPR1 complex in its oligomeric state, excludes the NPR1 from the nucleus and thus inhibits PR gene expression.162 Once activated, the NPR1-TGA1 system is an ideal example for a redox- and ROS-regulated signaling pair involved in local and systemic defense repair.163 In addition, nitric oxide is a redox regulator of the NPR1/TGA1 system.168

ROS and Redox

Many cellular signaling events are based on redox reactions, therefore, it is likely that ROS is directly linked to or at least integrated into the cellular redox metabolism. ROS are harmful to cells because they cause the oxidation of lipids, proteins and DNA and other components in their environment. Consequently, it is important for a cell to maintain the redox homeostasis, e.g. by antioxidants (ascorbate and glutathione) and/or antioxidant enzyme systems. The antioxidants are predominantly maintained in the reduced state. Changes in the balance of reduced vs. oxidised forms of the antioxidants may be used as a sensor for changes in the environment, and changes in ROS levels may directly affect the redox situation in the cell. Elevated ROS levels may lead to the oxidation of the antioxidant systems and this results in changes in the redox balance in the cell or apoplast. However, it is not understood yet, how changes in ROS levels influences the redox situation and ultimately activate ROS/redox-induced signaling events. In the plastids, for instance, high light intensity induces ROS production and the reduction of the plastoquinone pool. Besides activation of protein kinases by the reduced plastoquinone pool and initation of a short-term photoacclimation, high light also induces changes in plastid and nuclear gene expression and this includes numerous genes involved in antioxidative processes and genes which code for ROS-scavenging enzymes.1,169 Whether these two processes (ROS production and reduction of plastoquinone) are linked to each other, is unknown.

Specificity of ROS effects

The ROS level is controlled by a complex protein network in all compartments to maintain redox balance. Mittler et al.169 have shown that this network includes at least 152 genes in Arabidopsis. Recent studies demonstrates that different ROS signals contain a certain degree of specificity, i.e. the response depends on the ROS that accumulates in or around a cell.170,171 Since chloroplasts, mitochondria and peroxisomes are the main ROS producers in a plant cell, the majority of the ROS-scavenging enzymes are located in these organelles. This includes the major cellular isoforms of the superoxide dismutases, ascorbate peroxidases and catalases, but also glutathione peroxidases, and peroxiredoxins (cf. 169). Thus besides the classical use of inhibitors or the isolation of mutants, manipulation of the antioxidant enzyme levels in specific cellular compartments by genetic tools can influence the accumulation of specific ROS in given cellular compartments.149,150,151,172,173 Arabidopsis plants compromised in specific antioxidant enzymes in combination with inhibitor studies and the flu mutant allows comparative analyses of targets of O2·–, H2O2, and 1O2.170 Powerful tools are comparative expression profile, proteome or metabolome analyses of these mutants or wild-type plants treated with ROS biosynthesis inhibitors.26,149,173-180 Examples are the flu mutant which generates 1O2 in plastids, the cat2 mutant which accumulates H2O2 in peroxisomes, or plants overexpressing glycolate oxidase in the plastids, which generates H2O2 in plastids. Another examples is the rbohC rbohD double knockout line which is strongly impaired in apoplastic ROS production in response to pathogen attack, or the rbohC mutant impaired in root hair development due to the lack of apoplastic ROS production in the appropriate cells.114 Considering the complex ROS scavenging network169 and gene families involved in ROS production, large number of new mutants, mutant combinations or overexpressor lines need to be analyzed in the future.169 There are also other problems in defining specificity in ROS signaling: For instance, transcriptional analyses from the flu mutant have shed some light on the specific effects of 1O2,26,181 but, unlike in the wild-type, the 1O2 production is not associated with excess excitation of the photosystem II.182 So far, these comparative studies clearly demonstrate that the cellular ROS responses depend on the type of ROS and the subcellular location where they are produced. The analyses also allowed the generation of a list of marker genes that were preferentially regulated by hydrogen peroxide, superoxide, or singlet oxygen of different cellular origin, whereas other genes were identified as general oxidative stress response markers (cf. 169).

ROS in the cytosol

Although the cytosol is not thought to be a major site of ROS production in plants is important in redox signal integration. Cytoplasmic NADPH is central for redox signaling because it generates reducing substrates for antioxidant enzymes that detoxify ROS. Furthermore, NADPH maintains the thiol/disulfide status and supplies electrons to ROS producing enzymes such as NADPH oxidases. Signaling from the apoplasts and from the organelles have to pass through the cytoplasm in order to alter gene expression in the nucleus. It is conceivable that ROS information, the redox state in the plant/cell is integrated into other regulators. An example for such a network is provided by Khandelwal et al.183 They generated a redox network to identify genes whose expression is tightly coordinated during adjustment of the cell to a new environment. In their experiments they applied stress to the plastids and identified ten subnetworks which participate in the re-adjustment of the cellular redox homeostasis to the new conditions. Many of the genes in these subnetworks code for cytoplasmic proteins. How these processes are coupled to ROS remains to be determined. Furthermore, Mhamdi et al.184 showed that different NADPH-dehydrogenases play non-overlapping roles in some stress responses. Thus, specificity from the different ROS species and locations overlap with the integration of ROS signaling into the redox homeostasis.

A major problem in agriculture is the effect of biotic and abiotic stress responses on plant growth and development. A common theme within these environmental factors is the perturbation of ROS homeostasis. The great economical importance highlights the importance of the ROS signaling network. In particular the genetic approaches undertaken in the last decade might substantially contribute to understand the underlying mechanisms. This may provide also a tool for genetic engineering of crop plants.

Glossary

Abbreviations:

ROS

reactive oxygen species

Footnotes

References

  • 1.Apel K, Hirt H. Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Annu Rev Plant Biol. 2004;55:373–99. doi: 10.1146/annurev.arplant.55.031903.141701. [DOI] [PubMed] [Google Scholar]
  • 2.Allan AC, Fluhr R. Two Distinct Sources of Elicited Reactive Oxygen Species in Tobacco Epidermal Cells. Plant Cell. 1997;9:1559–72. doi: 10.1105/tpc.9.9.1559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Hu X, Bidney DL, Yalpani N, Duvick JP, Crasta O, Folkerts O, et al. Overexpression of a gene encoding hydrogen peroxide-generating oxalate oxidase evokes defense responses in sunflower. Plant Physiol. 2003;133:170–81. doi: 10.1104/pp.103.024026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Walters DR. Polyamines and plant disease. Phytochemistry. 2003;64:97–107. doi: 10.1016/S0031-9422(03)00329-7. [DOI] [PubMed] [Google Scholar]
  • 5.Purvis AC. Role of the alternative oxidase in limiting superoxide production by plant mitochondria. Physiol Plant. 1997;100:165–70. doi: 10.1111/j.1399-3054.1997.tb03468.x. [DOI] [Google Scholar]
  • 6.Maxwell DP, Wang Y, McIntosh L. The alternative oxidase lowers mitochondrial reactive oxygen production in plant cells. Proc Natl Acad Sci U S A. 1999;96:8271–6. doi: 10.1073/pnas.96.14.8271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Robson CA, Vanlerberghe GC. Transgenic plant cells lacking mitochondrial alternative oxidase have increased susceptibility to mitochondria-dependent and -independent pathways of programmed cell death. Plant Physiol. 2002;129:1908–20. doi: 10.1104/pp.004853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Sweetlove LJ, Fell D, Fernie AR. Getting to grips with the plant metabolic network. Biochem J. 2008;409:27–41. doi: 10.1042/BJ20071115. [DOI] [PubMed] [Google Scholar]
  • 9.Skovsen E, Snyder JW, Lambert JD, Ogilby PR. Lifetime and diffusion of singlet oxygen in a cell. J Phys Chem B. 2005;109:8570–3. doi: 10.1021/jp051163i. [DOI] [PubMed] [Google Scholar]
  • 10.Frank HA, Young AJ, Britton G, Cogdell RJ, eds. The Photochemistry of Carotenoids. In: Advances in Photosynthesis [and Respiration], Volume 8, Kluwer Academic Publishers (now Springer), Dordrecht 1999. [Google Scholar]
  • 11.Barber J, Andersson B. Too much of a good thing: light can be bad for photosynthesis. Trends Biochem Sci. 1992;17:61–6. doi: 10.1016/0968-0004(92)90503-2. [DOI] [PubMed] [Google Scholar]
  • 12.Aro EM, Virgin I, Andersson B. Photoinhibition of Photosystem II. Inactivation, protein damage and turnover. Biochim Biophys Acta. 1993;1143:113–34. doi: 10.1016/0005-2728(93)90134-2. [DOI] [PubMed] [Google Scholar]
  • 13.Ohad I, Keren N, Zer H, Gong H, Mor TS. Light induced degradation of the photosystem II reaction center D1 protein in vivo: an integrative approach. Baker NR, Bowyer JR, eds. Photoinhibition of Photosynthesis: From Molecular Mechanisms to the Field. BIOS Scientific Publishers, Oxford 1994, pp. 161–177. [Google Scholar]
  • 14.Foote CS, Shook FC, Abakerli RB. Characterization of singlet oxygen. Methods Enzymol. 1984;105:36–47. doi: 10.1016/S0076-6879(84)05006-0. [DOI] [PubMed] [Google Scholar]
  • 15.Baroli I, Niyogi KK. Molecular genetics of xanthophyll-dependent photoprotection in green algae and plants. Philos Trans R Soc Lond B Biol Sci. 2000;355:1385–94. doi: 10.1098/rstb.2000.0700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Baroli I, Do AD, Yamane T, Niyogi KK. Zeaxanthin accumulation in the absence of a functional xanthophyll cycle protects Chlamydomonas reinhardtii from photooxidative stress. Plant Cell. 2003;15:992–1008. doi: 10.1105/tpc.010405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Baroli I, Gutman BL, Ledford HK, Shin JW, Chin BL, Havaux M, et al. Photo-oxidative stress in a xanthophyll-deficient mutant of Chlamydomonas. J Biol Chem. 2004;279:6337–44. doi: 10.1074/jbc.M312919200. [DOI] [PubMed] [Google Scholar]
  • 18.Davison PA, Hunter CN, Horton P. Overexpression of beta-carotene hydroxylase enhances stress tolerance in Arabidopsis. Nature. 2002;418:203–6. doi: 10.1038/nature00861. [DOI] [PubMed] [Google Scholar]
  • 19.Pogson BJ, Rissler HM. Genetic manipulation of carotenoid biosynthesis and photoprotection. Philos Trans R Soc Lond B Biol Sci. 2000;355:1395–403. doi: 10.1098/rstb.2000.0701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Macpherson AN, Telfer A, Barber J, Truscott TG. Direct detection of singlet oxygen from isolated photosystem II reaction centres. Biochim Biophys Acta. 1993;1143:301–9. doi: 10.1016/0005-2728(93)90201-P. [DOI] [Google Scholar]
  • 21.Chakraborty N, Tripathy BC. Involvement of singlet oxygen in photodynamic damage of isolated chloroplasts of cucumber (Cucumis sativus L.) cotyledons. Plant Physiol. 1992;98:7–11. doi: 10.1104/pp.98.1.7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Chakraborty N, Tripathy BC. 5-aminolevulinic acid-induced photodynamic reactions in thylakoid membranes of cucumber (Cucumis sativus L.) cotyledons. J Plant Biochem Biotechnol. 1992;1:65–8. [Google Scholar]
  • 23.Fryer MJ, Oxborough K, Mullineaux PM, Baker NR. Imaging of photo-oxidative stress responses in leaves. J Exp Bot. 2002;53:1249–54. doi: 10.1093/jexbot/53.372.1249. [DOI] [PubMed] [Google Scholar]
  • 24.Hideg E, Kálai T, Hideg K, Vass I. Photoinhibition of photosynthesis in vivo results in singlet oxygen production detection via nitroxide-induced fluorescence quenching in broad bean leaves. Biochemistry. 1998;37:11405–11. doi: 10.1021/bi972890+. [DOI] [PubMed] [Google Scholar]
  • 25.Suh HJ, Kim CS, Jung J. Cytochrome b6/f complex as an indigenous photodynamic generator of singlet oxygen in thylakoid membranes. Photochem Photobiol. 2000;71:103–9. doi: 10.1562/0031-8655(2000)0710103CBFCAA2.0.CO2. [DOI] [PubMed] [Google Scholar]
  • 26.op den Camp RG, Przybyla D, Ochsenbein C, Laloi C, Kim C, Danon A, et al. Rapid induction of distinct stress responses after the release of singlet oxygen in Arabidopsis. Plant Cell. 2003;15:2320–32. doi: 10.1105/tpc.014662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Tripathy BC, Mohapatra A and Gupta I. Impairment of the photosynthetic apparatus by oxidative stress induced by photosensitization reaction of protoporphyrin IX. Biochim Biophys Acta 2007; 1767: 860–868 [DOI] [PubMed]
  • 28.Jung S, Lee HJ, Lee Y, Kang K, Kim YS, Grimm B, et al. Toxic tetrapyrrole accumulation in protoporphyrinogen IX oxidase-overexpressing transgenic rice plants. Plant Mol Biol. 2008;67:535–46. doi: 10.1007/s11103-008-9338-0. [DOI] [PubMed] [Google Scholar]
  • 29.Lermontova I, Grimm B. Reduced activity of plastid protoporphyrinogen oxidase causes attenuated photodynamic damage during high-light compared to low-light exposure. Plant J. 2006;48:499–510. doi: 10.1111/j.1365-313X.2006.02894.x. [DOI] [PubMed] [Google Scholar]
  • 30.Mock HP, Grimm B. Reduction of uroporphyrinogen decarboxylase by antisense RNA expression affects activities of other enzymes involved in tetrapyrrole biosynthesis and leads to light dependent necrosis. Plant Physiol. 1997;113:1101–12. doi: 10.1104/pp.113.4.1101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Rebeiz CA. Montazer-ZouhoorA, Hopen HJ,Wu SM. Photodynamic herbicides: 1. Concept and phenomenology. Enzyme Microb Technol. 1984;6:390–401. doi: 10.1016/0141-0229(84)90012-7. [DOI] [Google Scholar]
  • 32.Rebeiz CA, Montazer-Zouhoor A, Mayasich JM, Tripathy BC, Wu SM, Rebeiz CC. Photodynamic herbicides. Recent development and molecular basis of selectivity. CRC Crit Rev Plant Sci. 1988;6:385–436. doi: 10.1080/07352688809382256. [DOI] [Google Scholar]
  • 33.Rebeiz CA, Nandihalli UB, Reddy KN. In: Topics in Photosynthesis -Volume 10, Herbicides. (Baker NR and Percival MP eds.), Elsevier, Amsterdam, 1991; 173–208. [Google Scholar]
  • 34.Shalygo NV, Mock HP, Averina NG, Grimm B. Photodynamic action of uroporphyrin and protochlorophyllide in greening barley leaves treated with cesium chloride. J Photochem Photobiol B. 1998;42:151–8. doi: 10.1016/S1011-1344(98)00067-0. [DOI] [PubMed] [Google Scholar]
  • 35.Tripathy BC, Chakraborty N. 5-Aminolevulinic Acid Induced Photodynamic Damage of the Photosynthetic Electron Transport Chain of Cucumber (Cucumis sativus L.) Cotyledons. Plant Physiol. 1991;96:761–7. doi: 10.1104/pp.96.3.761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Mohapatra A, Tripathy BC. Detection of protoporphyrin IX in envelope membranes of pea chloroplasts. Biochem Biophys Res Commun. 2002;299:751–4. doi: 10.1016/S0006-291X(02)02703-1. [DOI] [PubMed] [Google Scholar]
  • 37.Mohapatra A, Tripathy BC. Differential distribution of chlorophyll biosynthetic intermediates in stroma, envelope and thylakoid membranes in Beta vulgaris. Photosynth Res. 2007;94:401–10. doi: 10.1007/s11120-007-9209-6. [DOI] [PubMed] [Google Scholar]
  • 38.Tripathy BC, Mohapatra A, Pattanayak GK. Subplastidic distribution of chlorophyll biosynthetic intermediates and characterization of protochlorophyllide oxidoreductase C. in Agricultural applications in green chemistry (W M. Nelson ed, ACS ss 887.) American Chemical Society, Washington DC 2004; 117–128. [Google Scholar]
  • 39.Havaux M, Dall’osto L, Bassi R. Zeaxanthin has enhanced antioxidant capacity with respect to all other xanthophylls in Arabidopsis leaves and functions independent of binding to PSII antennae. Plant Physiol. 2007;145:1506–20. doi: 10.1104/pp.107.108480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Mozzo M, Dall’Osto L, Hienerwadel R, Bassi R, Croce R. Photoprotection in the antenna complexes of photosystem II: role of individual xanthophylls in chlorophyll triplet quenching. J Biol Chem. 2008;283:6184–92. doi: 10.1074/jbc.M708961200. [DOI] [PubMed] [Google Scholar]
  • 41.Daub ME, Hangarter RP. Light-induced production of singlet oxygen and superoxide by the fungal toxin, cercosporin. Plant Physiol. 1983;73:855–7. doi: 10.1104/pp.73.3.855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Haworth P, Hess FD. The generation of singlet oxygen (1O2) by the nitrodiphenyl ether herbicide oxyfluorfen is independent of photosynthesis. Plant Physiol. 1988;86:672–6. doi: 10.1104/pp.86.3.672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Stobart AK, Ameen-Bukhari I. Regulation of delta-aminolaevulinic acid synthesis and protochlorophyllide regeneration in the leaves of dark-grown barley (Hordeum vulgare) seedlings. Biochem J. 1984;222:419–26. doi: 10.1042/bj2220419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Stobart AK, Ameen-Bukhari I. Photoreduction of protochlorophyllide and its relationship to delta-aminolaevulinic acid synthesis in the leaves of dark-grown barley (Hordeum vulgare) seedlings. Biochem J. 1986;236:741–8. doi: 10.1042/bj2360741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Granik S. Magnesium porphyrins formed by barley seedlings treated with delta-aminolevulinic acid. Plant Physiol. 1959;34:XVIII. [Google Scholar]
  • 46.Matringe M, Camadro JM, Labbe P, Scalla R. Protoporphyrinogen oxidase as a molecular target for diphenyl ether herbicides. Biochem J. 1989;260:231–5. doi: 10.1042/bj2600231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Duke SO, Becerril JM, Sherman TD, Matsumoto H. Photosensitizing porphyrins as herbicides. In: Hedin PA (eds) Naturally Occurring Pest Bioregulators. ACS Symposium series No. 449, 371–386 1991; ACS, Washington. [Google Scholar]
  • 48.Kunert KJ, Sandmann G, Böger P. Mode of action of diphenyl ethers. Weed Sci. 1987;3:35–55. [Google Scholar]
  • 49.Lydon J, Duke SO. Porphyrin synthesis is required for photobleaching activity of the p-nitrosubstituted diphenyl ether herbicides. Pestic Biochem Physiol. 1988;31:74–83. doi: 10.1016/0048-3575(88)90031-4. [DOI] [Google Scholar]
  • 50.Ensminger MP, Hess FD. Action spectrum of the activity of aciflurofen-methyl, a diphenyl ether herbicide, in Chlamydomonas eugametos. Plant Physiol. 1985;77:503–5. doi: 10.1104/pp.77.2.503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Kunert KJ, Böger P. The bleaching effect of diphenyl ether oxyfluorefen. Weed Sci. 1981;29:169–72. [Google Scholar]
  • 52.Yoshimoto T, Matsunaka S. The mode of action of diphenyl ether herbicides: The scavenging of active oxygen is not inhibited by diphenyl ethers. In: “Abstracts of the 5th International Congress of Pesticide Chemistry” 1982, Vol. 5, p. 62. [Google Scholar]
  • 53.Gupta I, Tripathy BC. Oxidative stress in cucumber (Cucumis sativus L.) Seedlings treated with Acifluorfen. Z Naturforsch. 1999;54C:771–81. [PubMed] [Google Scholar]
  • 54.Gupta I, Tripathy BC. Oxidative stress in cucumber (Cucumis sativus L.) seedlings treated with acifluorfen. Indian J Biochem Biophys. 2000;37:498–505. [PubMed] [Google Scholar]
  • 55.Graham MY. The diphenylether herbicide lactofen induces cell death and expression of defense-related genes in soybean. Plant Physiol. 2005;139:1784–94. doi: 10.1104/pp.105.068676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Spikes DK, Bommer JC. Chlorophyll and related pigments as photosensitizers in biology and medicine. In: Scheer H (ed.) Chlorophylls, pp. 1181–1204, CRC Press, Boca Raton, FL 1991. [Google Scholar]
  • 57.Edge R, McGarvey DJ, Truscott TG. The carotenoids as anti-oxidants--a review. J Photochem Photobiol B. 1997;41:189–200. doi: 10.1016/S1011-1344(97)00092-4. [DOI] [PubMed] [Google Scholar]
  • 58.Havaux M, Eymery F, Porfirova S, Rey P, Dörmann P. Vitamin E protects against photoinhibition and photooxidative stress in Arabidopsis thaliana. Plant Cell. 2005;17:3451–69. doi: 10.1105/tpc.105.037036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Triantaphylidès C, Krischke M, Hoeberichts FA, Ksas B, Gresser G, Havaux M, et al. Singlet oxygen is the major reactive oxygen species involved in photooxidative damage to plants. Plant Physiol. 2008;148:960–8. doi: 10.1104/pp.108.125690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Trebst A. Function of beta-carotene and tocopherol in photosystem II. Z Naturforsch C. 2003;58:609–20. doi: 10.1515/znc-2003-9-1001. [DOI] [PubMed] [Google Scholar]
  • 61.Kramarenko GG, Hummel SG, Martin SM, Buettner GR. Ascorbate reacts with singlet oxygen to produce hydrogen peroxide. Photochem Photobiol. 2006;82:1634–7. doi: 10.1562/2006-01-12-RN-774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Bilski P, Li MY, Ehrenshaft M, Daub ME, Chignell CF. Vitamin B6 (pyridoxine) and its derivatives are efficient singlet oxygen quenchers and potential fungal antioxidants. Photochem Photobiol. 2000;71:129–34. doi: 10.1562/0031-8655(2000)071<0129:SIPVBP>2.0.CO;2. [DOI] [PubMed] [Google Scholar]
  • 63.Ehrenshaft M, Bilski P, Li MY, Chignell CF, Daub ME. A highly conserved sequence is a novel gene involved in de novo vitamin B6 biosynthesis. Proc Natl Acad Sci U S A. 1999;96:9374–8. doi: 10.1073/pnas.96.16.9374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Chen H, Xiong L. Pyridoxine is required for post-embryonic root development and tolerance to osmotic and oxidative stresses. Plant J. 2005;44:396–408. doi: 10.1111/j.1365-313X.2005.02538.x. [DOI] [PubMed] [Google Scholar]
  • 65.Danon A, Miersch O, Felix G, Camp RG, Apel K. Concurrent activation of cell death-regulating signaling pathways by singlet oxygen in Arabidopsis thaliana. Plant J. 2005;41:68–80. doi: 10.1111/j.1365-313X.2004.02276.x. [DOI] [PubMed] [Google Scholar]
  • 66.Affek HP, Yakir D. Protection by isoprene against singlet oxygen in leaves. Plant Physiol. 2002;129:269–77. doi: 10.1104/pp.010909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Agati G, Matteini P, Goti A, Tattini M. Chloroplast-located flavonoids can scavenge singlet oxygen. New Phytol. 2007;174:77–89. doi: 10.1111/j.1469-8137.2007.01986.x. [DOI] [PubMed] [Google Scholar]
  • 68.Nagai S, Ohara K, Mukai K. Kinetic study of the quenching reaction of singlet oxygen by flavonoids in ethanol solution. J Phys Chem B. 2005;109:4234–40. doi: 10.1021/jp0451389. [DOI] [PubMed] [Google Scholar]
  • 69.Velikova V, Edreva A, Loreto F. Endogenous isoprene protects Phragmites australis leaves against singlet oxygen. Physiol Plant. 2004;122:219–25. doi: 10.1111/j.0031-9317.2004.00392.x. [DOI] [Google Scholar]
  • 70.Schmidt K, Fufezan C, Krieger-Liszkay A, Satoh H, Paulsen H. Recombinant water-soluble chlorophyll protein from Brassica oleracea var. Botrys binds various chlorophyll derivatives. Biochemistry. 2003;42:7427–33. doi: 10.1021/bi034207r. [DOI] [PubMed] [Google Scholar]
  • 71.Nielsen OF. Macromolecular physiology of plastids. XII. Tigrina mutants in barley: genetic, spectroscopic and structural characterization. Hereditas. 1974;76:269–304. doi: 10.1111/j.1601-5223.1974.tb01345.x. [DOI] [PubMed] [Google Scholar]
  • 72.Von Wettstein D, Gough S, Kannangara CG. Chlorophyll Biosynthesis. Plant Cell. 1995;7:1039–57. doi: 10.1105/tpc.7.7.1039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Runge S, van Cleve B, Lebedev N, Armstrong G, Apel K. Isolation and classification of chlorophyll-deficient xantha mutants of Arabidopsis thaliana. Planta. 1995;197:490–500. doi: 10.1007/BF00196671. [DOI] [PubMed] [Google Scholar]
  • 74.Armstrong GA, Runge S, Frick G, Sperling U, Apel K. Identification of NADPH:protochlorophyllide oxidoreductases A and B: a branched pathway for light-dependent chlorophyll biosynthesis in Arabidopsis thaliana . . Plant Physiol. 1995;108:1505–17. doi: 10.1104/pp.108.4.1505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Frick G, Su Q, Apel K, Armstrong GA. An Arabidopsis porB porC double mutant lacking light-dependent NADPH:protochlorophyllide oxidoreductases B and C is highly chlorophyll-deficient and developmentally arrested. Plant J. 2003;35:141–53. doi: 10.1046/j.1365-313X.2003.01798.x. [DOI] [PubMed] [Google Scholar]
  • 76.Meskauskiene R, Apel K. Interaction of FLU, a negative regulator of tetrapyrrole biosynthesis, with the glutamyl-tRNA reductase requires the tetratricopeptide repeat domain of FLU. FEBS Lett. 2002;532:27–30. doi: 10.1016/S0014-5793(02)03617-7. [DOI] [PubMed] [Google Scholar]
  • 77.Lee KP, Kim C, Lee DW, Apel K. TIGRINA d, required for regulating the biosynthesis of tetrapyrroles in barley, is an ortholog of the FLU gene of Arabidopsis thaliana. FEBS Lett. 2003;553:119–24. doi: 10.1016/S0014-5793(03)00983-9. [DOI] [PubMed] [Google Scholar]
  • 78.Wagner D, Przybyla D, Op den Camp R, Kim C, Landgraf F, Lee KP, et al. The genetic basis of singlet oxygen-induced stress responses of Arabidopsis thaliana. Science. 2004;306:1183–5. doi: 10.1126/science.1103178. [DOI] [PubMed] [Google Scholar]
  • 79.Lee KP, Kim C, Landgraf F, Apel K. EXECUTER1- and EXECUTER2-dependent transfer of stress-related signals from the plastid to the nucleus of Arabidopsis thaliana. Proc Natl Acad Sci U S A. 2007;104:10270–5. doi: 10.1073/pnas.0702061104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Kim C, Lee KP, Baruah A, Nater M, Göbel C, Feussner I, et al. (1O2-mediated retrograde signaling during late embryogenesis predetermines plastid differentiation in seedlings by recruiting abscisic acid. Proc Natl Acad Sci U S A. 2009;106:9920–4. doi: 10.1073/pnas.0901315106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Coll NS, Danon A, Meurer J, Cho WK, Apel K. Characterization of soldat8, a suppressor of singlet oxygen-induced cell death in Arabidopsis seedlings. Plant Cell Physiol. 2009;50:707–18. doi: 10.1093/pcp/pcp036. [DOI] [PubMed] [Google Scholar]
  • 82.Baruah A, Simková K, Hincha DK, Apel K, Laloi C. Modulation of O-mediated retrograde signaling by the PLEIOTROPIC RESPONSE LOCUS 1 (PRL1) protein, a central integrator of stress and energy signaling. Plant J. 2009;60:22–32. doi: 10.1111/j.1365-313X.2009.03935.x. [DOI] [PubMed] [Google Scholar]
  • 83.Ishikawa A, Okamoto H, Iwasaki Y, Asahi T. A deficiency of coproporphyrinogen III oxidase causes lesion formation in Arabidopsis. Plant J. 2001;27:89–99. doi: 10.1046/j.1365-313x.2001.01058.x. [DOI] [PubMed] [Google Scholar]
  • 84.Kruse E, Mock HP, Grimm B. Reduction of coproporphyrinogen oxidase level by antisense RNA synthesis leads to deregulated gene expression of plastid proteins and affects the oxidative defense system. EMBO J. 1995;14:3712–20. doi: 10.1002/j.1460-2075.1995.tb00041.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Kruse E, Mock HP, Grimm B. Coproporphyrinogen III oxidase from barley and tobacco--sequence analysis and initial expression studies. Planta. 1995;196:796–803. doi: 10.1007/BF01106776. [DOI] [PubMed] [Google Scholar]
  • 86.Williams P, Hardeman K, Fowler J, Rivin C. Divergence of duplicated genes in maize: evolution of contrasting targeting information for enzymes in the porphyrin pathway. Plant J. 2006;45:727–39. doi: 10.1111/j.1365-313X.2005.02632.x. [DOI] [PubMed] [Google Scholar]
  • 87.Ayliffe MA, Agostino A, Clarke BC, Furbank R, von Caemmerer S, Pryor AJ. Suppression of the barley uroporphyrinogen III synthase gene by a Ds activation tagging element generates developmental photosensitivity. Plant Cell. 2009;21:814–31. doi: 10.1105/tpc.108.063685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Mock HP, Heller W, Molina A, Neubohn B, Sandermann H, Jr., Grimm B. Expression of uroporphyrinogen decarboxylase or coproporphyrinogen oxidase antisense RNA in tobacco induces pathogen defense responses conferring increased resistance to tobacco mosaic virus. J Biol Chem. 1999;274:4231–8. doi: 10.1074/jbc.274.7.4231. [DOI] [PubMed] [Google Scholar]
  • 89.Hu G, Yalpani N, Briggs SP, Johal GS. A porphyrin pathway impairment is responsible for the phenotype of a dominant disease lesion mimic mutant of maize. Plant Cell. 1998;10:1095–105. doi: 10.1105/tpc.10.7.1095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Molina A, Volrath S, Guyer D, Maleck K, Ryals J, Ward E. Inhibition of protoporphyrinogen oxidase expression in Arabidopsis causes a lesion-mimic phenotype that induces systemic acquired resistance. Plant J. 1999;17:667–78. doi: 10.1046/j.1365-313X.1999.00420.x. [DOI] [PubMed] [Google Scholar]
  • 91.Lermontova I, Grimm B. Overexpression of plastidic protoporphyrinogen IX oxidase leads to resistance to the diphenyl-ether herbicide acifluorfen. Plant Physiol. 2000;122:75–84. doi: 10.1104/pp.122.1.75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Papenbrock J, Mishra S, Mock HP, Kruse E, Schmidt EK, Petersmann A, et al. Impaired expression of the plastidic ferrochelatase by antisense RNA synthesis leads to a necrotic phenotype of transformed tobacco plants. Plant J. 2001;28:41–50. doi: 10.1046/j.1365-313X.2001.01126.x. [DOI] [PubMed] [Google Scholar]
  • 93.Harpaz-Saad S, Azoulay T, Arazi T, Ben-Yaakov E, Mett A, Shiboleth YM, et al. Chlorophyllase is a rate-limiting enzyme in chlorophyll catabolism and is posttranslationally regulated. Plant Cell. 2007;19:1007–22. doi: 10.1105/tpc.107.050633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Pruzinská A, Tanner G, Anders I, Roca M, Hörtensteiner S. Chlorophyll breakdown: pheophorbide a oxygenase is a Rieske-type iron-sulfur protein, encoded by the accelerated cell death 1 gene. Proc Natl Acad Sci U S A. 2003;100:15259–64. doi: 10.1073/pnas.2036571100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Hirashima M, Tanaka R, Tanaka A. Light-independent cell death induced by accumulation of pheophorbide a in Arabidopsis thaliana. Plant Cell Physiol. 2009;50:719–29. doi: 10.1093/pcp/pcp035. [DOI] [PubMed] [Google Scholar]
  • 96.Gray J, Close PS, Briggs SP, Johal GS. A novel suppressor of cell death in plants encoded by the Lls1 gene of maize. Cell. 1997;89:25–31. doi: 10.1016/S0092-8674(00)80179-8. [DOI] [PubMed] [Google Scholar]
  • 97.Mach JM, Castillo AR, Hoogstraten R, Greenberg JT. The Arabidopsis-accelerated cell death gene ACD2 encodes red chlorophyll catabolite reductase and suppresses the spread of disease symptoms. Proc Natl Acad Sci U S A. 2001;98:771–6. doi: 10.1073/pnas.98.2.771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Yao N, Greenberg JT. Arabidopsis ACCELERATED CELL DEATH2 modulates programmed cell death. Plant Cell. 2006;18:397–411. doi: 10.1105/tpc.105.036251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Pruzinská A, Anders I, Aubry S, Schenk N, Tapernoux-Lüthi E, Müller T, et al. In vivo participation of red chlorophyll catabolite reductase in chlorophyll breakdown. Plant Cell. 2007;19:369–87. doi: 10.1105/tpc.106.044404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.McKersie BD, Leshem YY. Stress and stress coping in cultivated plants Kluwer Academic Publishers, Dordrecht-Boston-London 1994, ISBN 0-7923-2827-2. [Google Scholar]
  • 101.Halliwell B. Toxic effects of oxygen on plant tissues. In: Chloroplast metabolism: the structure and function of chloroplasts in green leaf cells. Oxford: Clarendon Press, 1984: 180-202. [Google Scholar]
  • 102.Smirnoff N. The role of active oxygen in the response of plants to water deficit and desiccation. New Phytol. 1993;125:27–58. doi: 10.1111/j.1469-8137.1993.tb03863.x. [DOI] [PubMed] [Google Scholar]
  • 103.Siefermann-Harms D. Chlorophyll, carotenoids and the activity of the xanthophyll cycle. Environ Pollut. 1990;68:293–303. doi: 10.1016/0269-7491(90)90032-8. [DOI] [PubMed] [Google Scholar]
  • 104.Demmig-Adams B, Adams WW., III The role of xanthophyll cycle carotenoids in the protection of photosynthesis. Trends Plant Sci. 1996;1:21–7. doi: 10.1016/S1360-1385(96)80019-7. [DOI] [Google Scholar]
  • 105.Frank HA, Cua A, Chynwat V, Young AJ, Gosztola D, Wasielewski MR. The lifetimes and energies of the first excited singlet states of diadinoxanthin and diatoxanthin: the role of these molecules in excess energy dissipation in algae. Biochim Biophys Acta. 1996;1277:243–52. doi: 10.1016/S0005-2728(96)00106-5. [DOI] [PubMed] [Google Scholar]
  • 106.Ogasawara Y, Kaya H, Hiraoka G, Yumoto F, Kimura S, Kadota Y, et al. Synergistic activation of the Arabidopsis NADPH oxidase AtrbohD by Ca2+ and phosphorylation. J Biol Chem. 2008;283:8885–92. doi: 10.1074/jbc.M708106200. [DOI] [PubMed] [Google Scholar]
  • 107.Takeda S, Gapper C, Kaya H, Bell E, Kuchitsu K, Dolan L. Local positive feedback regulation determines cell shape in root hair cells. Science. 2008;319:1241–4. doi: 10.1126/science.1152505. [DOI] [PubMed] [Google Scholar]
  • 108.Kobayashi M, Ohura I, Kawakita K, Yokota N, Fujiwara M, Shimamoto K, et al. Calcium-dependent protein kinases regulate the production of reactive oxygen species by potato NADPH oxidase. Plant Cell. 2007;19:1065–80. doi: 10.1105/tpc.106.048884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Yun BW, Feechan A, Yin M, Saidi NB, Le Bihan T, Yu M, et al. S-nitrosylation of NADPH oxidase regulates cell death in plant immunity. Nature. 2011;478:264–8. doi: 10.1038/nature10427. [DOI] [PubMed] [Google Scholar]
  • 110.Lin A, Wang Y, Tang J, Xue P, Li C, Liu L, et al. Nitric oxide and protein S-nitrosylation are integral to hydrogen peroxide-induced leaf cell death in rice. Plant Physiol. 2012;158:451–64. doi: 10.1104/pp.111.184531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Torres MA, Dangl JL, Jones JD. Arabidopsis gp91phox homologues AtrbohD and AtrbohF are required for accumulation of reactive oxygen intermediates in the plant defense response. Proc Natl Acad Sci U S A. 2002;99:517–22. doi: 10.1073/pnas.012452499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Zhang JF, Shao F, Li Y, Cui H, Chen L, Li H, et al. A Pseudomonas syringae effector inactivates MAPKs to suppress PAMP-induced immunity in plants. Cell Host Microbe. 2007;1:175–85. doi: 10.1016/j.chom.2007.03.006. [DOI] [PubMed] [Google Scholar]
  • 113.Yoshioka H, Asai S, Yoshioka M, Kobayashi M. Molecular mechanisms of generation for nitric oxide and reactive oxygen species, and role of the radical burst in plant immunity. Mol Cells. 2009;28:321–9. doi: 10.1007/s10059-009-0156-2. [DOI] [PubMed] [Google Scholar]
  • 114.Proels RK, Oberhollenzer K, Pathuri IP, Hensel G, Kumlehn J, Hückelhoven R. RBOHF2 of barley is required for normal development of penetration resistance to the parasitic fungus Blumeria graminis f. sp. hordei. Mol Plant Microbe Interact. 2010;23:1143–50. doi: 10.1094/MPMI-23-9-1143. [DOI] [PubMed] [Google Scholar]
  • 115.Denness L, McKenna JF, Segonzac C, Wormit A, Madhou P, Bennett M, et al. Cell wall damage-induced lignin biosynthesis is regulated by a reactive oxygen species- and jasmonic acid-dependent process in Arabidopsis. Plant Physiol. 2011;156:1364–74. doi: 10.1104/pp.111.175737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Marino D, Andrio E, Danchin EG, Oger E, Gucciardo S, Lambert A, et al. A Medicago truncatula NADPH oxidase is involved in symbiotic nodule functioning. New Phytol. 2011;189:580–92. doi: 10.1111/j.1469-8137.2010.03509.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Kwak JM, Mori IC, Pei ZM, Leonhardt N, Torres MA, Dangl JL, et al. NADPH oxidase AtrbohD and AtrbohF genes function in ROS-dependent ABA signaling in Arabidopsis. EMBO J. 2003;22:2623–33. doi: 10.1093/emboj/cdg277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Miller G, Schlauch K, Tam R, Cortes D, Torres MA, Shulaev V, et al. The plant NADPH oxidase RBOHD mediates rapid systemic signaling in response to diverse stimuli. Sci Signal. 2009;2:ra45. doi: 10.1126/scisignal.2000448. [DOI] [PubMed] [Google Scholar]
  • 119.Sagi M, Davydov O, Orazova S, Yesbergenova Z, Ophir R, Stratmann JW, et al. Plant respiratory burst oxidase homologs impinge on wound responsiveness and development in Lycopersicon esculentum. Plant Cell. 2004;16:616–28. doi: 10.1105/tpc.019398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Foreman J, Demidchik V, Bothwell JH, Mylona P, Miedema H, Torres MA, et al. Reactive oxygen species produced by NADPH oxidase regulate plant cell growth. Nature. 2003;422:442–6. doi: 10.1038/nature01485. [DOI] [PubMed] [Google Scholar]
  • 121.Potocký M, Jones MA, Bezvoda R, Smirnoff N, Zárský V. Reactive oxygen species produced by NADPH oxidase are involved in pollen tube growth. New Phytol. 2007;174:742–51. doi: 10.1111/j.1469-8137.2007.02042.x. [DOI] [PubMed] [Google Scholar]
  • 122.Müller K, Carstens AC, Linkies A, Torres MA, Leubner-Metzger G. The NADPH-oxidase AtrbohB plays a role in Arabidopsis seed after-ripening. New Phytol. 2009;184:885–97. doi: 10.1111/j.1469-8137.2009.03005.x. [DOI] [PubMed] [Google Scholar]
  • 123.Monshausen GB, Bibikova TN, Weisenseel MH, Gilroy S. Ca2+ regulates reactive oxygen species production and pH during mechanosensing in Arabidopsis roots. Plant Cell. 2009;21:2341–56. doi: 10.1105/tpc.109.068395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Pogány M, von Rad U, Grün S, Dongó A, Pintye A, Simoneau P, et al. Dual roles of reactive oxygen species and NADPH oxidase RBOHD in an Arabidopsis-Alternaria pathosystem. Plant Physiol. 2009;151:1459–75. doi: 10.1104/pp.109.141994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Williams B, Kabbage M, Kim HJ, Britt R, Dickman MB. Tipping the balance: Sclerotinia sclerotiorum secreted oxalic acid suppresses host defenses by manipulating the host redox environment. PLoS Pathog. 2011;7:e1002107. doi: 10.1371/journal.ppat.1002107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Baltruschat H, Fodor J, Harrach BD, Niemczyk E, Barna B, Gullner G, et al. Salt tolerance of barley induced by the root endophyte Piriformospora indica is associated with a strong increase in antioxidants. . New Phytol. 2008;180:501–10. doi: 10.1111/j.1469-8137.2008.02583.x. [DOI] [PubMed] [Google Scholar]
  • 127.Vadassery J, Tripathi S, Prasad R, Varma A, Oelmüller R. Monodehydroascorbate reductase 2 and dehydroascorbate reductase 5 are crucial for a mutualistic interaction between Piriformospora indica and Arabidopsis. . J Plant Physiol. 2009;166:1263–74. doi: 10.1016/j.jplph.2008.12.016. [DOI] [PubMed] [Google Scholar]
  • 128.Torres MA. ROS in biotic interactions. Physiol Plant. 2010;138:414–29. doi: 10.1111/j.1399-3054.2009.01326.x. [DOI] [PubMed] [Google Scholar]
  • 129.Fester T, Hause G. Accumulation of reactive oxygen species in arbuscular mycorrhizal roots. Mycorrhiza. 2005;15:373–9. doi: 10.1007/s00572-005-0363-4. [DOI] [PubMed] [Google Scholar]
  • 130.Tanaka A, Christensen MJ, Takemoto D, Park P, Scott B. Reactive oxygen species play a role in regulating a fungus-perennial ryegrass mutualistic interaction. Plant Cell. 2006;18:1052–66. doi: 10.1105/tpc.105.039263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Padmasree K, Padmavathi L, Raghavendra AS. Essentiality of mitochondrial oxidative metabolism for photosynthesis: optimization of carbon assimilation and protection against photoinhibition. Crit Rev Biochem Mol Biol. 2002;37:71–119. doi: 10.1080/10409230290771465. [DOI] [PubMed] [Google Scholar]
  • 132.Laloi C, Apel K, Danon A. Reactive oxygen signalling: the latest news. Curr Opin Plant Biol. 2004;7:323–8. doi: 10.1016/j.pbi.2004.03.005. [DOI] [PubMed] [Google Scholar]
  • 133.Brosché M, Merilo E, Mayer F, Pechter P, Puzõrjova I, Brader G, et al. Natural variation in ozone sensitivity among Arabidopsis thaliana accessions and its relation to stomatal conductance. Plant Cell Environ. 2010;33:914–25. doi: 10.1111/j.1365-3040.2010.02116.x. [DOI] [PubMed] [Google Scholar]
  • 134.Wrzaczek M, Brosché M, Salojärvi J, Kangasjärvi S, Idänheimo N, Mersmann S, et al. Transcriptional regulation of the CRK/DUF26 group of receptor-like protein kinases by ozone and plant hormones in Arabidopsis. BMC Plant Biol. 2010;10:95. doi: 10.1186/1471-2229-10-95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Gauthier A, Idänheimo N, Brosché M, Kollist H, Wrzaczek M, Kangasjärvi J. Characterization of RLSs in Arabidopsis thaliana Proceedings of the 10th International Conference on Reactive Oxygen and Nitrogen Species in Plants, 2011; P5. [Google Scholar]
  • 136.Wrzaczek M, Brosché M, Kollist H, Kangasjärvi J. Arabidopsis GRI is involved in the regulation of cell death induced by extracellular ROS. Proc Natl Acad Sci U S A. 2009;106:5412–7. doi: 10.1073/pnas.0808980106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Knight H, Knight MR. Abiotic stress signalling pathways: specificity and cross-talk. Trends Plant Sci. 2001;6:262–7. doi: 10.1016/S1360-1385(01)01946-X. [DOI] [PubMed] [Google Scholar]
  • 138.Bowler C, Fluhr R. The role of calcium and activated oxygens as signals for controlling cross-tolerance. Trends Plant Sci. 2000;5:241–6. doi: 10.1016/S1360-1385(00)01628-9. [DOI] [PubMed] [Google Scholar]
  • 139.Coelho SM, Taylor AR, Ryan KP, Sousa-Pinto I, Brown MT, Brownlee C. Spatiotemporal patterning of reactive oxygen production and Ca2+ wave propagation in fucus rhizoid cells. Plant Cell. 2002;14:2369–81. doi: 10.1105/tpc.003285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Baxter-Burrell A, Yang Z, Springer PS, Bailey-Serres J. RopGAP4-dependent Rop GTPase rheostat control of Arabidopsis oxygen deprivation tolerance. Science. 2002;296:2026–8. doi: 10.1126/science.1071505. [DOI] [PubMed] [Google Scholar]
  • 141.Anthony RG, Henriques R, Helfer A, Mészáros T, Rios G, Testerink C, et al. A protein kinase target of a PDK1 signalling pathway is involved in root hair growth in Arabidopsis. EMBO J. 2004;23:572–81. doi: 10.1038/sj.emboj.7600068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Rentel MC, Lecourieux D, Ouaked F, Usher SL, Petersen L, Okamoto H, et al. OXI1 kinase is necessary for oxidative burst-mediated signalling in Arabidopsis. Nature. 2004;427:858–61. doi: 10.1038/nature02353. [DOI] [PubMed] [Google Scholar]
  • 143.Hirt H, Garcia AV, Oelmüller R. AGC kinases in plant development and defense. Plant Signal Behav. 2011;6:1030–3. doi: 10.4161/psb.6.7.15580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Petersen LN, Ingle RA, Knight MR, Denby KJ. OXI1 protein kinase is required for plant immunity against Pseudomonas syringae in Arabidopsis. J Exp Bot. 2009;60:3727–35. doi: 10.1093/jxb/erp219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Howden AJ, Salek M, Miguet L, Pullen M, Thomas B, Knight MR, et al. The phosphoproteome of Arabidopsis plants lacking the oxidative signal-inducible1 (OXI1) protein kinase. New Phytol. 2010 doi: 10.1111/j.1469-8137. [DOI] [PubMed] [Google Scholar]
  • 146.Kovtun Y, Chiu W-L, Tena G, Sheen J. Functional analysis of oxidative stress-activated mitogen-activated protein kinase cascade in plants. Proc Natl Acad Sci U S A. 2000;97:2940–5. doi: 10.1073/pnas.97.6.2940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Jonak C, Okrész L, Bögre L, Hirt H. Complexity, cross talk and integration of plant MAP kinase signalling. Curr Opin Plant Biol. 2002;5:415–24. doi: 10.1016/S1369-5266(02)00285-6. [DOI] [PubMed] [Google Scholar]
  • 148.Moon H, Lee B, Choi G, Shin D, Prasad DT, Lee O, et al. NDP kinase 2 interacts with two oxidative stress-activated MAPKs to regulate cellular redox state and enhances multiple stress tolerance in transgenic plants. Proc Natl Acad Sci U S A. 2003;100:358–63. doi: 10.1073/pnas.252641899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Rizhsky L, Liang H, Mittler R. The water-water cycle is essential for chloroplast protection in the absence of stress. J Biol Chem. 2003;278:38921–5. doi: 10.1074/jbc.M304987200. [DOI] [PubMed] [Google Scholar]
  • 150.Pnueli L, Liang H, Rozenberg M, Mittler R. Growth suppression, altered stomatal responses, and augmented induction of heat shock proteins in cytosolic ascorbate peroxidase (Apx1)-deficient Arabidopsis plants. Plant J. 2003;34:187–203. doi: 10.1046/j.1365-313X.2003.01715.x. [DOI] [PubMed] [Google Scholar]
  • 151.Rizhsky L, Davletova S, Liang H, Mittler R. The zinc finger protein Zat12 is required for cytosolic ascorbate peroxidase 1 expression during oxidative stress in Arabidopsis. J Biol Chem. 2004;279:11736–43. doi: 10.1074/jbc.M313350200. [DOI] [PubMed] [Google Scholar]
  • 152.Epple P, Mack AA, Morris VR, Dangl JL. Antagonistic control of oxidative stress-induced cell death in Arabidopsis by two related, plant-specific zinc finger proteins. Proc Natl Acad Sci U S A. 2003;100:6831–6. doi: 10.1073/pnas.1130421100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Vranová E, Atichartpongkul S, Villarroel R, Van Montagu M, Inzé D, Van Camp W. Comprehensive analysis of gene expression in Nicotiana tabacum leaves acclimated to oxidative stress. . Proc Natl Acad Sci U S A. 2002;99:10870–5. doi: 10.1073/pnas.152337999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Desikan R, A-H-Mackerness S, Hancock JT, Neill SJ. Regulation of the Arabidopsis transcriptome by oxidative stress. Plant Physiol. 2001;127:159–72. doi: 10.1104/pp.127.1.159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Zheng M, Aslund F, Storz G. Activation of the OxyR transcription factor by reversible disulfide bond formation. Science. 1998;279:1718–21. doi: 10.1126/science.279.5357.1718. [DOI] [PubMed] [Google Scholar]
  • 156.Delaunay A, Isnard AD, Toledano MB. H2O2 sensing through oxidation of the Yap1 transcription factor. EMBO J. 2000;19:5157–66. doi: 10.1093/emboj/19.19.5157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Delaunay A, Pflieger D, Barrault MB, Vinh J, Toledano MB. A thiol peroxidase is an H2O2 receptor and redox-transducer in gene activation. Cell. 2002;111:471–81. doi: 10.1016/S0092-8674(02)01048-6. [DOI] [PubMed] [Google Scholar]
  • 158.van Montfort RL, Congreve M, Tisi D, Carr R, Jhoti H. Oxidation state of the active-site cysteine in protein tyrosine phosphatase 1B. Nature. 2003;423:773–7. doi: 10.1038/nature01681. [DOI] [PubMed] [Google Scholar]
  • 159.Gupta R, Luan S. Redox control of protein tyrosine phosphatases and mitogen-activated protein kinases in plants. Plant Physiol. 2003;132:1149–52. doi: 10.1104/pp.103.020792. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Meinhard M, Grill E. Hydrogen peroxide is a regulator of ABI1, a protein phosphatase 2C from Arabidopsis. FEBS Lett. 2001;508:443–6. doi: 10.1016/S0014-5793(01)03106-4. [DOI] [PubMed] [Google Scholar]
  • 161.Meinhard M, Rodriguez PL, Grill E. The sensitivity of ABI2 to hydrogen peroxide links the abscisic acid-response regulator to redox signalling. Planta. 2002;214:775–82. doi: 10.1007/s00425-001-0675-3. [DOI] [PubMed] [Google Scholar]
  • 162.Peleg-Grossman S, Melamed-Book N, Cohen G, Levine A. Cytoplasmic H2O2 prevents translocation of NPR1 to the nucleus and inhibits the induction of PR genes in Arabidopsis. . Plant Signal Behav. 2010;5:1401–6. doi: 10.4161/psb.5.11.13209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Rochon A, Boyle P, Wignes T, Fobert PR, Després C. The coactivator function of Arabidopsis NPR1 requires the core of its BTB/POZ domain and the oxidation of C-terminal cysteines. Plant Cell. 2006;18:3670–85. doi: 10.1105/tpc.106.046953. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Mou Z, Fan W, Dong X. Inducers of plant systemic acquired resistance regulate NPR1 function through redox changes. Cell. 2003;113:935–44. doi: 10.1016/S0092-8674(03)00429-X. [DOI] [PubMed] [Google Scholar]
  • 165.Després C, DeLong C, Glaze S, Liu E, Fobert PR. The Arabidopsis NPR1/NIM1 protein enhances the DNA binding activity of a subgroup of the TGA family of bZIP transcription factors. Plant Cell. 2000;12:279–90. [PMC free article] [PubMed] [Google Scholar]
  • 166.Zhang Y, Fan W, Kinkema M, Li X, Dong X. Interaction of NPR1 with basic leucine zipper protein transcription factors that bind sequences required for salicylic acid induction of the PR-1 gene. Proc Natl Acad Sci U S A. 1999;96:6523–8. doi: 10.1073/pnas.96.11.6523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Cumming RC, Andon NL, Haynes PA, Park M, Fischer WH, Schubert D. Protein disulfide bond formation in the cytoplasm during oxidative stress. J Biol Chem. 2004;279:21749–58. doi: 10.1074/jbc.M312267200. [DOI] [PubMed] [Google Scholar]
  • 168.Lindermayr C, Sell S, Müller B, Leister D, Durner J. Redox regulation of the NPR1-TGA1 system of Arabidopsis thaliana by nitric oxide. Plant Cell. 2010;22:2894–907. doi: 10.1105/tpc.109.066464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Mittler R, Vanderauwera S, Gollery M, Van Breusegem F. Reactive oxygen gene network of plants. Trends Plant Sci. 2004;9:490–8. doi: 10.1016/j.tplants.2004.08.009. [DOI] [PubMed] [Google Scholar]
  • 170.Gadjev I, Vanderauwera S, Gechev TS, Laloi C, Minkov IN, Shulaev V, et al. Transcriptomic footprints disclose specificity of reactive oxygen species signaling in Arabidopsis. Plant Physiol. 2006;141:436–45. doi: 10.1104/pp.106.078717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Mehterov N, Balazadeh S, Hille J, Toneva V, Mueller-Roeber B, Gechev T. Oxidative stress provokes distinct transcriptional responses in the stress-tolerant atr7 and stress-sensitive loh2Arabidopsis thaliana mutants as revealed by multi-parallel quantitative real-time PCR analysis of ROS marker and antioxidant genes. . Plant Physiol Biochem. 2012;59:20–9. doi: 10.1016/j.plaphy.2012.05.024. [DOI] [PubMed] [Google Scholar]
  • 172.Meskauskiene R, Nater M, Goslings D, Kessler F, op den Camp R, Apel K. FLU: a negative regulator of chlorophyll biosynthesis in Arabidopsis thaliana. Proc Natl Acad Sci U S A. 2001;98:12826–31. doi: 10.1073/pnas.221252798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Vandenabeele S, Vanderauwera S, Vuylsteke M, Rombauts S, Langebartels C, Seidlitz HK, et al. Catalase deficiency drastically affects gene expression induced by high light in Arabidopsis thaliana. Plant J. 2004;39:45–58. doi: 10.1111/j.1365-313X.2004.02105.x. [DOI] [PubMed] [Google Scholar]
  • 174.Davletova S, Rizhsky L, Liang H, Shengqiang Z, Oliver DJ, Coutu J, et al. Cytosolic ascorbate peroxidase 1 is a central component of the reactive oxygen gene network of Arabidopsis. Plant Cell. 2005;17:268–81. doi: 10.1105/tpc.104.026971. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Gechev TS, Gadjev IZ, Hille J. An extensive microarray analysis of AAL-toxin-induced cell death in Arabidopsis thaliana brings new insights into the complexity of programmed cell death in plants. Cell Mol Life Sci. 2004;61:1185–97. doi: 10.1007/s00018-004-4067-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Gechev TS, Hille J. Hydrogen peroxide as a signal controlling plant programmed cell death. J Cell Biol. 2005;168:17–20. doi: 10.1083/jcb.200409170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Gechev TS, Minkov IN, Hille J. Hydrogen peroxide-induced cell death in Arabidopsis: transcriptional and mutant analysis reveals a role of an oxoglutarate-dependent dioxygenase gene in the cell death process. IUBMB Life. 2005;57:181–8. doi: 10.1080/15216540500090793. [DOI] [PubMed] [Google Scholar]
  • 178.Vanderauwera S, Zimmermann P, Rombauts S, Vandenabeele S, Langebartels C, Gruissem W, et al. Genome-wide analysis of hydrogen peroxide-regulated gene expression in Arabidopsis reveals a high light-induced transcriptional cluster involved in anthocyanin biosynthesis. Plant Physiol. 2005;139:806–21. doi: 10.1104/pp.105.065896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Umbach AL, Fiorani F, Siedow JN. Characterization of transformed Arabidopsis with altered alternative oxidase levels and analysis of effects on reactive oxygen species in tissue. Plant Physiol. 2005;139:1806–20. doi: 10.1104/pp.105.070763. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Fahnenstich H, Scarpeci TE, Valle EM, Flügge UI, Maurino VG. Generation of hydrogen peroxide in chloroplasts of Arabidopsis overexpressing glycolate oxidase as an inducible system to study oxidative stress. Plant Physiol. 2008;148:719–29. doi: 10.1104/pp.108.126789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Laloi C, Stachowiak M, Pers-Kamczyc E, Warzych E, Murgia I, Apel K. Cross-talk between singlet oxygen- and hydrogen peroxide-dependent signaling of stress responses in Arabidopsis thaliana. Proc Natl Acad Sci U S A. 2007;104:672–7. doi: 10.1073/pnas.0609063103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Mullineaux PM, Baker NR. Oxidative stress: antagonistic signaling for acclimation or cell death? Plant Physiol. 2010;154:521–5. doi: 10.1104/pp.110.161406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Khandelwal A, Elvitigala T, Ghosh B, Quatrano RS. Arabidopsis transcriptome reveals control circuits regulating redox homeostasis and the role of an AP2 transcription factor. Plant Physiol. 2008;148:2050–8. doi: 10.1104/pp.108.128488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Mhamdi A, Mauve C, Gouia H, Saindrenan P, Hodges M, Noctor G. Cytosolic NADP-dependent isocitrate dehydrogenase contributes to redox homeostasis and the regulation of pathogen responses in Arabidopsis leaves. Plant Cell Environ. 2010;33:1112–23. doi: 10.1111/j.1365-3040.2010.02133.x. [DOI] [PubMed] [Google Scholar]

Articles from Plant Signaling & Behavior are provided here courtesy of Taylor & Francis

RESOURCES