Reactive Oxygen Species: Reactions and Detection from Photosynthetic Tissues Heta Mattila, Sergey Khorobrykh, Vesa Havurinne, Esa Tyystj¨arvi PII: DOI: Reference:
S1011-1344(15)00322-X doi: 10.1016/j.jphotobiol.2015.10.001 JPB 10158
To appear in: Received date: Revised date: Accepted date:
11 June 2015 30 September 2015 1 October 2015
Please cite this article as: Heta Mattila, Sergey Khorobrykh, Vesa Havurinne, Esa Tyystj¨ arvi, Reactive Oxygen Species: Reactions and Detection from Photosynthetic Tissues, (2015), doi: 10.1016/j.jphotobiol.2015.10.001
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT 1
RI P
Heta Mattila, Sergey Khorobrykh, Vesa Havurinne and Esa Tyystjärvi*
T
Reactive Oxygen Species: Reactions and Detection from Photosynthetic Tissues Department of Biochemistry / Molecular Plant Biology, University of Turku, 20014 Turku, Finland
AC CE
PT
ED
MA
NU
SC
*Corresponding author, e-mail
[email protected]
ACCEPTED MANUSCRIPT 2
Abstract
AC CE
PT
ED
MA
NU
SC
RI P
T
Reactive oxygen species (ROS) have long been recognized as compounds with dual roles. They cause cellular damage by reacting with biomolecules but they also function as agents of cellular signaling. Several different oxygen-containing compounds are classified as ROS because they react, at least with certain partners, more rapidly than ground-state molecular oxygen or because they are known to have biological effects. The present review describes the typical reactions of the most important ROS. The reactions are the basis for both the detection methods and for prediction of reactions between ROS and biomolecules. Chemical and physical methods used for detection, visualization and quantification of ROS from plants, algae and cyanobacteria will be reviewed. The main focus will be on photosynthetic tissues, and limitations of the methods will be discussed.
ACCEPTED MANUSCRIPT 3
Contents 1. Reactive oxygen species (ROS) and free radicals ................................................................................ 5 1.1. Properties of ROS ......................................................................................................................... 5
T
1.2. General principles of detection and monitoring of ROS .............................................................. 7
RI P
2. Singlet oxygen, 1O2 .............................................................................................................................. 8 2.1. Definition and properties of 1O2 .................................................................................................. 8 2.2. Formation and action of 1O2 in photosynthetic tissues ............................................................... 8
SC
2.3. Reactions of 1O2 ........................................................................................................................... 9 2.3.1. Physical deactivation of 1O2 by radiative and non-radiative mechanisms............................ 9
NU
2.3.2. Chemical reactions of 1O2 ................................................................................................... 10 2.4. Lifetime and diffusion distance of 1O2 ....................................................................................... 12 2.5. Detection of 1O2 ......................................................................................................................... 17
MA
2.5.1. Luminescence at 1270 nm .................................................................................................. 17 2.5.2. EPR detectable probes ........................................................................................................ 17
ED
2.5.3. Dyes ..................................................................................................................................... 18 2.5.4. Fluorescent probes ............................................................................................................. 18 2.5.5. Oxygen consumption in a reaction between 1O2 and histidine .......................................... 19
PT
2.5.6. Genetically encoded probes ............................................................................................... 19 2.5.7. Further gene expression methods ...................................................................................... 23
AC CE
2.5.8. Indirect methods of 1O2 detection ...................................................................................... 23 2.5.9. Summary of 1O2 detection .................................................................................................. 24 3. Hydrogen peroxide, H2O2 .................................................................................................................. 24 3.1. Definition and properties of H2O2 .............................................................................................. 24 3.2. Formation and action of H2O2 in photosynthetic tissues........................................................... 24 3.3. Reactions of H2O2 ....................................................................................................................... 25 3.4. Lifetime and diffusion distance of H2O2 ..................................................................................... 26 3.5. Detection of H2O2 ....................................................................................................................... 26 3.5.1. Principles of H2O2 detection................................................................................................ 26 3.5.2. Precipitate-forming compounds ......................................................................................... 26 3.5.3. Dyes ..................................................................................................................................... 27 3.5.4. Fluorescent probes ............................................................................................................. 27 3.5.5. Chemiluminescent probe .................................................................................................... 28 3.5.6. Microelectrodes .................................................................................................................. 28 3.5.7. Genetically encoded probes ............................................................................................... 29
ACCEPTED MANUSCRIPT 4
3.5.8. Unspecific probes................................................................................................................ 29 3.5.9. Indirect methods of H2O2 detection ................................................................................... 29 3.5.10. Summary of H2O2 detection .............................................................................................. 29
T
4. Superoxide anion radical, O2•, and hydroperoxyl radical, HO2• ...................................................... 33
RI P
4.1. Definitions and properties of O2• and HO2• .............................................................................. 33 4.2. Formation and action of O2• and HO2• in photosynthetic tissues ............................................ 33 4.3. Reactions of O2• and HO2• ........................................................................................................ 33
SC
4.4. Lifetime and diffusion distance of O2• ...................................................................................... 36 4.5. Detection methods of O2• ......................................................................................................... 37
NU
4.5.1. Spectroscopic methods ....................................................................................................... 37 4.5.2. EPR detectable probes ........................................................................................................ 37
MA
4.5.3. Dyes ..................................................................................................................................... 38 4.5.4. Fluorescent probes ............................................................................................................. 40 4.5.5. Chemiluminescent probes .................................................................................................. 40
ED
4.5.6. Microelectrodes .................................................................................................................. 41 4.5.7. Genetically encoded probes ............................................................................................... 41
PT
4.5.8. Indirect methods of O2• measurement............................................................................. 41 4.5.9. Summary of O2• detection ................................................................................................. 41 5. Hydroxyl radical, HO• ........................................................................................................................ 46
AC CE
5.1. Definition and properties of HO• ............................................................................................... 46 5.2. Formation and action of HO• in photosynthetic tissues ............................................................ 46 5.3. Reactions of HO• ........................................................................................................................ 47 5.4. Lifetime and diffusion distance of HO• ...................................................................................... 48 5.5. Detection methods of HO• ......................................................................................................... 48 5.5.1 Spectroscopic methods ........................................................................................................ 48 5.5.2. EPR detectable probes ........................................................................................................ 49 5.5.3. Fluorescent probes ............................................................................................................. 50 5.5.4. Microelectrode.................................................................................................................... 51 5.5.5. Summary of HO• detection ................................................................................................. 51 6. Peroxyl radical (ROO•), alkoxyl radical (RO•) and hydroperoxides ................................................... 54 6.1. Definition and formation ........................................................................................................... 54 6.2. Reactions of ROO•, RO• and hydroperoxides ............................................................................. 54 6.3. Lifetime and diffusion distance of ROOHs ................................................................................. 56 6.4. Detection methods of ROOHs .................................................................................................... 56
ACCEPTED MANUSCRIPT 5
6.4.1. Spectroscopic methods ....................................................................................................... 56 6.4.2. Chemiluminescence and thermoluminescence .................................................................. 56 6.4.3. Dyes ..................................................................................................................................... 57
T
6.4.4. Fluorescent probes ............................................................................................................. 57
RI P
6.4.5. Measurements of ROOH-derived species ........................................................................... 58 7. Ozone, O3 .......................................................................................................................................... 61 8. Non-specific methods ....................................................................................................................... 63
SC
8.1. Non-specific methods indicating the presence of ROS .............................................................. 63 8.2. Non-specific methods indicating reactions of ROS with biomolecules ..................................... 64
NU
9. Concluding remarks .......................................................................................................................... 64 10. Acknowledgements......................................................................................................................... 65
MA
11. References ...................................................................................................................................... 65
1.1. Properties of ROS
ED
1. Reactive oxygen species (ROS) and free radicals
AC CE
PT
The term "reactive oxygen species" refers to naturally occurring, oxygen-containing chemical species whose reactivity, at least towards some substances, is higher than that of ground-state oxygen [1]. In the present review, we will limit the discussion to singlet oxygen, superoxide (superoxide anion radical and hydroperoxyl radical), hydrogen peroxide, hydroxyl radical, ozone, and lipid peroxyl radicals, lipid peroxides and alcoxides. ROS are inevitably produced in photosynthetic organisms [2,3] and cause oxidative damage by reacting with biomolecules [1,4‒7], but ROS also function as cellular signals [8‒10]. Photosynthetic organisms have extensive quenching and scavenging mechanisms that alleviate the effects of ROS [11‒14]. Free radicals are defined as chemical species that contain one or more unpaired electrons and are capable of independent existence [1]. Ground-state O2 is classified as a free radical (a diradical) because the two degenerate outermost molecular orbitals are occupied by one electron each, and in accordance to Hund's rule, the spins of these two electrons are parallel in the lowest energy state of the system. Superoxide anion radical (O2•), hydroxyl radical (HO•) and perhydroxyl radical (HO2•) are also free radicals whereas singlet oxygen (1O2), hydrogen peroxide (H2O2), ozone (O3), peroxides and alcoxides are not radicals. Some free radicals (like HO• and the hydrogen radical H•) are extremely reactive, whereas some free radicals, like ions of transition metals, exhibit weak reactivity. Free radicals are affected by external magnetic fields because the net spin magnetic momentum of the unpaired electrons interacts with a magnetic field. The energy of a free radical depends on the spin configuration and on the strength of the magnetic field. Free radicals are therefore visible in electron paramagnetic resonance (EPR, also called electron spin resonance, ESR) spectroscopy. A radical with one unpaired electron (doublet, total spin 1/2) has two energy states (designated as the spin up and the spin down states) in a magnetic field, and a free radical with two unpaired electrons
ACCEPTED MANUSCRIPT 6
(triplet, total spin 1), has three energy levels. The total spin of a molecule containing only paired electrons is zero, and such a molecule is described as a singlet state. Singlets interact with external magnetic fields only weakly.
RI P
T
In addition to the classification as radicals and non-radicals, ROS can be roughly divided into two types: (1) Free ROS, small molecules composed of oxygen and hydrogen only, and (2) incorporated ROS in which oxygen is bound to other molecules to form reactive oxygen derivatives. Table 1 presents the most important ROS. The most important reactive nitrogen species (for a review, see [15]) are listed in Table 1 but will not be further discussed.
SC
Table 1. Most significant ROS and reactive nitrogen species. R is a residual of an organic molecule. Non-radicals Free ROS
NU
Radicals
Singlet oxygen, 1O2
Superoxide anion radical, O2• Perhydroxyl radical, HO2•
MA
Hydroxyl radical, HO•
Hydrogen peroxide, H2O2
Ozone, O3 Incorporated ROS and reactive nitrogen species Organic peroxides, ROOH Peroxynitrite ion, ONOO Alkyl peroxynitrite, ROONO
PT
ED
Peroxyl radical, ROO• Alkoxyl radical, RO• Nitric Oxide, NO• Nitrogen dioxide,NO2•
AC CE
The biological actions of ROS are selective because the reactivity of each ROS strongly depends on the ROS and on reaction conditions. Some ROS like HO• have a very high chemical activity and are able to oxidize most organic molecules. On the other hand, H2O2 has moderate reactivity and is able to migrate far from site of production. The general reactivity of the main ROS in biological environment decreases in the order HO• > 1O2 > H2O2 > O2•, although care should be taken in using this generalization, as the rate of reaction depends on the reaction partner. Ability to interconverse is an essential property of ROS. For example, 1-electron reduction of H2O2 leads to the formation of HO•, indicating that formation of a less reactive form can promote the formation of more reactive forms. Due to the interconversions, cellular effects of ROS should be considered as complex sets of reactions. Given these reactions, action mechanisms of oxygen in the cell can be understood. Production, detoxification and detection of ROS, as well as the roles of ROS as both agents of damage and as cellular signals are widely discussed (for reviews see e.g. [16‒18]). In the following, we will give an overview of the reactions of each ROS and describe methods that have been used to detect the ROS, focusing on photosynthetic organisms. The biological roles of the ROS in photosynthetic organisms will be discussed only to the extent needed for the understanding of the detection methods.
ACCEPTED MANUSCRIPT 7
1.2. General principles of detection and monitoring of ROS
T
In pure solvents or in the gas state, concentrations of various ROS can be directly measured by absorbance, fluorescence or EPR spectroscopy. However, the most common methods for monitoring the relative amounts of ROS require the use of chemicals that preferentially react with a particular ROS, yielding a detectable product.
ED
MA
NU
SC
RI P
Most ROS do not equilibrate throughout a complex biological material within their lifetime but either react near to the site of origin or remain confined to the producing compartment. Therefore the cellular localization of the detector substance is often of importance for understanding the results. Unfortunately, experimentally verified information about the localization of the various detector substances is seldom available, especially in plant material. In the present review, we list the available information. The logarithm of 1-octanol-water partition coefficient (LogP) is widely used to estimate hydrophobicity. As a rule, the lower the LogP value of a compound is, the more hydrophilic it is. All experimentally determined and some estimated LogP values in this review have been obtained from online chemical structure database, Chemspider [19]; for compounds not found in Chempsider, LogP values were estimated using an online tool ALOGPS 2.1 [20]. LogP values for different ROS detection compounds are presented in Tables 3‒7. Caution should be taken in interpreting LogP values as indicators of cellular localization or intake of a molecule, because cellular environment cannot be dichotomously divided into two simple immiscible phases. Furthermore, the plant and cyanobacterial cell wall may lower the cell permeability of lipophilic molecules.
AC CE
PT
Hydrophilic chemicals can be fed to intact plant samples in the transpiration stream but lipophilic substances may require vacuum infiltration or scratching the plant surface. Side effects of these aggressive methods should be separately evaluated. Biotechnology offers the possibility to make a mutant organism that synthesizes a specific genetically encoded ROS probe in the desired tissue or cellular compartment [21]. The success of biotechnology is based on the availability of specific cellular ROS probes and so far only probes for H2O2, based on the specific sensitivity of transcription factors (OxyR from Escherichia coli and Orp1 from Saccharomyces cerevisiae) to H2O2, have been tested for the detection of endogenous ROS (see [16,22] for reviews). Promising results have been obtained also in the development of genetically encoded 1O2 probes (see Chapter 2.5.6). Because certain ROS function as cellular signals, specific changes in gene expression can often be correlated with specific ROS. However, conclusions about the causal agent causing a change in gene expression must be drawn with great care, unless the mechanism of regulation is well understood. Enzymes present in all cells can sometimes be used to modify the sensor substances, as in the design of fluorescein diacetate dyes from which cellular esterases remove the acetate moiety, producing a fluorescent, charged dye that cannot diffuse back to the extracellular space [23]. In the case of aminophenyl and hydroxyphenyl fluorescein dyes, the strength of the ROS signal may also depend on the action of a peroxidase [24]. Detection of ROS with chemical reactions is often difficult because several ROS may react with the detector chemical. Unspecific detectors can be used in the rare cases that the system is known to produce only one type of ROS and that only one type of ROS is of interest. However, measurement of the total amount of ROS in the sample cannot be done with an unspecific detector because the reaction rates of the detector with different ROS vary. If the concentration ranges of various ROS
ACCEPTED MANUSCRIPT 8
species in the sample have been well characterized in advance, it may also be possible to use several unspecific detector substances and utilize knowledge about their reaction rates with different ROS.
RI P
T
In addition to the specificity and localization in the biological sample, ROS detector substances may have problems of biological incompatibility, stability in the light and stability of the reaction product. Some ROS detector substances are known to produce ROS. These topics will be discussed when information is available.
SC
2. Singlet oxygen, 1O2 2.1. Definition and properties of 1O2
A ground-state O2 molecule ( g O2) is a triplet form (also designated as 3O2) and can therefore 3
MA
NU
accept two electrons from a species containing paired electrons only if a spin conversion occurs. Thus, the reactivity of O2 at room temperature is moderate because the molecule has two electrons with parallel spins. The singlet forms of O2 containing only paired electrons are much more reactive than ground-state O2. Molecular oxygen has two singlet forms. The two electrons with antiparallel spins may reside
either on two different orbitals ( g O2) or both on one orbital ( g O2). g O2 decays rapidly to 1
ED
1
1
form either ground-state O2 or g , and therefore the designation 1O2 will be used for g O2, 1
1
PT
unless otherwise indicated.
1
AC CE
O2 is a long-lived excited state, but its decay is greatly speeded up by collisions with almost any other molecule [25]. The radiative decay of 1O2 to ground-state O2 yields luminescence with maximum at 1268 nm in the gas phase [26]; in water and deuterium oxide (D2O), the peak is at 1274 nm [27]. Although 1O2 is a singlet state, it has an EPR spectrum [28]. The spectrum has been detected only in the gas phase [29]. Rapid non-radiative decay of 1O2 occurs via (i) conversion of excitation energy of 1O2 to vibrational and rotational energy of both the quencher (e.g. H2O) and O2; (ii) a charge-transfer mechanism between the quencher (e.g. phenol or diazabicyclo[2.2.2]octane, DABCO) and 1O2; or (iii) an electronic energy transfer mechanism in which the reaction between 1O2 and a singlet ground state of the quencher (e.g. a carotenoid) produces a triplet state of both the quencher and oxygen (i.e. ground-state oxygen) [25].
2.2. Formation and action of 1O2 in photosynthetic tissues In photosynthetic tissues, 1O2 is mainly formed by photosensitization (for a review see [30]). Encounter of O2 with either a singlet or a triplet excited state of a pigment molecule (sensitizer) may lead to generation of a singlet form of oxygen. A reaction between a singlet excited state of a sensitizer and O2 yields 1O2 and a triplet state of the sensitizer, whereas a reaction between O2 and a triplet state of the sensitizer produces 1O2 and a singlet ground state of the sensitizer [25]. The reactions between O2 and singlet and triplet excited states of sensitizers are also affected by the energy gap between the singlet and triplet states of the sensitizer [31]. In practice, reactions with singlet excited states are of low significance because singlet excited states are short-lived. Chlorophylls, bacteriochlorophylls, hemes and other tetrapyrroles, especially protoporphyrin IX (an intermediate of synthesis of heme and chlorophyll), as well as iron sulfur centers are known
ACCEPTED MANUSCRIPT 9
NU
SC
RI P
T
photosensitizers [4]. However, chlorophyll (Chl) a and b, and bacteriochlorophylls in non-oxygenic photosynthesis, are the most important natural photosensitizers of 1O2 formation in photosynthetic organisms. According to the available data, production of 1O2 in plants is almost always sensitized by the triplet state of the primary donor (3P680*) of Photosystem II (PSII), formed by charge recombination reactions (for a review see [3]). Thus, neither the triplet state of Chl a (3Chl a) produced by intersystem crossing in the antennae [32] nor the triplet state of the primary donor of Photosystem I (PSI) (3P700*) seem to sensitize 1O2 formation [33,34]. Weak triplet formation in the antennae due to photochemical and non-photochemical quenching of excitation energy, low oxygen concentration in the antenna, and quenching of the triplet states by carotenoids may contribute to the apparently negligible sensitization by many Chl triplets in plants. Experiments with the histidine method (see 2.5.5) show that at photosynthetic photon flux density (PPFD) of 2300 µmol m-2s-1 the 1 O2 quencher histidine lowers oxygen production of the cyanobacterium Synechocystis sp. PCC 6803 by 13 % [35]. This result may actually suggest massive production of 1O2 within the membranes, as most 1O2 is expected to react at the site of production [1].
MA
Other sources of 1O2 than PSII in plants are related to plant defense and e.g. heat stress (for reviews, see [3,36,37]). Non-photochemical formation of 1O2 with the Russell mechanism, via recombination of two peroxy radicals, has also been shown to occur in plant tissues [38,39] but the importance of this pathway has not been studied extensively.
PT
ED
Detoxification of 1O2 in plant cells occurs by carotenoids and tocopherols in the thylakoid membranes (for reviews, see [5,14]) and in the cytosol by soluble scavengers including vitamins C and B6 and glutathione (GSH) [40,41]. 1
AC CE
O2 has been speculated to be the most important ROS formed in the light reactions of photosynthesis, as the triplet state of the primary donor of PSII is ubiquitously formed by charge recombination reactions. Work on the flu mutant of Arabidopsis thaliana that produces 1O2 when grown in a light-dark rhythm has revealed that the reason for cell death is not cellular damage mediated by 1O2 but 1O2-induced signaling events [42‒44]. However, the fact that programmed cell death depends on 1O2 signaling, does not rule out the possibility that 1O2 causes cellular damage.
2.3. Reactions of 1O2
2.3.1. Physical deactivation of 1O2 by radiative and non-radiative mechanisms Only radiative deactivation (reaction 1) is available for an isolated and unperturbed 1O2 molecule, and the radiative lifetime O2(1g) without any collisions is extremely long, 72 min [25]. 1
O2 3O2+ h
(1)
Enhancement of deactivation of 1O2 by collisions of 1O2 with other molecules is one reason for variation of the lifetime of 1O2 in different media. Deactivation of 1O2 by collisions with other molecules having sufficiently high triplet-state energy can occur by conversion of excitation energy of 1O2 to vibrational energy of the other molecule. This non-radiative deactivation process, known as electronic-vibrational energy transfer, limits the lifetime of 1O2 in many solvents. The lifetime of 1O2 in H2O is 3.1 s, about 20 times less than in D2O (68 s) [25,45,46]. Effects of solvent on the lifetime of 1O2 can often provide evidence for participation of 1O2 in chemical reactions.
ACCEPTED MANUSCRIPT 10
T
In addition to the electronic-vibrational non-radiative deactivation, 1O2 can be deactivated via an electron exchange mechanism (reaction 2). This spin-allowed electronic energy transfer to a molecule with a low triplet state energy is a very efficient mechanism of 1O2 deactivation. For example, the rate of deactivation of 1O2 with carotenoids is limited by diffusion. The rate constant for β-carotene and lutein is about 1010 M-1 s-1 [47], O2 + A 3(O2 A) 3O2 + 3A,
RI P
1
(2)
SC
where A is an acceptor molecule, for example a carotenoid. Reaction (2), also known as physical quenching of 1O2, is the main mechanism of quenching of 1O2 by carotenoids. Bimolecular rate constants of reactions between carotenoids and 1O2 can be found in [48].
MA
NU
In addition to their ability to quench 1O2, carotenoids protect the photosynthetic machinery also by quenching both the triplet and singlet excited states of Chls, thereby lowering the probability of 1O2 formation. In plant photosynthesis, the xanthophyll violaxanthin is de-epoxidated in the light to anteraxanthin and zeaxanthin which participate in so called non-photochemical quenching of excitation energy, a mechanism converting the energy of the singlet excited state of Chl a to heat. For review of the photoprotective roles of carotenoids in photosynthesis, see [49].
ED
2.3.2. Chemical reactions of 1O2 1 O2 is an electrophilic agent and reacts preferably with electron rich organic molecules by addition to double bonds. This chemical deactivation of 1O2 proceeds via well-known mechanisms:
AC CE
PT
ene reaction associated with formation of a hydroperoxide (3);
(3)
cycloaddition associated with dioxetane formation (4); (4)
cycloaddition reaction with aromatic compounds and formation of endoperoxides via the Diels-Alder mechanism (5); (5)
formation of charge-transfer intermediates. 1O2 can react with compounds containing heteroatom such as N or S to form charge-transfer intermediates, reactions (6) and (7) respectively. Chargetransfer intermediates are particularly important for the interaction of 1O2 with phenols and tocopherols [50]. O2 + NR3 [1O2--- +NR3]
1
(6)
ACCEPTED MANUSCRIPT 11
O2 + SR2 [1O2--- +SR2]
1
(7)
T
The charge-transfer complex can be disintegrated either by an electron exchange mechanism (physical quenching of 1O2) or by oxidation of a heteroatom, reaction (8).
RI P
(8)
(9)
MA
NU
SC
In addition to oxidation of a heteroatom, the reaction with 1O2 can lead to carbon-heteroatom bond cleavage and formation of a corresponding aldehyde [51]. 1O2 can efficiently oxidize amines to imines with formation of HO2•, reaction (9) [52].
1
AC CE
PT
ED
The interaction of 1O2 with the heteroatom of 2,2,6,6-tetramethly-4-piperidone (TEMP) produces 2,2,6,6-tetramethyl-4-piperidone-1-oxyl (TEMPO, TAN or TEMPONE), a stable radical (reaction 10) [53,54]. The reaction of 1O2 with TEMP (k=4 x 107 M-1 s-1 in phosphate buffer at pH 8, [53]) can be used to detect 1O2 with EPR or mass spectrometry (see 2.5.2).
(10)
O2 can react with many biologically important compounds like unsaturated fatty acids of membrane lipids [55]. The reaction of 1O2 with a cis-double bond of an unsaturated fatty acid can form both conjugated and non-conjugated diene hydroperoxides (reaction 11). The rate constants of reactions of 1O2 with unsaturated fatty acids are slightly higher than 104 M-1 s-1 [41].
(11)
ACCEPTED MANUSCRIPT 12
RI P
T
Methionine, tryptophan, tyrosine, cysteine and histidine react with 1O2 to form semistable products with a rate constant around 107 M-1 s-1 [56]. Histidine has been used for detection of 1O2 [35]. 1O2 also reacts with GSH and other thiol containing compounds with rate constants around 106 M-1 s-1. Reaction of 1O2 with GSH is associated with formation of glutathione sulphinate (GSO2H), glutathione sulphoxide (GSSOG), glutathione sulphonate (GSO3H) and glutathione disulphide (GSSG) [57]. Biologically active forms of vitamin B6 react with 1O2 with rate constants around 1 x 108 M-1s-1 [58]. Oxidation of ascorbic acid and plastohydroquinone by 1O2 can proceed as 2-electron reduction of 1O2 to H2O2 [59,60].
2.4. Lifetime and diffusion distance of 1O2
MA
NU
SC
The lifetime of 1O2 is of interest because a short lifetime would predict that 1O2 reacts at the site of its production whereas a long lifetime would allow 1O2 to act as a signal molecule. The diffusion distance can be approximated from the lifetime (see e.g. [25,61]). The lifetime of 1O2 in cellular environments has been several times measured in D2O based buffers but these measurements do not give a correct picture of what happens in vivo. Current sensor technology has, however, made it possible to measure the kinetics of the 1270 nm luminescence peak after a short laser pulse from some biological materials in H2O based buffers.
AC CE
PT
ED
The formation and decay of the 1O2 peak, obtained by exciting a sensitizer with a short laser pulse, remain unaltered if the values of the rate constant of the decay of the triplet state of the sensitizer (kT) and the rate constant of the decay of 1O2 (kΔ) are exchanged [62] (see Appendix in Supplementary information for the modeling of the kinetics of the 1270 nm peak). For this reason, measurements of 1O2 luminescence kinetics always need knowledge about the value of kT in the same experimental system. The long lifetime of 1O2 in D2O alleviates the problem because sensitizers with a short triplet lifetime (1/kT) can be used. If the sensitizer is an intrinsic constituent of the sample, like Chl in plant photosynthesis, the requirement of the knowledge of kT places a particular weight for the correct identification of the sensitizer. Further complications arise from the heterogenity of the biological sample. Theoretical considerations based on general cellular concentrations of histidine, tryptophan and methionine residues in proteins [63] or on the concentrations of various biomolecules like carotenoids and ascorbate that react with 1O2 [64,65] suggested that the lifetime is 0.05‒1 µs in cells [63,64] and 0.2 µs in chloroplasts [65] (Table 2). Measurements of 1270 nm luminescence kinetics from animal cells have, however, yielded longer but extremely variable lifetimes, ranging from 0.4 to 10 µs in cell suspensions or in individual cells (Table 2). Ragas et al. [66] estimate that the lifetime of 1 O2 inside Escherichia coli cells is only 7 ns. The large variation may partially depend on differences between the samples, as the concentrations of substances reacting with 1O2 may vary between cell types. Also the cellular localization of the sensitizer has been shown to affect the measured lifetime [67]. An apparent source of differences is that in some studies, whole tissues have been used whereas in other studies the current laser technology has been applied to measure 1O2 lifetime from single cells (Table 2). One possible reason for long lifetimes obtained in biological materials is the presence of membranes and lipid bodies in cells. The lifetime of 1O2 is 20‒25 µs in micelles [68] and 12.2 µs in liposomes [69]. Interaction between 1O2 with a quencher or a scavenger molecule requires physical contact of the molecules, and therefore the concentrations of quenchers and scavengers in the lipophilic environment would determine how much they shorten the lifetime of 1O2. Hackbarth et al. [70] see two different lifetime components in a mammalian cell system and suggest that one
ACCEPTED MANUSCRIPT 13
T
corresponds to the membrane and the other to the aqueous phase. However, most experimental data can be explained with only one lifetime component, which suggests that 1O2 equilibrates between different types of compartments within its decay time in the heteronegeous biological environment. Strong laser pulses have also been shown to prolong the measured lifetime of 1O2, probably via depletion of cellular 1O2 quenchers [70].
MA
NU
SC
RI P
So far all 1O2 lifetime measurements from plant and cyanobacteria samples have been done from isolated PSII preparations, either in D2O based buffers [71] or more recently also in water based buffer solutions [65,72,73]. In the case of the simplest PSII reaction centre preparations that lack the QA electron acceptor, the reaction centre triplet is obviously also 1O2 sensitizer. However, the triplet yield of the recombination of the primary radical pair is very low if the electron can proceed from pheophytin to a further acceptor [71], and therefore identification of the sensitizer triplet is more difficult when functional PSII preparations are used. The lifetimes per se mainly reflect the properties of the buffer (D2O or H2O). Correlation between kinetics of decay of the triplet form of the primary donor of PSII and 1270 nm luminescence rise at different temperatures and oxygen concentrations showed that recombination of the primary radical pair produces the sensitizer triplet in PSII particles [71]. Tomo et al. [73] show that much more 1O2 is produced by PSII preparations from a cyanobacterial mutant that contains divinyl Chl a instead of the monovinyl form occurring in the wild type.
AC CE
PT
ED
With regard to the potential signaling role, a triplicate lifetime, after which 5 % of 1O2 still exists, could be a practical limit for the reach of the signal. Assuming, in the absence of measurements, that the lifetime of 1O2 is 3 µs in aqueous stroma of chloroplasts and further assuming that 1O2 can pass through the envelope of a typical chloroplast (5‒7 x 2.5 µm; membrane thickness ~4 nm [74]), then approximately 20 % of the volume of the chloroplast could produce 1O2 of which a significant fraction could diffuse to the cytosol. However, 1O2 produced externally to mammalian cells failed to cause DNA strand breaks while 1O2 produced inside the cells did cause breaks [75,76], indicating that in animal cells the amount of 1O2 diffusing through the plasma membrane is small. Diffusion of 1O2 through the plasma membrane or cell wall has not been measured in plants or cyanobacteria. The diffusion distance of 1O2 may also determine the effectiveness of quenchers and scavengers. In this respect, the finding that added β-carotene protects mammalian cells against 1O2-induced cell death but does not shorten the lifetime of 1O2 [77] is highly interesting. The high viscosity of the cellular environment can partly explain the lack of an effect, as a three-fold increase in viscosity from C6H14 to CCl4 is associated with a 30 % decrease of the quenching constant of 1O2 by β-carotene [48]. The short diffusion distance of 1O2 also implies that cell-impermeable detector substances can only measure extracellularly produced 1O2.
ACCEPTED MANUSCRIPT
T
14
Sensitizer
Method of estimation
Buffer solution
Independent Lifetime measurement
0.05 to 1 µs [64]
0.4‒0.18
Effective signaling distance in viscose cellular environment, µm 0.013‒0.060
< 1 µs [63]
<0.180
<0.06
0.2 µs [65]
0.08
0.03
No
10 µs [79]
0.5
0.2
No
6 µs [79]
0.4
0.15
D2O
Yes
45 µs [80]
1.2
0.4
D2O
Yes
17 µs [67]
0.7
0.25
US
System
CR
IP
Table 2. Measurements of 1O2 lifetime with 1270 nm luminescence kinetics in biological systems. An estimated effective signaling distance, at which 5 % of 1 O2 survives, is given for water (D=1.8 x 10-5 cm2 s-1 at 20 °C; [78]) and for viscose cellular environment (D=2 x 10-6 cm2s-1; [61]).
Nucleus of a single neuron HeLa cells (single cell)
CE P
Photofrin 9-acetoxy-2,7,12,17tetrakis(beta-methoxyethyl)porphycene 5,10,15,20tetrakis(N-methyl-4-pyridyl)21H,23H-porphine Chlorin
AC
Cells
Chloroplast stroma Mammalian cells Mammalian cells
Theoretical, based on known concentrations of quenchers/ scavengers Theoretical (His, Trp, Met residues) Theoretical (ascorbate) 1270 nm H2O luminescence 1270 nm H2O luminescence
TE D
Various cellular environments
MA N
T
Effective signaling distance in water, µm
1270 nm luminescence
ACCEPTED MANUSCRIPT 15
Yes
30‒40 µs [67]
1‒1.1
0.3‒0.4
3.2 µs [81]
0.3
0.1
0.4 µs outside of membrane [70] 0.6 µs [82]
0.1
0.04
0.1
0.04
0.03-0.18 µs [82]
0.03‒0.08
0.01‒0.025
3.7 µs; estimated 0.007 µs inside the cells [66] 44‒59 µs (airsaturated medium, 20 °C [71] 3.2 µs [65]
0.3 (based on the extracellular lifetime)
0.005 (based on the estimated cellular lifetime)
IP
Extrapolation D2O/H2O Yes from mixtures measurements in D2O/H2O mixtures 1270 nm H2O Yes luminescence
CR
Neuron (single cell)
D2O
T
5,10,15,20tetrakis(N-methyl-4-pyridyl)21H,23H-porphine
US
HeLa cells (single cell)
Pheophorbide a
Leukemia cells
Aluminum tetrasulphonated phthalocyanine
1270 nm luminescence
H2O
Rat tissues in vivo
Aluminum tetrasulphonated phthalocyanine
1270 nm luminescence
H2O
Escherichia coli cells
New Methylene Blue and Zinc tetramethyltetrapyridino[3,4b:39,49-g:30,40-l:3-,4q]porphyrazinium salt
1270 nm luminescence
H2O
PSII reaction centre preparations
PSII reaction centre triplet
1270 nm luminescence
D2O
Yes
PSII reaction centre preparations
PSII reaction centre triplet (assumed)
1270 nm luminescence
H2O
No (assignment supported by effect of
AC
CE P
TE D
MA N
Jurkat cells
No (fitting and literature values used) No (fitting and literature values used) No (fitting and literature values used)
ACCEPTED MANUSCRIPT 16
H2O, D2O
Functional PSII particles from wild type and a mutant with divinyl Chl a
PSII reaction centre triplet (assumed)
1270 nm luminescence
H2O
oxygen) Yes
MA N
US
No
TE D CE P AC
3.4‒4.3 µs in H2O; 17‒18.7 µs in D2O [72] 3.94‒4.80 µs [73]
T
1270 nm luminescence
IP
Chls (assumed)
CR
PSII particles
ACCEPTED MANUSCRIPT 17
2.5. Detection of 1O2
SC
RI P
T
2.5.1. Luminescence at 1270 nm Several recent reviews discuss the detection of 1O2 from photosyhthetic organisms [3,83‒85]. 1O2 can be measured by recording the weak 1270 nm luminescence emitted when 1O2 returns to the ground state (reaction 1; Fig. 1; for reviews see [25,84,85]). The detection of 1270 nm luminescence from biological materials is limited by the material thickness, as absorption of water [86] and biomolecules would attenuate the signal to less than one tenth per 1 cm. Therefore, thin samples are preferred. D2O is far more transparent than H2O in the near-infrared range [87].
MA
NU
Due to the low phosphorescence yield and low concentration of 1O2 in biological systems, measurements of 1270 nm luminescence in photosynthetic material have been limited to generation of 1O2 with strong laser pulses, usually in isolated samples, including PSII reaction centres of pea or PSII particles from cyanobacteria [72,73,88]. Significantly more sensitive detectors than are available now will be needed for the measurement of 1270 nm luminescence from biological material under constant illumination [65]. An additional difficulty is to distinguish 1O2 phosphorescence from the tail of Chl a fluorescence, but the separation can be done by spectral comparison or by comparing the decay times [72].
PT
ED
In addition to the 1270 nm "monomol" luminescence, 1O2 also emits luminescence at 634 and 703 nm. This "dimol luminescence" is emitted after a collisional formation of a (1O2)2 dimol in the gas phase [89]. Emissions originating from transitions between several excited states of the oxygen molecule are important in atmospheric science [90] but too weak for studies from biological material.
AC CE
2.5.2. EPR detectable probes TEMP and TEMPD. The two most popular chemicals used to measure 1O2 from photosynthetic material have been 2,2,6,6-tetramethylpiperidine (TEMP) and the more water-soluble 2,2,6,6tetramethyl-4-piperidinone (TEMPD); both commercially available (see Table 3 for review of chemicals used for 1O2 detection). Reaction between 1O2 and TEMP (reaction 10) or TEMPD produces a stable oxyl radical, and the reaction has been shown to be specific to 1O2 [53,91]. In the case of TEMP, the radical is called TEMPO. TEMPD is available both as a pure compound and as a hydrochloride (TEMPD-HCl); only the hydrochloride form can be used for photosynthetic applications because the other form strongly inhibits PSII [92,93]. Due to the low solubility to water, TEMP cannot be used at temperatures below 20 °C. Both TEMP (experimentally determined LogP = 2.15 [19] and TEMPD (LogP = 0.43, as estimated with ALOGPS2.1 [20]) are thought to penetrate into cyanobacterial cells [94,95]. The amount of the TEMPO radical has traditionally been measured with EPR spectroscopy [96,97]. EPR is not suitable for automated measurements of large numbers of samples, and a mass spectrometry method was recently developed for the quantification of TEMPO [98]. TEMP was found not to inhibit the PSII activity of isolated thylakoid membranes [96]. However, storage of TEMP enriches unknown impurities. In the commercially available TEMP, and also in TEMPD, these impurities are present at such a quantity that they cause strong inhibition of PSII, as demonstrated with fluorescence, thermoluminescence and oxygen evolution measurements [92].
ACCEPTED MANUSCRIPT 18
T
Vacuum distillation removes the impurities to such extent that TEMP can be used with isolated plant thylakoids without any adverse effects [93], but even distilled TEMP causes an almost immediate inhibition of PSII in live cyanobacteria (Synechocystis sp. PCC 6803, [5]) and in live diatoms Phaeodactylum tricornutum (Fig. 2).
SC
RI P
With distilled TEMP, 1O2 can be detected from isolated plant thylakoid membranes (e.g. [93,96,98]). TEMP has been used also with purple bacteria [99] and TEMPD has been used to detect 1O2 from isolated thylakoids of the green alga Chlamydomonas [100], but in vivo effects of the probes on the photosystems of the organism were not reported. In intact plant material, the TEMPO radical can be reduced to an EPR silent hydroxylamine, and therefore TEMPO is usually extracted with an organic solvent if EPR is used as the detection method [101].
ED
MA
NU
2.5.3. Dyes Imidazole/RNO. Imidazole-containing compounds, including imidazole itself, are rapidly oxygenated by 1O2 [102,103], and 1O2 can also be detected by measuring bleaching of p-nitroso-dimethylaniline (RNO) by imidazole [104‒106]. The method has been used to measure the action spectrum of 1O2 production in isolated spinach thylakoids [107]. The RNO method should be used with caution in plant material because imidazole has been shown to inhibit PSII oxygen evolution and carbonic anhydrase in a pH-dependent manner [108]. Both imidazole and RNO are commercially available.
AC CE
PT
2.5.4. Fluorescent probes SOSG. Singlet Oxygen Sensor Green®, a proprietary fluorescent probe, is highly selective for 1O2. According to the manufacturer's data, SOSG does not react with H2O2 or O2•, and our unpublished experiments have confirmed the high selectivity. Reaction with 1O2 increases the fluorescence yield of SOSG at 520‒540 nm. According to the manufacturer's data, SOSG is cell-impermeable, but a study with tobacco leaves showed that a pinhole administration of SOSG to tobacco leaves leads to preferential association of the dye with the nuclei of epidermal cells [109]. Gollmer et al. [110] showed that SOSG is also able to penetrate into human cells; therefore SOSG can be used in confocal microscopy [111]. A disadvantage of SOSG is that when illuminated with wavelengths below 600 nm, the fluorescence yield of SOSG increases similarly as during exposure to 1O2 (Fig. 3; [92,109]). Furthermore, the reaction product of SOSG and 1O2 sensitizes 1O2 production [112]. Surprisingly few researchers have taken the photosensitivity of SOSG into account (for correct use of SOSG, see [92,109,113]). SOSG caused 15 % inhibition of the photochemical yield of PSII in tobacco leaves [109]. It is unclear if SOSG can be used with live cyanobacteria, as SOSG seems to penetrate only to a small fraction of the incubated cells and fluorescence from both SOSG and Chl a is only rarely detected from the same cell [114]. Illumination of isolated pumpkin thylakoids in the presence of SOSG with >650 nm light caused a very small signal which was not sensitive to D2O, anaerobicity or azide, suggesting that SOSG cannot be used to measure 1O2 from isolated thylakoids [92]. However, SOSG has been used for detecting 1O2 from isolated plant thylakoids and from intact leaves [109,113,115]. Reasons for the contrasting results about the sensitivity of SOSG in plant material are presently unknown. DanePy. Dansyl-2,2,5,5-tetramethyl-2,5-dihydro-1H-pyrrole (DanePy), another fluorescent probe, is sensitive to 1O2 with minor reactivity with O2•, H2O2 and lipid radicals [116,117]. The probe is not available commercially. DanePy has been used with bean, pea, tobacco, Arabidopsis thaliana and
ACCEPTED MANUSCRIPT 19
SC
RI P
T
Phillyrea latifolia leaves in vivo [109,118‒121] and with the photosynthetic proteobacterium Rhodobacter sphaeroides [122]. DanePy can be used for imaging [118‒120]. DanePy is fluorescent (500‒600 nm, maximum at 545 nm) but forms a non-fluorescent adduct with 1O2, and therefore the presence of 1O2 is inferred from the decrease in fluorescence yield from untreated control to a treated sample. DanePy is harmless for the PSII yield if administered to a leaf either through a pinhole or in the transpiration stream and localizes intracellularly, mainly to chloroplasts [119,120]. DanePy is not light-sensitive [109]. An oxalate derivative of DanePy was used with live Chlamydomonas cells [100] but this dye does not penetrate to leaf cells [109].
NU
DMAX and DPAX. The probes 9-[2-(3-carboxy-9,10- dimethyl)anthryl]-6-hydroxy-3H-xanthen-3-one (DMAX) and 9-[2-(3-carboxy-9,10- dimethyl)anthryl]-6-hydroxy-3H-xanthen-3-one (DPAX) can be used to detect 1O2 with fluorescence [123,124]. DMAX and DPAX were reported not to react with H2O2, O2• or NO• [123,124], and DPAX has been used with intact chloroplasts [125]. DMAX and DPAX are not commercially available.
AC CE
PT
ED
MA
2.5.5. Oxygen consumption in a reaction between 1O2 and histidine Histidine is known as a powerful, biocompatible antioxidant because its imidazole ring reacts rapidly with 1O2 (k=5 x 107 M-1 s-1; [126]; for a review, see [127]). Histidine also protects against HO•, but this protection may be based on the efficient scavenging of metals by histidine rather than reactions between histidine and HO• [127]. Recently, a simple method in which 1O2 production is measured by monitoring the decrease of oxygen concentration in the presence of histidine [128] has been applied to 1O2 measurements from live cyanobacteria [35,129,130]. Cyanobacteria take up histidine from the medium and therefore histidine can be used for 1O2 measurements in live cyanobacteria [35]. The method was shown to be insensitive to O2• and H2O2 [35]. No reports of the use of the histidine method in plants or algae in vivo have been published, and it is possible that in eukaryotes histidine would not enter the chloroplasts that normally export histidine. The reaction between histidine and 1 O2 occurs in neutral to alkaline solutions [126]. 2.5.6. Genetically encoded probes Specific, genetically encoded probes have not been used for the detection of 1O2. However, the IFP1.4 protein containing a biliverdin (an oxidized form of bilirubin) chromophore [131] has been shown to react with 1O2 [132] and not with H2O2, O2• or HO• [132]. The protein emits far-red fluorescence (maximum at 710 nm) when excited at 684 nm, and might therefore act as a perfect genetically encoded 1O2 sensor. The combination of IFP1.4 and a 1O2 sensitizer protein has been used for measurement of protein-protein interaction [132]. Furthermore, fluorescence of the UnaG protein that contains bilirubin as a chromophore was shown to decrease due to a reaction between bilirubin and 1O2 [133], and thus this protein may also be developed into a genetically encoded 1O2 probe. These probes have not yet been applied for detection of endogenous 1O2 production.
ACCEPTED MANUSCRIPT 20
Specificity
Stability
Toxicity
Quantification of TEMPO with EPR [96] or mass spectrometry [98] in vitro
Specific [53,91]
Stable but instant measurements recommended [98]
Not toxic to isolated plant thylakoids [93]; Inhibits PSII in live cyanobacteria [5] and algae (Fig. 2)
Notes
Cell permeable (Synechocystis) [95]
Not toxic to isolated plant thylakoids [93]
-
1.24 (estimated)
Cell permeable (Synechococcus) [94]
Vacuum distil before use [93,92]; Extraction of TEMPO to organic solvent avoids formation of EPR silent hydroxylamine [101] Use TEMPD-HCl [93]
Does not react Stable with H2O2 or O2• [35]
Not toxic
-
-3.32 [19]
-
Only in neutral or alkalic pH [126]
No reaction with H2O2 or O2• (manufacturer's
Little or no effect on photosynthesis
Only red light (>660 nm) can be used
-
Nuclei of epidermal cells; does not localize
The reaction product sensitizes 1O2
US
Light sensitivity
MA N
TE D
Specific, similar Stable to TEMP
CE P
TEMPD (2,2,6,6- Quantification tetramethyl-4of the reaction piperidinone) product with EPR in vitro [100] Histidine Consumption of oxygen by histidine, in vitro; in vivo from cyanobacteria [35] Singlet Oxygen Fluorescence (Ex Sensor Green, 505, Em 525), in SOSG vitro and in vivo
Localization
-
Hydrophobicity (LogP) 2.15 [19]
CR
Method
AC
Compound (IUPAC name) TEMP (2,2,6,6tetramethylpipe ridine)
IP
T
Table 3. Chemicals used to detect 1O2. The stability data indicate the stability of the signal after reaction with 1O2 in plant tissue. LogP values have been obtained from Chemspider [19]; estimated LogP values were determined using a LogP prediction tool ALOGPS 2.1 [20]. - = no data.
Stable
ACCEPTED MANUSCRIPT 21
in plants [109]
[92,109,113]
Minor reaction with H2O2, O2• and lipid radicals [116,117]
No effect on PSII efficiency [109]
Imaging
-
-
-
Decrease in absorbance at 440 nm
-
-
DPAX (9-[2-(3carboxy-9,10dimethyl)anthryl ]6-hydroxy-3Hxanthen-3-one) IFP1.4 (genetically encoded probe with biliverdin chromophore) UnaG (genetically encoded probe with bilirubin
Fluorescence (Ex 492, Em 516‒517), intact chloroplasts [124]
Does not react with NO, H2O2 or O2• [123]
Decrease in fluorescence (Ex 684, Em 710)
Does not react with NO, H2O2 or O2• [132]
Decrease in absorbance (~475) and fluorescence
Aproprotein reacts with O2• at high concentration
to chloroplasts [109] Cellular; mainly chloroplasts [109,119,120]
production [112] -
Extracellular in plants [109]; cytoplasmic in Chlamydomonas [100] -
Less sensitive than DanePy [109]
-
-
IP
Fluorescence (Ex 364, Em 545 nm) is quenched by 1O2, imaging
Not sensitive
-
-
-
Imidazole inhibits PSII [108]
-
-
-
-0.08 imidazole, [19]/ 1.45 p-nitrosodimethylaniline (estimated) Soluble to water at 10 µM [123]
-
-
-
-
-
Not yet tested for endogenous 1 O2
-
-
-
-
-
Not yet tested for endogenous 1 O2
CE P
TE D
-
AC
-
US
CR
DanePy (dansyl2,2,5,5tetramethyl-2,5dihydro-1Hpyrrole) DanePy-oxalate (dansyl-2,2,5,5tetramethyl-2,5dihydro-1Hpyrrole oxalate) RNO (imidazole and p-nitrosodimethylaniline)
MA N
T
data)
ACCEPTED MANUSCRIPT 22
CE P
TE D
MA N
US
CR
IP
T
(~525)
AC
chromophore)
ACCEPTED MANUSCRIPT 23
RI P
T
2.5.7. Further gene expression methods Ability to sense 1O2 may be important both in avoiding damage and in signaling [134], and photosynthetic organisms have several genes that react specifically to 1O2 (e.g. [135]), as well as to other ROS (see e.g. [136]). The mRNA levels of 1O2 inducible genes have been used to mark 1O2 levels in the cells [137]. Also reporter systems, where proteins induced by 1O2 have been fused to marker proteins, have been developed, for example in Chlamydomonas reinhardtii and Arabidopsis thaliana [138,139], respectively.
MA
NU
SC
2.5.8. Indirect methods of 1O2 detection Damage caused by 1O2. If 1O2 is not quenched or scavenged, it can damage proteins, pigments and lipids [1,140]. The amount of 1O2 can be estimated by the extent of the damage, as reactions between cell components and 1O2 sometimes have specific reaction products. Products of lipid peroxidation, as well as oxidation products of carotenoids, plastoquinone and tocopherol, have been suggested to be used for detection of 1O2 [55,137,141,142], respectively. However, it is not usually very easy to make strong conclusions based on these data, and the ability of the photosynthetic organisms to quench 1O2 can also vary. The capacity of leaf extracts to quench 1O2 can be measured for example according to [143].
AC CE
PT
ED
Modifications of the amount or lifetime of 1O2. As the lifetime of 1O2 is lengthened 5‒20 fold in D2O compared to water [46,72], measurements can be done in D2O to facilitate 1O2 detection (e.g. [72]). However, the use of D2O in place of H2O has consequences in biological materials. Proteins are often more stable and more rigid in D2O than in water [144]. It has been shown that with photosynthetic organisms D2O reduces growth and oxygen evolution as well as slows down PSII electron transfer [145] and carbon fixation [146]. Chl is a natural sensitizer of 1O2 in photosynthetic systems but the amount of 1O2 can be increased by using with artificial 1O2 (photo)sensitizers [147]. However, the popular 1O2 generators Rose Bengal and methylene blue have been shown to affect PSII electron transfer [148]. Different 1O2 generators localize in different cellular compartments. For example, Neutral Red and Rose Bengal penetrate to the chloroplast, in contrast to methylene blue [148]. If the sensitizer penetrates weakly to the target tissue, then the amount of 1O2 produced may not be physically relevant. The extent of protection against cellular damage, obtained by a treatment with an 1O2 scavenger like azide, histidine or DABCO, has also been used to assess the importance of 1O2 in the damage mechanism (e.g. [149]). Due to the high reactivity of 1O2, scavengers that localize to the site of 1O2 production, often within a membrane, are preferred. Effects of natural quenchers. Carotenoids and tocopherols that function as the principal quenchers and scavengers of 1O2 can often be used to draw conclusions about the participation of 1O2. For example, participation of 1O2 in the oxidative inhibition of the repair of PSII after photoinhibition was supported by the finding that both in plants [150] and in cyanobacteria [151] the ability to synthesize -tocopherol protects the PSII repair mechanism but does not affect the rate of light-induced damage. The same tocopherol data suggested that 1O2 is not involved in the damaging reaction of photoinhibition, but on the other hand, a cyanobacterial mutant with a high carotenoid content is less sensitive to photoinhibition than the wild type [130].
ACCEPTED MANUSCRIPT 24
3. Hydrogen peroxide, H2O2 3.1. Definition and properties of H2O2
NU
SC
RI P
T
2.5.9. Summary of 1O2 detection To conclude, 1O2 measurements are a tricky business. Direct measurements of 1O2 are still not possible from physiological conditions. When using chemical traps (see Table 3) it is important to check for the possible adverse effects on the organism or biological function, in addition to checking the sensitivity and reliability of the probe. The place where the sensor compound localizes can also affect the results. More 1O2 probes exist than are used in plant research and also new probes are being synthesized but the usability of the new probes in photosynthetic systems should be tested before extensive use [117,152,153]. For example, the 1O2 probe MVP (trans-1-(2’methoxyvinyl)pyrene) has long been used e.g. in medicine, but it could not detect 1O2 from leaves, maybe because its chemiluminescence (465 nm) overlaps with Chl a fluorescence [109].
MA
H2O2 is the result of reduction of oxygen by two electrons. In biological systems, H2O2 is mostly present in the neutral form because of its high pKa value (pKa1 is 11.8). H2O2 is a moderate oxidant (E0 of the pair H2O2+2H++2e-/2H2O is 1.776 V; [154]) that can oxidize compounds like thiols.
ED
3.2. Formation and action of H2O2 in photosynthetic tissues
AC CE
PT
In biological systems, including photosynthetic tissues, H2O2 is mainly formed via dismutation of O2• (reaction 29), and because superoxide dismutases (SODs) are present in most cellular compartments, the formation of H2O2 depends on the formation of O2•. Furthermore, the direct 2electron reaction between 1O2 and plastohydroquinone was recently shown to produce a stoichiometric amount of H2O2 (Fig. 4; [60]). The production of H2O2 by the interaction of 1O2 and plastohydroquinone may have importance in chloroplast signaling because H2O2 produced within the membrane would avoid the stromal scavenging mechanisms [60]. In the soluble phase, 1O2 produces H2O2 in a reaction with ascorbate [59]. H2O2 can be enzymatically scavenged by catalases (CATs), peroxidases and peroxiredoxins. CATs have a relatively low affinity for the substrate, reflected by a high Michaelis constant Km for H2O2; usually around 50 mM in plant catalases [155,156], although a lower value of 17.6 mM was also reported [157]. Peroxidases have lower Km values (0.57 mM for ASC, 0.11 mM for H2O2; [158]). Ascorbate and GSH function as electron donors for different peroxidases. Peroxiredoxins form a third group of H2O2 scavenging enzymes. In peroxiredoxins, the oxidizable thiol group is a cysteine residue of the protein itself, and therefore peroxiredoxins require an activating reduction step between the catalytic rounds. H2O2 is known to be associated with reactions causing cellular damage. For example, the action of the herbicide paraquat, known also as methyl viologen, is based on formation of O2•, which inevitably leads to formation of H2O2, too. However, it is usually difficult to determine if the actual damaging agent is H2O2, O2• or HO• formed in the Fenton reaction (reaction 15). Cyanobacteria are highly sensitive to H2O2 [159] and this sensitivity can actually be used to destroy potentially poisonous cyanobacteria from lakes [160].
ACCEPTED MANUSCRIPT 25
In plants, H2O2 functions, together with O2•, as a signal molecule in stomatal opening, programmed cell death and various stress responses (for review, see [161]).
3.3. Reactions of H2O2
NU
SC
RI P
T
H2O2 is particularly reactive against thiol groups [162], especially those of cysteine residues of proteins and other compounds [16], and the reactivity of H2O2 towards thiols explains the inhibitory effect of H2O2 on the thiol-group-containing enzymes of the Calvin-Benson cycle. H2O2 inactivates fructose-1,6-bisphosphatase, glyceraldehyde 3-phosphate dehydrogenase and xylulose-5-phosphate kinase [163] and the Krebs cycle enzyme aconitase [164] and reacts with several other plant proteins [165,166]. The Escherichia coli transcription factor OxyR recognizes H2O2 via a reaction with a cysteine residue [16,167]. The reaction of H2O2 with thiol groups is also a H2O2 elimination mechanism, accomplished by peroxiredoxins in plant cells [13]. Reduction of H2O2 by thiol groups is a two-step reaction passing through the mechanism described by reactions 12.1 and 12.2.
MA
H2O2 + RSH RSOH + H2O RSOH + RSH RSSR + H2O
(12.1) (12.2)
ED
The rate constants of reaction 12.1 range from 102 to 107 M-1 s-1. The thiolate anion (RS), formed in high pH, reacts more rapidly with H2O2 than the respective thiol (RSH) [162,168], and the reactivities of thiols depend on both pH and the pKa value of the thiol [162].
AC CE
PT
A well-known reaction of H2O2 is its reduction via the Haber-Weiss mechanism (sum of reactions 13 and 14) and Fenton mechanism (reaction 15). Both mechanisms generate the very powerful oxidant HO•. The rate constant of reaction 13 in aqueous medium ranges from 0.13 to 0.23 M-1 s-1 [169,170] and the rate constant of reduction of H2O2 by HO2• (reaction 14) is 0.5 M-1 s-1. Thus, reactions (13) and (14) are relatively slow in biological material where concentrations of O2• and H2O2 are low. O2• + H2O2 O2 + OH + HO•
(13)
HO2• + H2O2 O2 + H2O + HO•
(14)
M+(n-1)+ H2O2 M+n + OH + HO•
(15)
The reduction of H2O2 by transition metals via the Fenton mechanism (reaction 15) is much more efficient than the Haber-Weiss reaction (reactions 13 and 14). The rate constant of Fenton reactions depend on the metal M (Fe and Cu are the biologically most important ones) or metal complex and pH, and range from 50 to 104 M-1 s-1 [171‒174]. The reduction of H2O2 via the Haber-Weiss and Fenton mechanisms (reactions 13‒15) causes subsequent reactions where interactions between O2•, HO2•, HO• and the reduced and oxidized metal ion can occur. In particular, O2• is able to reduce Fe3+ to Fe2+ and Cu2+ to Cu+, thereby promoting continuing production of HO• via reaction (15). The reduction of H2O2 by peroxidases is one of the main routes of H2O2 scavenging; organic compounds including ascorbate, glutathione and guaiacol can be used as electron donors. In the chloroplast, H2O2 scavenging is mainly catalyzed with an ascorbate-specific peroxidase (APX, (EC
ACCEPTED MANUSCRIPT 26
1.11.1.11)). The catalytic reduction of H2O2 by peroxidases proceeds via the peroxidase ping-pong mechanism (reactions 16‒18) [175]. PX-Fe(III)-P + H2O2 PX-Fe(IV)=O-P+ + H2O
T
(16)
RI P
PX-Fe(IV)=O-P+ + AH PX-Fe(IV)=O-P + A• +H+ PX-Fe(IV)=O-P+ AH PX-Fe(III)-P + A• + OH,
(18)
SC
where P is porphyrin.
(17)
MA
NU
The rate constant of the reaction of H2O2 with APX is around 107 M-1 s-1 and the Km for H2O2 is 80 M [176,177]. Thus, APX is an efficient enzyme with high affinity to the substrate and a very fast rate of the reaction. CAT-depended scavenging of H2O2 also occurs through the peroxidase-like ping-pong mechanism where one H2O2 molecule is used as an electron donor (reactions 19.1 and 19.2). The decomposition of H2O2 by CAT, unlike the reaction catalyzed by peroxidase, is associated with O2 production. (19.1)
H2O2 + Fe(IV)=O-CAT+→ H2O + Fe(III)-CAT + O2
(19.2)
ED
H2O2 + Fe(III)-CAT H2O + Fe(IV)=O-CAT+
PT
The peroxidase-catalyzed reaction of H2O2 with a specific electron donor like luminol can be used for detection of H2O2. In this case, electron donor is oxidized to an optically detectable product (see 3.5.5). In addition to the above, H2O2 irreversibly inactivates two of the three types of SOD (Cu,ZnSOD, Fe-SOD), and APX in the absence of ascorbate [178].
AC CE
3.4. Lifetime and diffusion distance of H2O2 As a small, neutral molecule, H2O2 is expected to readily diffuse through membranes. Judging from experiments with the genetically encoded probe HyPer, the lifetime of H2O2 is 1‒3 min in mammalian cells [179] and around 10 s in Arabidopsis thaliana guard cells [180]. Both lifetimes are so long that 5 % of H2O2 would be present at 0.2‒1 mm from the site of generation in a viscous biological material.
3.5. Detection of H2O2 3.5.1. Principles of H2O2 detection H2O2 can be detected with several different compounds functioning as hydrogen donors for the reduction of H2O2. Some detector substances require a catalyst, usually a peroxidase or a metal ion, and if the study object does not contain endogenous peroxidases, horseradish peroxidase can be used together with the detector substance. 3.5.2. Precipitate-forming compounds DAB. The most widely used H2O2 detection chemical in plant research has been 3,3'diaminobenzidine (DAB) (see Table 4 for a list of chemicals used for H2O2 detection). DAB reacts with H2O2, forming a brown precipitate that can be made visible in plant material by removing Chl (Fig. 5; [181‒183]). DAB penetrates well to plant cells and it is not sensitive to light [184], the precipitate is stable [183], and DAB localizes partially to the chloroplast [185]. Unfortunately, DAB strongly lowers
ACCEPTED MANUSCRIPT 27
the quantum yield of PSII electron transfer in the concentration range required for imaging [184]. The inhibition of photosynthesis by DAB should be taken into account in the interpretation of results obtained with DAB from photosynthetic material.
RI P
T
Cerium. H2O2 reacts with cerium chloride, forming electron-dense precipitates of cerium perhydroxides [186]. This has been used to localize H2O2 production e.g. in wheat roots [187] and Zinnia elegans stems [188] from electron transmission microscope images.
ED
MA
NU
SC
3.5.3. Dyes MBTH/DMAB. A peroxidase-catalyzed oxidation by H2O2 couples 3-methyl-2-benzothiazoline hydrazone (MBTH) and 3-(dimethylamino) benzoic acid (DMAB) to a deep purple indamine compound that can be assayed spectrophotometrically at 590 nm [189]. Veljovic-Jovanovic et al. [190] showed that a plant leaf extract causes, in addition to a reaction caused by the H2O2 present in the extract, slow but significant unspecific increase in 590 nm absorption. The unspecific increase could be avoided with polyvinylpyrrolidone treatment. Furthermore, ascorbate inhibits the MBTHDMAB signal [190]. The effects of ubiquitous plant metabolites on the MBTH-DMAB signal call for special care to assure that the signal actually reflects the H2O2 content of the study material. The MBTH/DMAB method has been used to measure the increase in extracellular peroxidase activity in sunflower roots during defense signaling [191].
PT
4-aminoantipyrine/phenol. H2O2 can be measured with 4-aminoantipyrine and phenol as H2O2 oxidatively couples with them forming colourful quinone-imine (absorbance at 505 nm) [192]. Peroxidase is needed for the reaction. The method has been used to measure H2O2 from pea chloroplasts [193] and from root extracts of tomato [194].
AC CE
Starch/KI. Iodide ions (I-) are oxidized by H2O2 after which iodine (I2) is complexed with (exogenous) starch, and H2O2 can be histochemically localized by the blue color of the starch/I-complexes. Starch and iodide do not penetrate to cells [188,195] but the starch/KI-method has been used to detect H2O2 production in cut stem surfaces [195] and in peeled or partially digested leaf segments [196] or to measure extracellular H2O2 production by whole roots [197]. Age and wounding of the plant affect the results obtained with the starch method [188]. Iodide can be oxidized also by organic peroxides (ROOH) [198] and other cellular electron acceptors besides H2O2 [195]. DHBS. In the presence of peroxidases 3,5-dichloro-2-hydroxybenzenesulfonic acid (DHBS) is oxidized by H2O2 to its quinone form which then reacts with 4-aminoantipyrine (4-AAP) to produce a colorful substance that can be measured spectrophotochemically at 510 nm. The DHBS method has been used to measure extracellular H2O2 from a tobacco cell suspension [199]. 3.5.4. Fluorescent probes Scopoletin. Decrease of the fluorescence yield of scopoletin (7-hydroxy-6-methoxy-2H-1benzopyran-2-one) is a traditional in vivo method of H2O2 detection in biological systems [200]. In plant leaves, scopoletin mainly localizes to the cell wall but also penetrates to the cells [184]. The usability of scopoletin in plants in vivo is limited by the need to use ultraviolet (UV) wavelengths (346 nm) for excitation, as plant leaves strongly absorb UV light [201]. The 435 nm emission of scopoletin is also strongly absorbed by Chl a. Scopoletin has been used for measurement of H2O2 in root cells
ACCEPTED MANUSCRIPT 28
[202] and in the incubation media of callus cells [203] but also from leaves of Pseudowintera colorata [204].
RI P
T
Eu3Tc. An europium-tetracycline complex (Eu3Tc) [205] reacts with H2O2 to form a highly fluorescent adduct (excitation at 360‒440 nm; emission at 616 nm). The fluorescence yield of Eu3Tc-H2O2 adduct is strongly dependent on pH [205] and therefore the complex should be used only in a buffer solutions at pH 6.6‒7.2. Eu3Tc is not photosensitive but lowers PSII yield [184]. Eu3Tc has not been used with other photosynthetic materials.
ED
MA
NU
SC
Amplex Red and Amplex Ultrared. Amplex Red is 10-acetyl-3,7-dihydroxyphenoxazine which, after oxidation by H2O2, fluoresces at ~580 nm when excited at 570 nm; Amplex Ultrared is a derivative of the original reagent. The commercial kits also contain horseradish peroxidase as a catalyst. Both compounds have been tested for H2O2 imaging from leaves, and Amplex Red penetrates to cells and even to chloroplasts whereas Amplex Ultrared was found to stay mostly in the apoplast [184]. Both compounds have a strong negative effect on PSII yield [184], which limits their use in plant material. Furthermore, both compounds are sensitive to light [184]. Cyanide, salicylhydroxamic acid, and propyl gallate have been shown to inactivate the Amplex Red method in vitro [206]. Antal et al. [207] used Amplex Ultrared to measure H2O2 formation by isolated thylakoid membranes after treatments with short flashes in the dark.
AC CE
PT
3.5.5. Chemiluminescent probe Luminol. 3-amino-phthal-hydrazide, known with the common name luminol, is oxidized by H2O2 in a peroxidase-catalyzed reaction to produce an excited state of 3-aminophthalate; the decay of the excited state yields blue luminescence peaking at 440 nm. Reaction with H2O2 requires a peroxidase or a metal ion as a catalyst, and a mixture of luminol and H2O2 has been used from the 1930's in crime scene investigation to detect the iron of haemoglobin present in blood stains. If the catalyst is provided, luminol can be used to measure H2O2, and a detection limit of 0.42 nM was obtained for cobalt-catalyzed measurement of H2O2 from seawater [208]. The luminescence of luminol is quenched by many cellular substances like ascorbic acid and phenols, which limits the usability of luminol for H2O2 measurements from intact biological materials [190]. Luminol chemiluminescence is also observed in the presence of peroxynitrite and O2• ([209]; see section 4.5.5). Plastoquinone was recently found to enhance the luminescence [60]. However, the Co(II) catalyzed reaction offers the very high sensitivity of 0.5 nM for plant extracts diluted to such extent that the quenchers and enhancers do not disturb the reaction [210]. The luminol/Co(II) method revealed an increase in H2O2 concentration in Arabidopsis thaliana leaf tissue after a shift to high light [211] and a hypoxiainduced increase in the H2O2 level of grapevine buds [212]. 3.5.6. Microelectrodes Microelectrodes. Electrochemical H2O2 microsensors (for O2• electrodes, see 4.5.6) have been developed (for a review see [213]). Ren et al. [214] developed a carbon fiber Pt-ultramicroelectrode coated with single-walled carbon nanotubes and hemoglobin (Hb/SWCNTs/CFUME), and were able to monitor the oxidative burst (H2O2) in Aloe leaves in vivo with a detection limit of 4 μM. Neither 0.1 mM ascorbic acid nor NaCl affected the electrode. A microelectrode with poly-ophenylenediamine and Pt microparticles (POPD–Pt-MP–Pt) has been used to measure H2O2 induced by Cd2+ treatment from leaves of oilseed rape in vivo [215]. Another Pt-electrode has been used in
ACCEPTED MANUSCRIPT 29
vitro with Arabidopsis thaliana cell suspensions [216]. The small size of current electrodes (diameter less than 10 µm [214]) facilitates measurements from intact plants.
ED
MA
NU
SC
RI P
T
3.5.7. Genetically encoded probes With genetically encoded sensors there would be no need to get foreign molecules inside the cells to detect H2O2 (or O2•, see 4.5.7) and they also offer ways for locating the ROS in the cell, and the use of (fluorescence) microscopy (for reviews see [217,218]). The ability of H2O2 to oxidize the thiol group of OxyR [167]; reactions 12.1 and 12.2] was used in designing the H2O2 sensor HyPer, in which circularly permuted yellow fluorescent protein (cpYFP) was fused with the regulatory domain of OxyR [21]. HyPer functions as a genetically encoded H2O2 sensor. Hyper is claimed to be specific to H2O2 [21], and it has been used also in Arabidopsis thaliana to measure H2O2 produced in the peroxisome lumen and in the cytoplasm [180]. HyPer has excitation maxima at 420 and 500 nm and emits at 516 nm, and the oxidation of HyPer by H2O2 enhances the 500 nm peak and lowers the 420 nm peak. However, the sensor (like many fluorescent proteins) shows strong pH-dependence [179], and therefore the measurement should be coupled with measurement of pH. Many fluorescent proteins themselves produce ROS when illuminated, which may lead to photobleaching [219,220], but Hyper appears to show little photobleaching [221]. Another issue with genetically encoded probes is that they act as antioxidants, and their continuous expression may affect the physiology of the organism [222]. Recently published genetically encoded H2O2 probes (OxyFRET and PerFRET) respectively apply the Orp1 and Yap1 proteins of the H2O2 sensing system of yeast [223]; these probes have not yet been tested with plant material.
AC CE
PT
3.5.8. Unspecific probes H2O2 has been measured from plant material [224,225] and from Chlorella cells [226] with fluorescein derivatives that react primarily with HO• [227]. Because these dyes can be used to detect more than one ROS, they will be described in Chapter 8. 3.5.9. Indirect methods of H2O2 detection Increase in the activities of H2O2 scavenging enzymes, including APX and CAT [180] or increase in the expression levels of genes coding for these enzymes may suggest an increase in the amount of H2O2 in the tissue. A considerable number of genes are affected by H2O2 in plants [136,228] and in cyanobacteria [229]. Many of these genes are also regulated by light [230], and plant chloroplasts produce H2O2 in the light. Purified H2O2 scavengers offer possibilities for simple measurements of H2O2 in vitro. For example, H2O2 production by a cell suspension has been measured via the reaction of a scavenger [231]. 3.5.10. Summary of H2O2 detection Some methods of H2O2 determination have been used long, and lots of knowledge about different aspects of their use has been accumulated whereas the caveats in newer, promising methods may still remain to be found. Unfortunately, most detection methods that have been studied with regard to their effects on photosynthesis seem to have adverse effects. This is particularly important for DAB, as this dye localizes to the chloroplasts in DAB-infiltrated leaves. Convincing determination of H2O2 requires the use of more than one method, and measurements of H2O2 from extracts are recommended to back up in vivo staining experiments.
ACCEPTED MANUSCRIPT 30
Specificity
Stability
Brown precipitate, Imaging, semiquantitative
Specific
Stable
Cerium
Electron transmission microscopy Absorbance at 590 nm
-
-
Ascorbate inhibits the signal [190]
-
Absorbance at 505 nm
-
Histochemical staining
Oxidized by ROOHs and cellular electron acceptors [195,198] -
Toxicity
Hydrophobicity (LogP) 0.24 (estimated)
Localization
Notes
Cellular; penetrates to chloroplasts [184] -
Too low sensitivity at acceptable concentration [184]
-
-
-
TE D
Strongly lowers PSII yield [184]
-
-
1.41 MBTH (estimated)/ 1.35 DMAB (estimated)
In vitro or from ground tissue
-
-
-
-
-
-
Peroxidase is needed for the reaction
-
-
-
1.04 iodide [19]
Cellimpermeable [195]
Wounding affects the results
-
-
-
-
-
Peroxidase is needed for the reaction
CE P
DHBS/4-AAP (3,5- Absorbance at 510 dichloro-2nm hydroxybenzenes
AC
MBTH (3-methyl2benzothiazoline hydrazone) and DMAB (3(dimethylamino) benzoic acid) 4aminoantipyrine/ phenol Starch/KI
Light sensitivity Not sensitive [184]
US
Method
MA N
Compound (IUPAC name) DAB (3,3'diaminobenzidin e)
CR
IP
T
Table 4. Methods of detection of H2O2. Light sensitivity data are for visible light, unless otherwise indicated. The stability data indicate the stability of the signal after reaction with H2O2 in plant tissue. LogP values have been obtained from Chemspider [19]; estimated LogP values were determined using a LogP prediction tool ALOGPS 2.1 [20]. - = no data.
-
ACCEPTED MANUSCRIPT
-
-
Signal affected by several metabolites and pH [205] -
Decrease Not sensitive to <50 % [184] in 4‒5 min in leaves t1/2 = 6‒8 min
Amplex Ultrared (modified from Amplex Red)
Fluorescence (Ex 567/ Em 580), Imaging
-
Luminol (3amino-phthalhydrazide)
Chemiluminescence at 430 nm
Hb/SWCNTs/CFU ME (Microelectrode
Amperometry
Interference by ascorbate, phenolics [190] and plastoquino ne [60] -
1.62 (estimated)
Cell wall, cytoplasm [184]
-
Lowers PSII yield [184]
-
Mostly intercellular [184]
-
Strongly lowers PSII yield [184]
1.93 (estimated)
Addition of peroxidases may be needed.
Strongly lowers PSII yield [184]
-
At least partly cellular, accumulates in chloroplasts [184] Limited penetration to cells [184]
-
1 % increase in signal in 30 min at PPFD 1200 µmol m2 -1 s [184] 30 % increase in signal in 30 min at PPFD 1200 µmol m2 -1 s [184] -
-
-0.41 (estimated)
In vitro or from ground tissue
Use of Co as catalyst facilitates measurements from plant tissues [210]
-
-
-
-
-
Not affected by ascorbic acid or NaCl [214]
-
US
MA N
TE D
CE P AC
Amplex Red (10acetyl-3,7dihydroxyphenox azine)
-
CR
Fluorescence (Ex 346/ Em 435) quenching, Quantification, imaging Reversible formation of fluorescent (Ex 400/ Em 615) Eu3Tc-H2O2 complex, Imaging Fluorescence (Ex 570/ Em 583), Imaging
Europiumtetracycline complex (Eu3Tc)
IP
ulfonic acid/ 4aminoantipyrine) Scopoletin
T
31
Linear decrease to <50 % in 15 min
Addition of peroxidases may be needed
ACCEPTED MANUSCRIPT
IP CR
Only little photobleachin g [221]
-
-
Can be targeted to desired tissue in vivo
Fluorescence depends strongly on pH [179].
-
-
Can be targeted to desired tissue in vivo
Moderate pH sensitivity [223]
MA N
US
-
-
-
TE D
Fluorescence ratio (Ex 435 nm, Em 535 nm/480 nm)
Specific to H2O2 [21] but may be affected by other oxidants [179] Specific to H2O2 in vivo [223] but not tested with isolated protein
CE P
OxyFRET and PerFRET (genetically encoded probes based on Orp1 and Yap1, respectively).
Fluorescence ratio (Ex 420/500 nm, Em 516 nm)
AC
based on haemoglobin and carbon nanotubes) HyPer (genetically encoded probe based on OxyR)
T
32
ACCEPTED MANUSCRIPT 33
4. Superoxide anion radical, O2•, and hydroperoxyl radical, HO2• 4.1. Definitions and properties of O2• and HO2•
MA
NU
SC
RI P
T
Reactions and biological activities of O2• have interested the scientific community for tens of years, and several reviews are available (see [232‒235]). O2• is 1-electron reduced form of O2. The midpoint redox potential (Em) of the pair O2/O2• at pH 7.0 is -160 mV vs Normal Hydrogen Electrode (NHE) in aqueous solution [236]. The Em of the pair O2/O2• becomes more positive if O2• is protonated to form HO2•, and therefore HO2• is a stronger oxidant than O2• [237]. However, O2• is a weak deprotonation agent in aqueous solutions. O2• can form O2•(H2O)n complexes (with n from 1 to 3 [233] and the free energy of hydration was estimated to be around 355 kJ mol-1 [238]. The pKa value for HO2• is 4.8 and therefore only 0.25 % of O2• is protonated at pH 7. In aprotic medium O2• forms a weaker solvation complex. The pKa value of HO2• in DMF was estimated to be 12, and therefore O2• can act as a strong deprotonation agent in an aprotic medium. Weak solvation of O2• in aprotic media leads to a shift of the redox potential of the pair O2/O2• in comparison with an aqueous solution. The redox potential of the pair O2/O2• in DMF was estimated to range between -550 and -600 mV vs NHE [233].
PT
ED
In addition to its role as an oxidant in reactions that produce H2O2 (see reactions 32.1, 34 and 35 below), O2• can act as a reductant, which leads to conversion of O2• to O2 (reactions 23, 24, 32.2 and 33); see [237] for a compilation of reactions of O2• and HO2•. Enzymatically catalyzed dismutation of O2• proceeds via oxidation of one O2• ion and reduction of another one. The dual role of O2• is unique among ROS, as other ROS function as oxidants in biological environment.
AC CE
4.2. Formation and action of O2• and HO2• in photosynthetic tissues 1-electron reduction of oxygen by PSI is the main photosynthetic source of O2•, but small amounts of O2• are also produced by PSII [2,239] and by reduced forms of plastoquinone [60]. O2• is invariably scavenged by SODs (EC 1.15.1.1), enzymes catalyzing the dismutation reaction (reaction 29). Almost all aerobic organisms contain SOD(s), and the enzymes fall in the phylogenetic groups of Mn and Fe-SODs, Cu,Zn-SODs and Ni-SOD [240]. The Mn and Fe-SODs form one clade. Chloroplasts contain mainly Cu,Zn-SODs [240]. All SODs speed up O2• dismutation, but the reaction proceeds even in the absence of an enzyme. In addition to SODs, redox active metals like manganese ions catalyze the dismutation.
4.3. Reactions of O2• and HO2• According to its thermodynamic properties, O2• expresses dual reactivity; it acts as a weak oxidant and as a strong reducing agent in electron transfer reactions in aprotic media in the absence of proton donors. The redox potential of O2•/HO2 (data are available only for aprotic medium) is very low, -1.75 V in DMSO [232]. On the other hand, O2• is able to oxidize organic molecules via deprotonation-oxidation mechanism in the presence of protons or proton donors. Thus, the behavior of O2• in redox reactions very strongly depends on its protonation. Furthermore, O2• reacts preferably with neutral or radical partners and only slowly with negatively charged ones.
ACCEPTED MANUSCRIPT 34
RI P
T
In addition to electron-proton transfer reactions, O2• can take part in nucleophilic substitution or addition if the substrate has a low electron affinity. O2• participates in nucleophilic reactions with alkyl halides, sulfonates, phosphates, esters, acyl halides, acyl anhydrides to form peroxyl (ROO•), alkoxyl (RO•) radicals and epoxides as intermediates [232,233,241,242]. Both the nucleophilicity and the basicity of O2• depend on medium. O2• acts as a powerful nucleophilic agent in aprotic medium due to rapid protonation and dismutation in aqueous solutions.
SC
Reactions of O2• with organic and inorganic molecules fall in five basic categories: protonation, electron transfer reaction, nucleophilic substitution and addition, attraction of hydrogen, and addition to metal or metal complex.
MA
NU
Protonation. Protonation of O2• is a highly important reaction because this reaction is the main source of HO2• and is involved in a complicated deprotonation-oxidation process as well. O2• can be protonated via a reaction with proton or a proton donor such as ascorbic acid, reactions (20) and (21) respectively. O2• + H+ HO2• O2• + AscH2 HO2• + AscH
(20) (21)
(22)
PT
O2• + -TocH HO2• + -Toc
ED
O2• can be also protonated by -tocopherol (-Toc) in aprotic medium, reaction (22):
AC CE
Electron transfer. O2• can reduce many organic and inorganic molecules such as quinones and cytochromes, reactions (23) and (24), respectively, via a 1-electron transfer mechanism. O2• + Q O2 + Q•
(23)
O2• + Cyt c(Fe3+) O2 + Cyt c(Fe2+)
(24)
Nucleophilic addition and substitution. Nucleophilic reactions of O2• with organic molecules are associated with formation of corresponding peroxy radicals which are more oxidizing than O2• itself. For example, O2• reacts with carbon tetrachloride (CCl4) to form the trichloromethyl peroxy radical [243], reaction (25). O2• + CCl4 Cl3COO• + Cl
(25)
Nucleophilic reaction of O2• with nitrones is widely used for detection of O2• with EPR. The reaction of O2• with 5-diethoxyphosphoryl-5-methyl-1-pyrroline N-oxide (DEPMPO) yields a stable DEPMPO/O2• adduct (reaction 26) [244].
(26)
ACCEPTED MANUSCRIPT 35
T
Attraction of hydrogen. Attraction of a hydrogen atom from organic molecules by O2• is an unlikely reaction. Most probably O2• dependent attraction of hydrogen is the result of its protonation. HO2• is a typical neutral radical and is able to attract a hydrogen from organic molecules such as linolenic acid [237], reaction (27).
RI P
HO2• + LH H2O2 + L•
(27)
SC
Addition to metal of metal complex. A reaction of O2• with transition metals or metal complexes yields unstable and stable intermediates [233,237]. For example, O2• can react with manganese complexes Mn2+-L to form MnOO+L complexes [245], reaction (28). Mn2+-L + O2• MnOO+-L
(28)
NU
Complex reactions involving O2•. O2• is involved in many complicated reactions. Below we will describe only the most important O2• depended reactions occurring in photosynthetic organisms.
O2• + O2• H2O2 + O2
PT
ED
MA
Dismutation is one of the main reactions determining the lifetime of O2• and its possibilities to take part in other biochemical processes. Dismutation is the reaction between two molecules of O2• and can occur via noncatalytic and catalytic routes. Noncatalytic dismutation of O2• proceeds, in general, in aqueous medium and it can be written as a reaction (29). Dismutation is a 2-step reaction; the first is protonation of O2• (reaction 20), and the second step is a radical-radical reaction between O2• and HO2• or between two molecules of HO2•, reactions (30) and (31) respectively. (29) (30)
HO2• + HO2• H2O2 + O2
(31)
AC CE
O2• + HO2• + H+ H2O2 + O2
Dismutation of O2• is very slow in alkaline pH (0.3 M-1 s-1) because O2• is mainly deprotonated. Protonation of O2• leads to an increase in the reaction rate constant with a maximum (108 M-1 s-1) at pH 4.8, the pKa value of HO2•. In acidic pH less than 4.8, when O2• is mostly protonated, the reaction rate constant is reduced to 8.6 x 105 M-1 s-1 [237]. The catalytic route of O2• dismutation involves the SOD enzyme (EC 1.15.1.1). The general reaction can be written as follows: M(n+1)+-SOD + O2• → Mn+-SOD + O2
(32.1)
Mn+-SOD + O2• + 2H+ → M(n+1)+-SOD + H2O2,
(32.2)
where M is a metal ion present in SOD. The rate constant of O2• dismutation catalyzed by SOD is about 6.4 x 109 M-1 s-1 [246]. O2• and HO2• can react with H2O2 via the Haber-Weiss reaction (reactions 13 and 14). O2• can also mediate the reduction of H2O2 in the Fenton reaction (reaction 15) by reducing a transition metal ion, e.g. Fe3+ (reaction 33) [247].
ACCEPTED MANUSCRIPT 36
O2• + Fe3+ O2 + Fe2+
(33)
RI P
T
In addition to the Haber-Weiss reaction, O2• can directly participate in decomposition of ROOHs. O2• can react with ROOH yielding the ozonide anion radical and aldehydes or ketones as final products [248]. O2• is able to react with benzoyl peroxide in toluene with formation of 1O2 [249]. In cells, decomposition of ROOH by O2• might be observed in hydrophobic regions like membranes.
O2• + AscH2 + H+ → H2O2 + AscH•,
(34)
MA
where AscH• is monodehydroascorbate radical.
NU
SC
In cells, O2• is particularly reactive against iron-sulfur centres, and the ability of the Escherichia coli transcription factor SoxR to sense O2• is based on this reactivity [16]. Ascorbic acid and reduced GSH can participate in scavenging of O2•. The oxidation of ascorbic acid by O2• is a complicated reaction concerted with transfer of a hydrogen atom and a proton [250]. The general reaction of O2• with ascorbic acid (reaction 34) can be represented in the following way:
O2• + GSH + H+ → H2O2 + GS•
ED
O2• can react with reduced GSH to form H2O2 and gluthyl radical (GS•), reaction (35). Addition of O2 to GS• is considered as a source of peroxysulfenyl radical (GSOO•) [251]. (35)
PT
The rate constant of the reaction of O2• with reduced GSH was estimated to be 103 M−1 s−1 at maximum [162,252]. It is much less than the rate constant for the reaction of O2• with ascorbic acid, 3.3 x 105 M−1·s−1 at pH 7.8 [253].
AC CE
The ability of O2• to bind with transition metal complexes determines its toxicity to metalloenzymes. O2• can inactivate the ferredoxin-dependent nitrate reductase, interacting with molybdenum in its active center [254]. CAT passes into an inactive form due to interaction of O2• with its heme [255].
4.4. Lifetime and diffusion distance of O2• High pH and low concentration of O2• increase its lifetime by affecting the dismutation rate (reactions 29‒30; [256]). According to Fridovich [257] the lifetimes of 0.1 mM and 0.1 nM solutions of O2• in water are 0.05 s and 14 h, respectively. In seawater, 87‒1120 pM concentrations of O2• have been measured, with half-life of 10‒100 s [258,259]. In animal cells the half-life of O2• has been calculated to be below 50 ms, resulting in 40 µm diffusion distance [260,261]. To our knowledge, no measurements of the lifetime of O2• in plant cells have been published; in general, the lifetime would be controlled by SOD (reaction 32; [257]). Due to its negative charge, O2• cannot easily pass through membranes [262], and in cellular pH only a minor fraction would be in the protonated form. Plant PSII particles have been shown to produce 1‒2 µmol O2• (mg chl)-1 h-1 in high light [263], and extracellular O2• production by Chlamydomonas in UV-light was measured to be 0.60‒2.40 µmol g-1 min-1 [264].
ACCEPTED MANUSCRIPT 37
4.5. Detection methods of O2•
RI P
T
4.5.1. Spectroscopic methods O2• can be detected directly by spectroscopy (absorption at 230–350 nm; [265]) or with EPR spectroscopy [266], but due to the strong absorption of cell components at UV-range and the low amount of O2•, these methods are rarely useful in biological applications. Changes in the EPR spectrum, induced by UV treatment of attached plant leaves, were attributed to O2• [267], but the origin of these signals is controversial [268].
AC CE
PT
ED
MA
NU
SC
4.5.2. EPR detectable probes DMPO, EMPO, DEPMPO, BMPO, DCP, TMT and Tiron. O2• produced by photosynthetic organisms has often been measured with probes that, after reacting with O2•, can be detected with EPR spectroscopy (for reviews, see [269,270]; see Table 5 for a list of chemicals used for detection of O2•). The most widely used probes are 5,5-dimethyl-1-pyrroline N-oxide (DMPO), 5(diethoxyphosphoryl)-5-methyl-1-pyrroline-N-oxide (DEPMPO; reaction 26), 5-(ethoxycarbonyl)-5methyl-1-pyrroline N-oxide (EMPO) and 5-tert-butoxycarbonyl-5-methyl-1-pyrroline N-oxide (BMPO; Fig. 6), all of which are commercially available. DMPO has been used to detect extracellular O2• production by intact cells of a red alga Chattonella marina [271] and with EMPO, O2• has been detected from isolated PSII membranes of spinach [272]. DEPMPO has been used for O2• detection from isolated tobacco thylakoids [273], from reaction center complexes of spinach [274] and from isolated plasma membranes of maize [275]. BMPO has been used to detect O2• from cell wall extracts of pea [276] and from isolated PSII membranes of spinach [277]. The lipophilic cyclic hydroxylamine probe 1-hydroxy-4-isobutyramido-2,2,6,6-tetramethylpiperidinium (TMT) and the more hydrophilic probe 1-hydroxy-2,2,5,5-tetramethylpyrrolidine-3,4-dicarboxylic acid (DCP) have been used to detect O2• from pea thylakoids [278]. In comparison to the traditional EPR probes, also commercially available TMT and DCP are said to be more sensitive, and can be also used to detect the location of O2• production [278,279], although TMT might be re-reduced by photosynthetic electron transport chain [278]. The spin trap 4,5-dihydroxy-1,3-benzene-disulfonic acid disodium salt (Tiron) has been used to detect O2• from wheat roots [280] and from lichen tissue [281]. OXANOH and PTM-TC. The spin traps 2-ethyl-1-hydroxy-2,5,5-trimethyl-3-oxazolidine (OXANOH) and perchlorotriphenylmethyl radical-tricarboxylic acid (PTM-TC) have been used to detect O2• from spinach thylakoids [282,283], respectively, and from plant roots [284]. O2• induces oxidation of OXANOH to OXANO, but it is also auto-oxidized in the presence of metal-ions [282]. In contrast to most of the spin traps, the EPR signal of PTM-TC decreases in the presence of O2•. OXANOH and PTM-TC are not commercially available. Specificity. DMPO, EMPO, BMPO and DEPMPO are not specific to O2• but react also with other radicals (especially HO• and carbon centered radicals). Fortunately, reaction products of the spin trap with different ROS often have different EPR spectra [273,285‒287] (Fig. 6; see also 5.5.2). Thus, the limited specificity of these EPR probes is not only a drawback but also an advantage, as several ROS can be detected with one experiment. For example, Bogdanovic Pristov et al. [276] used DEPMPO to detect the carbon dioxide radical (CO2•), O2• and HO•. In contrast, the ability of cyclic hydroxylamines (e.g. TMT and DCP) with other than O2• limits the usability of these compounds
ACCEPTED MANUSCRIPT 38
because the different products usually have similar EPR spectra [279]. Interference with transition metal ions can be reduced by using chelating agents [279]. PTM-TC does not react with HO•, ROO•, H2O2, NO•, GSH or with L-ascorbate [288].
ED
MA
NU
SC
RI P
T
Lifetime of spin adducts. A challenging feature of the EPR probes is that the spin adducts usually decay rapidly [289,290], especially the spin adducts of DMPO (t1/2 is about 45 s; [291]). The spin adducts of DEPMPO and EMPO are more stable, for EMPO t1/2 is about 8 min and for DEPMPO t1/2 is about 14 min [286,287,289], and of these two the EPR spectrum of the EMPO adduct is less complicated [287]. The half-lifetime of the BMPO adduct is 23 min [292]. The spin adduct of PTM-TC is stable [288]. Also Tiron has been reported to be quite stable (t1/2 is about 15 min), but the need for basic (pH 8.5) solution might restrict its use [280]. Efforts have been made to design more stable spin probes (for example [293]). Furthermore, cyclodextrins can be used to stabilize the spin adducts [290]. Cyclodextrins have also been shown not to significantly inhibit photosynthesis when used in appropriate concentrations [290]. With plant material, DMPO is more commonly used as HO• probe (see 5.5.2), however, the DMPO/O2• spin adduct decays to the HO• adduct and also to HO• itself [294]. Furthermore, UV-radiation can cause generation of the DMPO cation radical, which after reaction with water generates the hydroxyl adduct of DMPO [295]. Spin adducts of DEPMPO or EMPO do not decay to the HO• adducts [286,288] or decay only slowly (>20 min with DEPMPO) [269].
AC CE
PT
Test of the effects of these probes on photosynthesis are rare. TMT and DCP do not affect the reduction of P700+ (although in the precence of DCMU, TMT was observed to slightly accelerate the reduction of P700+) [278]. Long incubations of plant roots with Tiron exert a negative effect on root growth [296]. Long (6 h) incubations of animal cells with DMPO, EMPO, BMPO and DEPMPO have been shown to slow down growth and cause death of the cells [297]. Higher concentrations than 90 mM of DMPO may be toxic to algae [271]. In vivo detection of O2• with EPR probes is still rare. Warwar et al. [284] were able to obtain onedimensional images of O2• production in a whole plant root using PTM-TC, and they also concluded that the probe stayed in the apoplastic space. More commonly used EPR probes, DMPO, BMPO and DEPMPO are thought to be membrane-permeable [298]. 4.5.3. Dyes NBT, NBD and XTT. Nitro blue tetrazolium (NBT) is a traditional O2• dye that has also been used with plant material. The reaction of NBT with O2• can be followed spectrophotometrically at 550‒560 nm. O2• has been detected with NBT from legume roots [299], pea leaves [300], diatoms [301] and from extracted leaf material of maize [302]. To quantify the signal, the biological material has to be made transparent by removing Chl [299] or centrifuged away [302], but the precipitated product of the NBT-O2• reaction can be histochemically localized. A method for quantification of O2• from stained leaves has been developed [303]. Both NBT and its reaction product are quite water insoluble. A more water soluble O2• probe sodium, 3´-(1-[phenylamino-carbonyl]-3,4- tetrazolium)-bis(4-methoxy-6-nitro) benzene-sulfonic acid hydrate (XTT), commonly used as a cell viability assay [304], has been introduced [305]. XTT has
ACCEPTED MANUSCRIPT 39
been used to detect O2• from tobacco cell suspensions and from cucumber leaf extracts [306,307], respectively.
SC
RI P
T
NBT can be reduced also by P680 of PSII and CO2- [83,308] and it has also been shown to generate O2•, although this reaction may not be significant [308]. Furthermore, detergents like Triton X-100 increase NBT reduction [309]. A related probe, 4-chloro-7-nitrobenzo-2-oxa-1,3-diazole (NBD), has been reported to be more selective towards O2• [310], but NBD has not been used, to our knowledge, with photosynthetic organisms. Neither is XTT an optimal O2• probe, as it has been shown that Escherichia coli suspensions can reduce XTT also anaerobically (with NADPH:XTT reductases; [311]) and short-chain sugars and phenolics reduce XTT, which makes the use of whole cell material difficult [311,312].
PT
ED
MA
NU
Cytochrome c. Reduction of (exogenous) (ferri)cytochrome c (Cyt c, commercial) has often been used to detect O2•, as the reduction of Cyt c can be monitored spectrophotometrically (around 550 nm; [313]). The Cyt c method has been used in vitro to measure O2• from PSII-enriched membrane fragments from higher plants [263], from apoplasts of wheat roots [314] and from Chlamydomonas sp. ICE-L suspension [264]. In addition to O2•, Cyt c can be reduced by the components of the photosynthetic electron transport chain [315,316] and by ascorbate, GSH and other reductants [317]. In low and moderate light intensities (at PPFD 5–200 µmol m-2 s-1) Cyt c was mainly reduced by the plastoquinone pool and not by O2• [316]. For this reason, comparison of the results obtained with and without added SOD is always required [263]. To enhance the specificity of Cyt c to O2•, it is possible to use acetylated Cyt c [318], for example with Chlamydomonas in [264]. However, reduced Cyt c can also be reoxidized by many cell components, like cytochrome oxidases and peroxidases and by oxidants like peroxynitrite (ONOO-) [317], which calls for care with the use of the method.
AC CE
Epinephrine. Epinephrine (also called adrenaline) has been used to detect O2• as after the reaction with O2•, the absorbance of the reaction product, adrenochrome, can be measured (at about 490 nm; [319]). Epinephrine has been used to measure O2• from chestnut seeds [320] and spinach chloroplasts [313], extracellular O2• production from bryophytes and lichens [281] and roots [191], and mitochondrial O2• production from barley roots [321]. This method is not specific to O2•, as epinephrine has been shown to react also with H2O2 [322], and therefore specificity assays are always needed (see e.g. [191]). In addition, adrenochrome can be metabolized by a NADPHdependent cytochrome P450 reductase [323] and take part in redox cycling with mitochondrial complex I, resulting in fast increase in O2• production [324]. For the latter reason the authors [324] advise to measure only the initial part (60 s) of O2• production after addition of epinephrine. Hydroxylamine. O2• can be detected by following the formation of nitrite from hydroxylamine (NH2OH) with absorbance at 530 nm, but pH should be below 8 to avoid generation of O2• by autooxidation of hydroxylamine [325]. Comparison with measurements in the presence of added SOD may be needed to exclude unspecific reactions. The method has been used to measure O2• from homogenized leaves of maize [326]. Hydroxylamine penetrates well to leaf cells [327]. A drawback of this method is that hydroxylamine is known to reduce the manganese ions of the oxygen-evolving complex of PSII, which leads to loss of PSII electron transfer capacity [327,328].
ACCEPTED MANUSCRIPT 40
ED
MA
NU
SC
RI P
T
4.5.4. Fluorescent probes HE. The commercially available probe hydroethidine (HE; or dihydroethidine DHE) has been widely used to detect O2•, as its reaction product emits yellow to red fluorescence. HE is permeable to both mammalian and plant cells [329,330] and oxidized HE intercalates with DNA, which enhances the fluorescence. HE also has mitochondria-specific analog, Mito-HE or MitoSOX red [331]. Unfortunately, also HE reacts with other substances than O2•, for example with cytochromes, hemeproteins, manganese-porphyrin complex, hypochlorous acid NO and ONOO− [332‒334], but the reaction product of HE and O2• differs from the unspecific reaction products [335]. To enhance the specificity of fluorescence detection, one should use 490 nm light for excitation and detect 560‒570 nm fluorescence. However the spectra of the O2• specific and unspecific reaction products overlap so strongly [335] that reliable detection requires high performance liquid chromatography (HPLC) measurements [335,336]. HE can also be oxidized in UV, and in some conditions also in visible light [337], which prevents detection of light-induced O2• formation. Additional issues weakening the usability of HE are that is has been shown to increase the dismutation rate of O2•, and be toxic for Escherichia coli [332]. Nevertheless, HE has been used to detect O2• from plants, for example from tobacco cells and pea roots with a fluorescence microscope [338] and from spinach leaves with HPLC [339]. For HPLC detection, a simplified method has been developed [340], though, at least to our knowledge, it has not been used with plant material. Martin et al. [341] measured O2• production by Arabidopsis thaliana mitochondria by Mito-HE.
AC CE
PT
4.5.5. Chemiluminescent probes Three commercially available chemiluminescent probes, lucigenin (N,N'-Dimethyl-9,9'-biacridinium dinitrate; luminescence at 470 nm), CLA (2-Methyl-6-phenyl-3,7-dihydroimidazo[1,2-a]pyrazin-3(7H)one; luminescence at 380 nm), a chemical derived from the luciferin of Cypridina, and luminol (see 3.5.5) have been used for detection of O2• from plant material. Lucigenin. Lucigenin has been used with suspension cultures of rose [342]. Lucigenin itself is capable of producing considerable amounts of O2• via redox cycling especially when used in a concentration exceeding 10 µM; unfortunately high concentrations are required for the detection of O2• [343,344]. Furthermore, lucigenin can also be reduced by NAD(P)H reductases [345]. Luminol and L-012. Luminol, better known as H2O2 detector substance (see e.g. [346]), has also been used to measure O2• from potato tuber discs [347], Vicia faba roots [348] and from tobacco leaves [349]. Luminol [350], and its analog 8-amino-5-chloro-7-phenylpyrido[3,4-d]pyridazine1,4(2H,3H)dione (L-012) [351] might also generate O2• [346]. All of these probes, lucigenin, luminol and L-012, also react with H2O2 and peroxynitrite, L-012 being the most insensitive towards H2O2 but having remarkable reactivity with peroxynitrite [209]. CLA. CLA has been used to detect O2• from tobacco cell suspension [352] and from the epidermal layer of Vicia faba leaves [353]. In addition to O2•, CLA has been reported to react with 1O2, and to some extent also with HO• and H2O2 [354]. CLA has even been used as a 1O2 probe (for example [354]). Addition of SOD suppressed the CLA signal completely in plant leaves [353], except for an initial spike.
ACCEPTED MANUSCRIPT 41
RI P
T
MCLA. Several CLA-derived O2• probes are commercially available, including 2-methyl-6-(4methoxyphenyl)-3,7-dihydroimidazo[1,2-a]pyrazin-3(7H)-one (MCLA). It has been reported that MCLA does not react with H2O2 or O3 but reacts with HO• and 1O2 [355]. Godrant et al. [356] measured extracellular production of O2• by cyanobacteria Trichodesmium erythraeum and concluded that at PPFD 50‒100 µmol m-2 s-1 there was no interference from 1O2, nor did addition of 100 nM H2O2 affect the signal, though some signal from blank was observed as well as quite remarkable amount of a SOD non-inhibitable signal. MCLA has also been used to measure the (extracellular) production of O2• by immobilized marine diatoms [357].
MA
NU
SC
4.5.6. Microelectrodes With a microelectrode (0.02 cm2) consisting of Cu,Zn-SOD immobilized on nanostructured ZnO disks Deng et al. [358] measured in vivo O2• production from bean sprouts. Direct electron transfer between SOD and the electrode was achieved, with detection limit of 0.2 µM. No interference from H2O2, O2, uric acid, ascorbic acid or 3,4-dihydroxyphenylacetic acid was detected. O2• has been detected in vitro from tomato cell suspension with a Pt electrode in which SOD was immobilized within carboxymethylcellulose-gelatin [359,360].
AC CE
PT
ED
4.5.7. Genetically encoded probes cpYFP. Recently a new fluorescent O2• probe, Mito-cpYFP, was presented [361]. It is a circularly permuted yellow fluorescent protein that was targeted to mitochondria of human osteosarcoma cells using the cytochrome C oxidase subunit IV targeting sequence, and it was reported to react only with O2•, not with other ROS, nor with Ca2+, ATP, ADP, NAD(P)+ or NAD(P)H [361]. The probe was also used to detect mitochondrial O2• production in Arabidopsis thaliana [362]. However, Schwarzländer et al. [363] reported that in the mitochondria of Arabidopsis thaliana the probe only responded to changes in pH and not to changes in O2• level. Furthermore, purified cpYFP appeared not to respond to O2• [364]. 4.5.8. Indirect methods of O2• measurement Gene expression. The transcriptomic response of Arabidopsis thaliana [136] and the dinoflagellate Pyrocystis lunula [365] to O2• has been published, but this knowledge has not been used for O2• detection. 4.5.9. Summary of O2• detection As almost all the methods used to measure O2• have flaws, reliable results can be best obtained by using at least two parallel methods [276,313,366]. Furthermore, use of SOD or other agents to confirm specificity to O2• is recommended, although the complex chemistry of some probes (NBT, luminol and lucigenin) may make the evaluation more difficult (see [346,367]). In particular, SOD cannot very well quench O2• produced inside the membranes (e.g. [282]), and the probes themselves may also generate O2• that is quenched by SOD. Different localization of different chemicals should be taken into account when comparing results obtained with different probes [278].
ACCEPTED MANUSCRIPT 42
IP
T
Table 5. Chemicals used for the detection of O2•. The stability data indicate the stability of the signal after reaction with O2• in plant tissue. LogP values have been obtained from Chemspider [19]; estimated LogP values were determined using a LogP prediction tool ALOGPS 2.1 [20]. - = no data. Method
Specificity
Stability (t1/2)
DMPO (5,5-dimethyl-1pyrroline N-oxide)
EPR
Also reacts with HO• and other radicals [273,285‒287]
45 s [286,287,289]
DEPMPO (5(diethoxyphosphoryl)-5methyl-1-pyrroline-Noxide) EMPO (5(ethoxycarbonyl)-5methyl-1-pyrroline Noxide) BMPO (5-tertbutoxycarbonyl-5-methyl1-pyrroline N-oxide)
EPR
Also reacts with HO• and other radicals [273,285‒287]
8 min [286,287,289]
EPR
Also racts with HO• and other radicals [273,285‒287]
14 min [286,287,289]
Slower growth of animal cells at 25 mM [297]
EPR
Also reacts with HO• and other radicals [273,285‒287]
23 min [292]
TMT (1-hydroxy-4isobutyramido-2,2,6,6tetramethylpiperidinium)
EPR
Also reacts with HO• and other radicals [273,285‒287]
DCP (probe 1-hydroxy2,2,5,5tetramethylpyrrolidine-
EPR
-
4 h, reduced by photosyntheti c electron transport [278] 4h
Hydrophobici Localization ty (LogP) 0.34 (estimated)
Slower growth of animal cells at 25 mM [297]; No effect on photosynthesis [96] Slower growth of 0.47 animal cells at 25 (estimated) mM [297]
US
MA N
TE D
AC
CE P
Toxicity
CR
Compound (IUPAC name)
Notes Spin adduct decays to HO• spin adduct and to HO• [294], UV-light sensitive [295]
-
Spin adduct may decay slowly to HO• spin adduct [269,287,289]
0.41 (estimated)
-
Spin adduct may decay slowly to HO• spin adduct [287,289]
Death of animal cells after long incubation, slower growth at 25 mM [297] No effect on reduction of P700+ [278]
0.86 (estimated)
-
-
1.5 [279] (experiment al)
-
-
No effect on reduction of P700+ [278]
-1.3 [279] (experiment al)
-
-
ACCEPTED MANUSCRIPT 43
Epinephrine
-
-
EPR
No reaction with HO•, ROO, H2O2, NO, GSH or L-ascorbate [288] Reduced by P680, and CO2- [83,308]
Stable [288]
T
EPR
Long incubation slows root growth [296] -
Absorbance at 550 nm
Absorbance at 490 nm
Sugars and phenolics reduce [311,312]
Reacts with ascorbate, GSH, plastoquinone, other components of photosynthetic ETC, oxidases and peroxidases [315‒317] Reacts with H2O2 [322]
US -
MA N
Absorbance at 550‒560 nm, histochemical staining Absorbance at 470 nm, histochemical staining
CR
IP
15 min [280]
Stable
TE D
XTT (Sodium,3*-(1[phenylamino-carbonyl]3,4- tetrazolium)-bis(4methoxy-6-nitro benzenesulfonic acid hydrate) Cytochrome c (Cyt c)
-
CE P
NBT (Nitro blue tetrazolium)
EPR
AC
3,4-dicarboxylic acid) Tiron (trap 4,5-dihydroxy1,3-benzene-disulfonic acid disodium salt) OXANOH (2-ethyl-1hydroxy-2,5,5-trimethyl3-oxazolidine) PTM-TC (perchlorotriphenylmethy l radical-tricarboxylic acid)
-0.15 (estimated)
-
0.97 (estimated)
Cell permeable
Have been used in vivo in roots, but detection in vitro [296] Interference from metal-ions [282]
Lipophilic, soluble to water at 1 mM [284] 3.49 (estimated)
Apoplast [284]
Has been used in vivo in roots [284]
Cell permeable
-
-
-
1.95 (estimated)
Cell permeable
-
Reaction product can be oxidized by cell components or peroxynitrite [317] Reaction product metabolized
-
0.30 (estimated)
-
Comparison with and without added SOD is needed, enhanced specificity by the use of acetylated Cyt c [263]
-
-1.37 [19]
-
Reaction product may produce O2• [324]
ACCEPTED MANUSCRIPT
Absorbance at 530 nm
-
HE or DHE (Hydroethidine)
Fluorescence (Ex 490, Em 520 nm)
Reacts with cytochromes, hemeproteins, Mnporphyrin complex, hypochlorous acid, NO and ONOO[332‒334] Reacts with H2O2 [209,345]
-1.23 (estimated)
Cell permeable [327]
4.03 (estimated)
Cellpermeable, Mito-HE localizes in mitochondri a
Can be reduced by NAD(P)H reductases [345] -
-
3.37 (estimated)
-
Produces O2• [343,344]
-
-0.41 (estimated)
-
May produce O2• [346]
-
-
1.10 (estimated)
-
M0ay produce O2• [346]
-
-
2.02 (estimated)
-
-
-
-
2.19
-
-
US MA N
-
TE D
CE P
Lucigenin (N,N′ -Dimethyl- Luminescenc 9,9′ -biacridinium e at 470 nm dinitrate)
Hydroxylamine inhibits oxygen evolving complex of PSII [327,328] Reaction product toxic to E. coli [332]
CR
Hydroxylamine
IP
by Cyt P450 reductase [323] -
T
44
Luminescenc e at 430 nm
L-012 (8-amino-5-chloro7-phenylpyrido[3,4d]pyridazine1,4(2H,3H)dione)
Luminescenc e at 460 nm
CLA (2-Methyl-6-phenyl3,7-dihydroimidazo[1,2a]pyrazin-3(7H)-one) MCLA (2-methyl-6-(4-
Luminescenc e at 380 nm
For reaction with H2O2, see 3.5.5; reacts with peroxynitrite [209] Reacts with H2O2 in the presence of peroxidase and with peroxynitrite [209,345] Reacts with 1O2, H2O2 and HO• [354]
Luminescenc
More specific than
AC
Luminol (5-amino-2,3dihydro-1,4phthalazinedione)
SOD may be needed to exclude unspecific reactions, produce O2• at the pH 8.5 or above [325] HPLC distinguishes specific and unspecific reaction products [292]; oxidized in UV and VIS light, enhances dismutation of O2• [332]
ACCEPTED MANUSCRIPT 45
CLA, reacts with 1O2 and HO• [355]
-
89.8% of sensitivity retained for 80 days [359]
-
-
-
-
-
-
-
Responds to pH, not to O2[364]
CR
-
TE D
-
-
US
Mitochondria targeted fluorescent (Ex 488, Em 505 nm) protein
-
MA N
Signal decreases less than 10 % in 6 days [358]
IP
T
(estimated)
Amperometry No interference by H2O2, O2, uric acid, ascorbic acid or 3,4dihydroxyphenylaceti c acid [358] Amperometry No response with 0.5 mM H2O2 [359]
CE P
CMC–G–SOD (Microelectrode with SOD on carboxymethylcellulosegelatin) Mito-cpYFP
e at 465 nm
AC
methoxyphenyl)-3,7dihydroimidazo[1,2a]pyrazin-3(7H)-one) Microelectrode with SOD on ZnO
ACCEPTED MANUSCRIPT 46
5. Hydroxyl radical, HO• 5.1. Definition and properties of HO•
RI
PT
Hydroxyl radical (HO•), the result of homolytic cleavage of water (H2O --> HO• + H•), is the most reactive ROS. HO• is a very powerful oxidant; E0 of the pair HO•/H2O is +2.3 V [368]. The rate constants of reactions of HO• with many molecules range from 109 to 1010 M-1 s-1 [369,370]. HO• has one unpaired electron and is able to accept an electron from almost any molecule in its immediate vicinity [371].
SC
5.2. Formation and action of HO• in photosynthetic tissues
MA
NU
Although HO• is known to be important especially in the chloroplast, HO• has received relatively little attention in plant biology. In biological systems the main HO• producing reaction is the metal-catalyzed Haber-Weiss reaction consisting of the Fenton reaction (reaction 15) and reduction of the transition metal ion by O2• (reaction 33) (Chapter 2; [1,372]). In spite of its high reactivity HO• plays a role in defense reactions in mammalian cells [373,374].
AC CE P
TE
D
Plants do not have specific scavengers for HO•, but the scavenging systems for H2O2 and O2• may protect against formation of HO• by limiting the Fenton reaction (reaction 15). It is assumed that plants prevent excess HO• production and subsequent damage by maintaining low levels of transition metals and sufficient levels of antioxidants in cellular compartments that might contain HO• [375]. The mechanisms maintaining appropriate levels of transition metals may also enable plants to target certain areas for controlled destruction [375]. For example, controlled destruction of cell wall pectins and xyloglucans for germination and elongative growth is believed to require HO• [376,377]. In the cell wall, HO• production has been proposed to involve soluble/cell wall bound proteins such as class III haemcontaining peroxidases (e.g. [372,378]), possibly in conjunction with cell wall associated SODs [379]. NADPH oxidases may also produce extracellular HO• [275,380]. Copper ions are readily available in cell walls and may catalyze Fenton chemistry, especially in roots [378,381]. Chloroplasts are likely the main intracellular sites of HO• production in plants due to the presence of transition metals (Fe, Cu, Mn) that are essential for the function of many transition metal binding proteins of the photosynthetic electron transfer chain [375,382]. Within chloroplasts the acceptor side of PSI is the principal site of H2O2 and O2• production [383,384], and therefore also HO• is thought to be mainly formed at PSI through the iron-catalyzed Haber-Weiss reaction (reactions 13‒15 and 33; [385]). However, with PSII being the main site for molecular oxygen evolution, also the role of PSII in HO• formation has been studied (reviewed by [2]). Formation of HO• at the acceptor side of PSII is thought to involve the reduction of metal center bound peroxides, such as non-heme iron bound peroxide. This reaction is similar to the Fenton reaction except that the reduction of the bound peroxide proceeds as an inner sphere electron transfer reaction [2,386]. Another possibility for HO• production at the acceptor side of PSII is the reduction of free peroxides in a typical Fenton reaction, with detached Mn2+ from the oxygen evolving complex of PSII serving as a possible reservoir of transition metal ions [386]. The oxygen evolving complex may also promote formation of HO• by acting as a source of H2O2 [387], especially after heat induced damage [388]. Yamashita et al. [388] proposed that HO• could be, along with 1O2, one
ACCEPTED MANUSCRIPT 47
PT
of the major ROS causing damage to PSII, particularly under stress. Also components of the mitochondrial electron transfer chain can produce HO•, but in plants the mitochondrial contribution is thought to consist mainly of production of H2O2 that can leak to other cell compartments to yield HO• in Fenton chemistry [375,389].
5.3. Reactions of HO•
RI
HO• participates in several typical reactions:
abstraction of hydrogen atom with formation of H2O and radical of substrate (36);
SC
HO• + RH H2O + R•
(36)
MA
NU
addition to double bonds with formation of a hydroxylated radical (37);
(37)
TE
AC CE P
HO• + SCN + OH + SCN•
D
electron transfer reactions leading to formation of a neutral radical (38.1) or a cation radical [390] (38.2); (38.1)
(38.2)
formation of aromatic-OH adduct due to a reaction of an aromatic compound with HO• is one of the methods for HO• detection with HPLC-MS. For example, HO• can react with phenylalanine to form isomers of tyrosine, reaction (39) [391]. Isomers of tyrosine are rather stable and not normally present in proteins and can serve as HO• traps in biological samples [392].
(39)
ACCEPTED MANUSCRIPT 48 Interaction of HO• with many aquated metal cations is an electron transfer reaction (40) that proceeds via two different mechanisms, both with a rate constant of ~108 M-1 s-1 [369]. HO• + Mn+ M(n+1)+ + OH
(40)
(41)
AC CE P
TE
D
MA
NU
SC
RI
PT
HO• initiates lipid peroxidation resulting in hydrogen abstraction from a pentyl group of an unsaturated fatty acid and formation of a radical that interacts with 3O2 to form a peroxyl radical (ROO•) with a rate constant of ~108 M-1 s-1 [393], reaction (41).
5.4. Lifetime and diffusion distance of HO• The rates of the reactions of HO• with biomolecules are usually limited by diffusion, and therefore reactions of HO• with biomolecules take place close to the site of HO• generation. The lifetime of HO• is thought to be in the order of nanoseconds [394,395], and recent simulations yielded a lifetime of only 30 ps in aqueous environment [371]. The diffusion distance of HO• has been estimated to be only a few molecular diameters from the place of origin [395]. The lifetime of HO• has not been measured in vivo.
5.5. Detection methods of HO• 5.5.1 Spectroscopic methods DMSO. Dimethyl sulfoxide (DMSO) has been used to detect HO• by colorimetry (absorbance at 420 nm) of methanesulfinic acid (MSA) formed in the oxidation of DMSO by HO• [396]. DMSO is only weakly toxic [397,398] and MSA is stable at pH 5‒9 [396]. Using DMSO, HO• formation has been witnessed in bean cowpea nodules, cucumber seedlings [399] and in leaves of duckweed and ryegrass [398]. In DMSO
ACCEPTED MANUSCRIPT 49 trapping, MSA must be extracted from the samples before colorimetry because a reaction between diazonium salts and MSA is required for the colorimetric assay [396‒400].
SC
RI
PT
Deoxyribose degradation. Free deoxyribose sugars yield malondialdehyde (MDA) upon degradation by HO• [401], and this observation has been utilized in HO• detection via detection of TBA-reactive substances (TBARS, see 6.4.5) [402]. Use of deoxyribose sugars as HO• detectors has been criticized for both its accuracy and its utilization in biological systems where free deoxyribose is not abundant [402]. In plant sciences this method has nevertheless been used to detect HO• formation in soybean leaves [403] and in wall fractions of isolated onion bulb cells [404].
AC CE P
TE
D
MA
NU
5.5.2. EPR detectable probes DMPO, DEPMPO, EMPO, 4-POBN. EPR spectroscopy is the most widely used method for detection of HO• in photosynthetic materials [2,269,375]. EPR probes that are used for HO• detection (DMPO, DEPMPO, EMPO) form adducts with O2• as well (see 4.5.2; [269]), but the EPR spectra of O2• and HO• probe adducts are so distinct that they allow differentiation between the two ROS, especially in the case of DMPO and EMPO. DEPMPO adduct spectra are more difficult to analyze [269,405]. Furthermore, HO• adducts produce diamagnetic species with Fe(III) and O2• and decay rapidly; both reactions result in loss of the EPR signal [152]. So even though HO• adducts can be distinguished, the actual amount of HO• being formed in a system may not be obtained with simplistic analyses of the EPR signal [269]. Taking into account the fact that decay of e.g. DMPO/O2• adducts yields HO• and that different probes act as scavengers of many ROS [269], special care needs to be taken in planning experiments where one and the same spin trap probe is used for the detection of O2• and HO•, as the naturally occurring formation of HO• is tightly linked to the concentrations of free O2•, free peroxides and transition metals [372,406]. Additional steps to confirm the precursors of HO• production in a studied system can be taken by adding SOD and CAT individually or together into the spin trapping mixture, thus pinpointing the prevalent HO• production pathway [407]. Also chelates such as ethylenediamine-N,N,N’,N’-tetraacetic acid (EDTA), diethylenetriamine-N,N,N¢,N’,N’’-pentaacetic acid (DTPA) and deferoxamine mesylate (Desferal) are commonly added into the buffers used in OH• measurements in order to reduce the amount of free transition metals, but use of these chelators should also be considered carefully as they themselves can stimulate ROS production [408,409]. 4-pyridyl-1-oxide-N-tert-butylnitrone (4-POBN) is an EPR probe that is used more spesifically for HO• detection. 4-POBN also reacts with O2• but the adduct is so unstable compared to a HO• adduct that almost no O2• adducts are detected [269]. In concentrations that are commonly used in biological studies (50 mM), 4-POBN can react with H2O2 in a reaction involving peroxidases, yielding 4-POBN•/4POBN spin adduct [410]. In addition to the specificity of 4-POBN, adducts produced in typical experiments with 4-POBN and HO• are considered to be stable when compared to other HO• reactive EPR probes [269,411]. The stability of 4-POBN adducts is, however, a result from the common practice of using HO• scavenging agents such as DMSO, ethanol or formate in conjunction with HO• probes [269,407,412‒415]. These scavengers react with HO• to form carbon centered radicals, such as αhydroxyethyl in the case of ethanol [413], which in turn form stable adducts with the spin trap. In the case of 4-POBN, the HO• adduct is stable because 4-POBN is practically always used with ethanol and the
ACCEPTED MANUSCRIPT 50
RI
PT
detected product is an adduct of 4-POBN and a carbon-centered hydroxyethyl radical [416‒420]. It is important to take the additional effects of ethanol (from 170 to 850 mM) into account when planning spin trap experiments with 4-POBN [269]. The half-life of the actual 4-POBN/HO• adduct has been shown to be only 1 min in Hank’s Buffer Solution, much shorter than the half-life of the DMPO/HO• adduct (2 h) in almost the same conditions [421]. The life time estimates for different spin trap adducts derived from organic chemistry (e.g. [152]) can be difficult to translate into biological context, and therefore EPR measurements of spin adducts should be done as fast as possible.
MA
NU
SC
Although there are many issues concerning in vivo detection of HO•, mainly the very short lifetime of HO• and the yet uncharacterized production pathways of other interfering ROS [375,415], EPR spectroscopy has been used in vivo in plants. In vivo measurements of HO• directly from tissues infiltrated with 4-POBN have been done by Renew et al. [422] and Deng et al. [420] on intact cucumber root and tobacco leaf cutouts, respectively. Measurements done by exposing the intact tissue to the spin trap and then measuring EPR from ground tissue are more numerous. With 4-POBN HO• has been detected from roots [407,418], coleoptiles [416,417], plant radicles and endosperm caps [423] and from potato tubers [419]. DMPO, in turn, has been used to detect HO• from roots by Demidchik et al. [424].
AC CE P
TE
D
In vitro HO• has been detected with DMPO from isolated PSII membranes [285,425] and from isolated rapeseed hypocotyl tissue [414]. DEPMPO [387,426‒428], EMPO [388,429,430] and 4-POBN [386,387] have also been used in isolated PSII membranes. DMPO has been utilized in many photosynthesis studies to examine HO• formation in isolated PSII antenna complexes [431] and isolated thylakoid membranes [97,432]. Both DEPMPO [275] and 4-POBN [380] have been used to trap HO• in isolated plant plasma membranes, whereas DEPMPO has been used to measure HO• from pea cell wall isolates [433,434] and from the apoplast of cucumber leaf [435]. 5.5.3. Fluorescent probes Fluorophorenitroxide proxyl fluorescamine (PF). Alternatives to EPR include 5-((2-carboxy)phenyl)-5hydroxy-1-((2,2,5,5-tetramethyl-1-oxypyrrolidin-3-yl)methyl)-3-phenyl-2-pyrrolin-4-one sodium salt (fluorophorenitroxide proxyl fluorescamine, PF) [421,436]. PF works as a spin trap for HO• in a similar manner as 4-POBN. A reaction of PF with the methyl radical produced in a reaction between HO• and DMSO eliminates the fluorescence quenching nitroxide group of PF [421]. Other radicals such as O2• do react with PF, but with the use of DMSO the probe is described by [421] as HO• specific. The excitation/emission wavelengths of PF are 380/485 nm. HO• formation has been fluorometrically detected from aliquots of PF media used for experimental incubations of barley root tip segments [436]. The use of UV wavelengths for excitation limits the use of PF in biological material. Terephthalic acid. Liu et al. [437] detected the formation of HO• by benzene-1,4-dicarboxylic acid (terepthalic acid, TPA) in supernatant of pea pod cell wall homogenate. TPA traps HO• and forms monohydroxy terephthalate (TPA-OH) that fluoresces at 350‒550 nm with excitation at 362 nm [437]. While the signal producing reaction of TPA is HO• specific and the formed adduct is relatively stable [437,438], TPA is known to react with other radicals without producing a fluorescent adduct [439]. UV absorption by biological materials may limit the use of TPA as well.
ACCEPTED MANUSCRIPT 51
PT
5.5.4. Microelectrode An electronic HO• microsensor, based on a change in the conductivity of polyaniline occurring due to a reaction with HO• was recently published [440], though it has not yet been used with photosynthetic material.
AC CE P
TE
D
MA
NU
SC
RI
5.5.5. Summary of HO• detection Because of its very short lifetime and unspecific reactivity, HO• is a difficult research subject and its function remains easily overshadowed by other ROS [441]. Furthermore, conclusions drawn from the physiological responses to H2O2 might be misleading due to the possible formation of HO• from H2O2 by the Fenton reaction [442]. Most methods used to measure HO• from plant material are not very specific (Table 6). Many methods relying on absorption/emission of HO• specific detector molecules are difficult to use in plants because the absorbance and fluorescence spectra of photosynthetic tissues often overlap strongly with the wavelengths of the signal.
ACCEPTED MANUSCRIPT 52
Specificity
Stability (t1/2)
Absorbance at 420 nm
HO• specific [396]
Stable [396]
Deoxyribose (2Deoxy-D-ribose) DMPO (5,5-dimethyl1-pyrroline N-oxide)
Absorbance at 530 nm EPR
Not very specific [402] Reacts also with O2• and other radicals [413]
-
Toxicity
RI
Method
MA
NU
SC
Compound (IUPAC name) DMSO
PT
Table 6. Chemicals used for the detection of HO•. The stability data indicate the stability of the signal after reaction with HO• in plant tissue. LogP values have been obtained from Chemspider [19]; estimated LogP values were determined using a LogP prediction tool ALOGPS 2.1 [20]. - = no data.
AC
CE DEPMPO (5EPR (diethoxyphosphoryl)5-methyl-1-pyrrolineN-oxide)
Reacts also with O2• and other radicals [286]
Stable [286]
EMPO (5(ethoxycarbonyl)-5methyl-1-pyrroline Noxide) 4-POBN (4-pyridyl-1oxide-N-tert-
EPR
Reacts also with O2• and other radicals [287]
-
EPR
HO• specific [269]
HO• adduct:
Not very toxic Slower growth of animal cells at 25 mM [297]; Plant root cells viable after treatment [441]; No effect on photosynthesis [96] Slower growth of animal cells at 25 mM [297]; No effect on photosynthesis [427]
PT ED
2 h [413]
Not very toxic [397]
αhydroxye
Hydrophobicity (LogP) -1.35 [19]
-1.90 (estimated) 0.34 (estimated)
Notes Optimal pH 5‒9; detected molecule is methanesulfic acid [396] Can be used to detect carbon centered radicals [413]
0.47 (estimated)
Can be used to detect carbon centered radicals [286]
Slower growth of animal cells at 25 mM [297]
0.41 (estimated)
-
-0.20 (estimated)
Can be used to detect carbon centered radicals [287] Commonly used with ETOH
ACCEPTED MANUSCRIPT 53
Fluorescence (Ex 380, Em 485 nm)
Reacts also with O2• [421]
Fluorescence (Ex 362, Em 350‒550 nm)
PANI electrode based on conductance change of polyaniline
Conductance measurement
May react with other radicals, but signal is HO• specific [439] Specific to HO• [440]
PT
[413] H 7 optimal for HO• adduct [411]
RI
thyl radical adduct: stable for >8 min [413]
-
-
Can be used to detect carbon centered radicals [436]
Stable [438]
-
2.00 [19]
-
Stable [440]
-
-
-
AC
CE
PT ED
MA
NU
Fluorophorenitroxide proxyl fluorescamine (5-((2carboxy)phenyl)-5hydroxy-1-((2,2,5,5tetramethyl-1oxypyrrolidin-3yl)methyl)-3-phenyl2-pyrrolin-4-one sodium salt) Terephthalic acid (Benzene-1,4dicarboxylic acid)
With O2• less than 1 min [413]; Without O2• 32 h [411] -
SC
butylnitrone)
ACCEPTED MANUSCRIPT 54
6. Peroxyl radical (ROO•), alkoxyl radical (RO•) and hydroperoxides 6.1. Definition and formation
PT
Peroxyl and alkoxyl radicals and hydroperoxides are best known as products of peroxidation of lipids. Other ROS, especially 1O2 and HO•, initiate reactions that lead to formation of these ROS.
RI
6.2. Reactions of ROO•, RO• and hydroperoxides
SC
The appearance of ROO• leads to formation of an alkoxyl radical (RO•) and lipid hydroperoxides (ROOH). The scheme of reactions leading to the appearance of these forms and, consequently, lipid peroxidation is well known, reactions (42‒44).
NU
HO• + RH R• + H2O
(42)
Initiation of the process and the appearance of lipid radicals;
MA
R• + 3O2 ROO•
(43)
addition of molecular oxygen to the carbon-centered radical (R•) yielding the ROO•;
D
ROO• + RH ROOH + R•
(44)
AC CE P
TE
abstraction of a hydrogen from another lipid molecule by ROO•, like reaction (42). Reaction (44) is known as the chain propagation reaction because this reaction supplies R• for reaction (43) and thus leads to accumulation of ROOH. The spontaneous decomposition of ROOH (reactions 45 and 46) is unfavorable because the dissociation energies of ROO-H and RO-OH bonds are 376 kJ/mol and 184 kJ/mol, respectively [443]. Spontaneous decomposition of ROOH can be observed in perturbed conditions, e.g. at a high temperature. ROOH ROO• + H•
(45)
ROOH RO• + HO•
(46)
Non-enzymatic decomposition ROOH via either their reduction or oxidation by transition metal is one of the most significant ways of ROOH decomposition (reactions 47.1 and 47.2). ROOH + Mn+ RO• + OH + M(n+1)+
(47.1)
ROOH + M(n+1)+ ROO• + H+ + Mn+
(47.2)
Interaction of ROOH with radicals can also lead to decomposition of ROOH due to abstraction of a hydrogen atom and formation of ROO• (reaction 48). ROOH + R• ROO• + RH
(48)
ROOH can also react with the C-H bond of saturated and alkylaromatic hydrocarbon via reaction (49) [444].
ACCEPTED MANUSCRIPT 55 ROOH + RH RO• + R• + H2O2
(49)
Dismutation of ROOH yields both RO• and ROO•, reaction (50). ROOH + ROOH RO• + ROO• + H2O
PT
(50)
RI
Decomposition of ROOH is a key reaction in the chain reaction of lipid peroxidation because of additional formation of ROO• or RO• species that can abstract a hydrogen from a lipid molecule, reactions (44) and (51). ROO• + RH ROOH + R•
(51)
NU
SC
The redox potentials (E0' at pH 7) of pairs RO•/ROOH and ROO•/ROOH are 1.6 V and from 0.77 V to 1.44 V, respectively [368]. The rate constant of the reaction of RO• with RH is 107 M-1 s-1, which is five orders higher than the rate constant of the reaction of ROO• with RH, 102 M-1 s-1 [168]. R•, ROO• and RO• can also interact with each other in several reactions [445,446]:
MA
The radical-radical reaction between two RO• can lead to formation of a carbonyl compound in excited state, reaction (52); RO• + RO• ROH + R=O*
(52)
AC CE P
TE
D
Reaction 52 is rather unlikely in biological systems because of the high reactivity of RO• toward lipids and proteins. A radical-radical reaction between two ROO• leads to the formation of an excited carbonyl compound (reaction 53.1) or 1O2 (reaction 53.2). The latter is called the Russell mechanism. The rate constant for the termination stage in the Russell reaction varies between 106 and 108 M-1 s-1 [445].
(53.1)
(53.2)
RO• and ROO• can react with R• to form dimers, reactions (54) and (55) [443,447]. RO• + R• ROR
(54)
ROO• + R• ROOR
(55)
Numerous experimental data on the kinetics of hydroperoxide chain process is represented in a recent monograph [444].
ACCEPTED MANUSCRIPT 56 Reactions (52) and (53.1) are associated with chemiluminescence (380–460 nm) due to relaxation of the excited carbonyl compound. This chemiluminescence is used to measure lipid peroxidation [448] (see 6.4.2).
SC
RI
PT
In reactions with lipid molecules, RO• can cause a homolytical scission (β-cleavage) of a C-C bond on either side of the alkoxyl group to form a wide range of aldehydes and oxo fatty acids [449]. Formaldehyde, acetaldehyde, propionaldehyde, acrolein, malondialdehyde, (E)-2-hexenal, (Z)-3-hexenal, 4-hydroxy-(E)-2-hexenal and other carbonyls are found to be products of degradation of polyunsaturated fatty acids [450], and therefore aldehydes can be used as markers of lipid peroxidation. Measurement of coloured substances produced in a reaction between aldehydes and thiobarbituric acid (TBA) is the most commonly applied method [451]. A new method for a comprehensive analysis of oxylipin carbonyls (quantitation and structural estimation) in plants with reverse-phase HPLC has recently been established [450,452].
MA
NU
Reduction of ROOH by thiol groups of peroxiredoxins is an important scavenging mechanism. This reaction proceeds as a 2-step reaction like reaction with H2O2, reactions (55.1) and (55.2). The Km values and rate constants of the reaction between peroxiredoxins and ROOHs depend on both ROOH species and on the peroxiredoxin, and the Km values range from 40 nM to 1.7 mM and the rate constants from 104 M-1 s-1 to 107 M-1 s-1 [453].
D
ROOH + RSH RSOH + ROH
(55.2)
TE
RSOH + RSH RSSR + H2O
(55.1)
Tocopherols and carotenoids can act as antioxidants preventing lipid peroxidation [454,455].
AC CE P
6.3. Lifetime and diffusion distance of ROOHs The lifetimes of ROO•or RO• in the cells have been estimated to be 7 s and 1 µs, respectively [395]; these lifetimes would allow 5 % of the original ROS to be present at 0.16 mm or 60 nm from the site of formation in a viscous cellular environment, respectively. ROOHs are generally less reactive than ROO•or RO•.
6.4. Detection methods of ROOHs 6.4.1. Spectroscopic methods Lipid hydroperoxides contain fatty acids with conjugated dienes, and lipid hydroperoxides can be detected spectroscopically by measuring the absorbance of the dienes at 230‒235 nm. A high sensitivity for conjugated dienes is achieved with double-derivative spectroscopy [456]. However, this method can only be used for pure extracted material. 6.4.2. Chemiluminescence and thermoluminescence Excited carbonyl compounds, produced due to degradation of ROOHs, emit light at 380–460 nm. The chemiluminescence method (see Table 7 for a list of chemical methods for ROOH detection) can be applied for detection of ROOHs under both spontaneous and catalytic degradation of ROOH. ROOH-derived chemiluminescence has been measured from plant leaves with a sensitive camera [446]. ROOH-derived chemiluminescence resulting from sodium hypochlorite-induced degradation of ROOH was proposed as a sensitive assay for ROOH [457]. In plant material, addition of acridine in reaction medium significantly improved the linear correlation between luminescence intensity and amount of ROOHs [458]. A high-
ACCEPTED MANUSCRIPT 57 temperature thermoluminescence band peaking at 115‒130 °C was shown to correlate with the amount of lipid peroxides in chlorophyll-containing material [459]. The method has been frequently used for the detection of lipid peroxides and assessment of oxidative stress especially in photosynthesis research; see e.g. [460,461].
MA
NU
SC
RI
PT
6.4.3. Dyes Xylenol orange. Xylenol orange, used as dye for detection of ROOH is know as the ferrous xylenol orange method (FOX method). The FOX method is based on the oxidation of ferrous ions Fe(II) to ferric ion Fe(III) by ROOH in acidic conditions. The reaction of ferric ion with xylenol orange results in formation of a complex with an absorption maximum at 560 nm [462‒464]. Unfortunately, the stoichiometry between the amount of the xylenol orange complex and the amount of peroxide depends on peroxide species, ranging from 1.8 to 5.4 moles of xylenol orange complex per 1 mole of peroxide [465]. Therefore, a separate calibration curve is needed for each type of peroxide. The FOX method can be applied for both aqueous and organic solvents and used for detection of both hydrophilic (FOX1 method) and lipophilic (FOX2 method) peroxides. The FOX1 and FOX2 methods are not specific to ROOH because ferrous ions can be oxidized by other oxidizing agents. To accurately quantify ROOHs, the measurements are done in the presence and absence of triphenylphosphine which selectively reduces ROOHs to corresponding alcohols. To prevent peroxidation during analysis, butylated hydroxytoluenes are usually used [466,467].
TE
D
Thiocyanate. Thyocyanate (SCN) has also been used for detection of ROOH with the oxidation of ferrous iron. The reaction of ferric ion with the thiocyanate ion leads to formation of complexes absorbing at 470‒505 nm [462]. In contrast to the FOX methods, the thiocyanate-iron method can be used for detection of ROOH in material containing carotenoids [468].
AC CE P
Iodide. ROOHs oxidize iodide to iodine and iodine react with iodide, yielding triiodide anion (I3-). I3- absorbs at 290 and 360 nm [198,469]. A modified iodometric method has recently been used for determination of ROOH in plants [470]. The iodometric method is sensitive to oxygen that must be removed from the reaction medium. Iodometry is not specific for ROOH because H2O2 can oxidize I- to I2 (see 3.5.3; [471]), and therefore H2O2 should be removed before measurement. The stoichiometry between amounts of peroxides and iodine is 1:1. This method is known as an International Dairy Federation method (IDF method, 74A:1991). 6.4.4. Fluorescent probes Diphenyl-1-pyrenylphosphine (DPPP). DPPP can be used as a fluorescent probe to monitor lipid peroxidation in cell membranes. DPPP reacts with organic hydroperoxides stoichiometrically to yield DPPP oxide (DPPPNO) which emits at 380 nm with excitation at 351 nm [472]. DPPP readily reacts with lipophilic peroxides like methyl linoleate hydroperoxide but not with tert-butyl hydroperoxide. The DPPP method can be used for measurement of lipid peroxidation within membranes. DPPP also rapidly reacts with H2O2, and therefore this method can only be used in vitro with H2O2 free samples. [3-(4-phenoxyphenylpyrenylphosphino) propyl]triphenylphosphonium iodide (MitoDPPP). The reactivity of MitoDPPP is similar as that of DPPP, and MitoDPPP can be oxidized by peroxides including H2O2 but not by free radicals like O2• or HO•. The oxidation of MitoDPPP with hydroperoxides yields MitoDPPPO which has 35 times higher fluorescence intensity than MitoDPPP (ex 350 nm, em 380 nm) [473]. The reaction rate of MitoDPPP
ACCEPTED MANUSCRIPT 58 with peroxides depends on the type of the peroxide. MitoDPPP is supposed to localize within lipid membranes and can be used for detection of ROOH in living cells.
SC
RI
PT
2-(4-diphenylphosphanylphenyl)-9-(1-hexylheptyl)anthra[2,1,9-def,6,5,10-d′ e′ f ′ ]diisoquinoline-1,3,8,10tetraone (Spy-HP). Spy-HP reacts rapidly with hydroperoxides such as m-chloroperbenzoic acid (MCPBA) and cumene hydroperoxide to form an oxidized derivative (Spy-HPO) that emits strong fluorescence at 535 nm when excitated at 524 nm. The formation of Spy-HPO shows strict 1:1 stoichiometry with the amount of peroxides and its affinity to lipophilic hydroperoxides is high, allowing measurements at nM concentration. Spy-HPO is stable against autoxidation and photobleaching and the reaction of Spy-HP with O2•, H2O2 and HO• is very slow compared to the reaction with lipophilic peroxides [474]. Spy-HP has been used to detect lipid peroxides in photosynthetic membranes [475].
D
MA
NU
2-(4-diphenylphosphanyl-phenyl)-9-(3,6,9,12-tetraoxatridecyl)-anthra[2,1,9-def:6,5,10-d9e9f9]diisoquinoline1,3,8,10-tetraone (Liperfluo). Liperfluo is oxidized by peroxides to Liperfluo-Ox with (ex 524 nm, em 535 nm). Liperfluo reacts much more slowly with various other ROS including H2O2, O2•, HO• and NO• than with methyl linoleate hydroperoxide [476]. Liperfluo shows only little cytotoxicity in SH-SY5Y cells and can be used for detection of peroxides in cells. It was demonstrated that Liperfluo is more useful as a probe than Spy-LHP for live-cell fluorescence imaging [476]. The fluorescence yield of Liperfluo-Ox is linearly dependent on the concentration of methyl linoleate hydroperoxide.
AC CE P
TE
Mito-Bodipy-TOH is a fluorescent peroxyl radical probe absorbing at 537 nm and emitting at 572 nm. The probe consists of an α-tocopheryl moiety, a fluorogenic reporter and a mitochondria-targeting segment by which the probe targets to the inner mitochondrial membrane [477]. The probe reacts with lipid peroxides but its specificity has not been extensively studied. Mito-Bodipy-TOH has not been used with photosynthetic material. 6.4.5. Measurements of ROOH-derived species TBA reactive substances (TBARS). Measurements of TBARS are commonly used for lipid peroxidation assays. The method is based on the ability of ROOH-derived products like malondialdehyde to react with TBA to form TBA-adducts at an elevated temperature in acidic conditions. TBA-adducts can be easily detected by absorbance at 532 nm [478,479] or by fluorescence at 553 nm [480]. Other ROOH-derived aldehydes can also react with TBA to form complexes [481]. Decomposition of lipid peroxides to TBARS during analysis results in formation of radicals that can promote additional lipid peroxidation, which may lead to overestimation of ROOH content. Protective substances (like butylated hydroxytoluene) are usually used to inhibit the lipid peroxidation [482,483]. Traces of iron (II) can also increase TBARS by accelerating decomposition of ROOHs, and therefore chelators are added to the medium [483]. HPLC analysis of TBA-adducts allows measurement of the MDA-(TBA)2-adduct only [484]. Using gas chromatography-mass spectrometry (GC-MS) it has recently been shown that lipid peroxidation is accurately measured with the TBA method in most plants [485]. 2,4-dinitrophenylhydrazine (DNPH). This method is based on extraction of lipid-derived carbonyls and their derivatization with 2,4-dinitrophenylhydrazine and separation with HPLC and LC/MS [452]. The carbonyl species can be identified from the MS/MS spectrum. The structure of ROOH-derived carbonyls is determined from the position of the hydroperoxide group in the lipid molecule. A HPLC based comprehensive analysis of carbonyls using DNPH for derivatization can be used for detection of fatty acids involved in peroxidation [486].
ACCEPTED MANUSCRIPT 59
IP
T
Table 7. Methods of detection of ROOH species. All mentioned chemicals show a stable, non-light-sensitive signal. LogP values have been obtained from Chemspider [19]; estimated LogP values were determined using a LogP prediction tool ALOGPS 2.1 [20]. - = no data. Method
Specificity
Hydrophobicity (LogP)
Notes
Ferric ion xylenol orange complex (absorption maximum at 560 nm) [462]
1.85 (estimated), can be used for both aqueous and organic solvents [466]
Various stoichiometries between amount of xylenol orange complex and amount of peroxide [465]
Thiocyanate (IDF method)
Ferric ion thiocyanate complex (absorption at 470–505 nm) [462,468]
0.22 (estimated), can be used for both aqueous and organic solvents [462,468]
Stoichiometry between amount of complex and amount of peroxide is not strict [462]
Iodide
Formation I3Absorption at 290–360 nm [198,469] Oxidized form has high fluorescence (Ex 351, Em 380) [472] Oxidized form has high fluorescence (Ex 350, Em 380) [473]
Nonspecific. Triphenylphosphine shuld be used to evaluate unspecificity [463,466] Nonspecific. Triphenylphosphine should be used to evaluate unspecificity [462] -
1.04 [19]
Experiments can be carried out in anaerobic condition dou to sensitivity to oxygen [198,469] -
MitoDPPP ([3-(4phenoxyphenylpyre nylphosphino) propyl]triphenylpho sphonium iodide) Spy-HP (2-(4Oxidized form has high diphenylphosphanyl fluorescence (Ex 524, phenyl)-9-(1Em 535) [474] hexylheptyl)anthra[2 ,1,9-def,6,5,10-
US
MA N
TE D
CE P
Can react with H2O2 [472]
8.52 (estimated), lipophilic [472]
Can react with H2O2 [473]
Lipophilic [473]
-
Specific for lipohilic peroxides. Slowly react with hydrophilic peroxides and H2O2 [474]
Very lipohilic [474]
Useful only for organic solvents [474]
AC
DPPP (Diphenyl-1pyrenylphosphine)
CR
Compound (IUPAC name) Xylenol orange (FOX method)
ACCEPTED MANUSCRIPT
TE D
-
CE P
-
IP
-
US
CR
-
MA N
Specific for lipohilic peroxides. Slowly react with hydrophilic peroxides and H2O2 [476]
AC
d′ e′ f ′ ]diisoquinoline1,3,8,10-tetraone) Liperfluo 2-(4High fluorescence (Ex diphenylphosphanyl- 524, Em 535) [476] phenyl)-9-(3,6,9,12tetraoxatridecyl)anthra[2,1,9def:6,5,10d9e9f9]diisoquinolin e-1,3,8,10-tetraone) TBA Complex with MDA with maximum of absorption at 532 nm [478,479] Mito-Bodipy-TOH Fluorescence (Ex 537, Em 572) [477]
T
60
-0.39 (estimated)
TBA can react with other carbonyls in addition to MDA [481]
-
Targeted for mitochondrial inner membrane. Not yet tested with photosynthetic material.
ACCEPTED MANUSCRIPT 61
7. Ozone, O3
SC
RI P
T
Ozone is an oxygen molecule consisting of three oxygen atoms (O3). O3 is a strong oxidant (E0 of the couple O3 + 2H+ + 2e-/O2 + H2O is 2.076 V [154]). O3 may also take part in 1-electron reactions, forming the ozonide radical O3-. Thermodynamic calculations suggest a value 1.6 V for the E0 value of the couple O3/O3• [487], although a lower value of 1.0 V could also be suggested [487]. External O3 does not penetrate to living plant cells [488] and therefore detection of O3 will not be discussed. However, external O3 penetrates to the apoplastic space through stomata and causes the formation of other ROS in the apoplast and thereby initiates plant defense signaling [8].
(56)
AC CE
PT
ED
MA
NU
The principal reaction of O3 with organic compounds is the addition to double bonds and cleavage of unsaturated compounds (ozonolysis). O3 reacts with a double bond to form ozonide and via the Criegee mechanism, reaction (56) [489].
Alternatively to formation of secondary ozonides, the decomposition of primary ozonides in aqueous medium can yield carbonyls and H2O2. The ozonolysis of ethylene yields two molecules of formaldehyde and one molecule of H2O2, reaction (57). Formation of H2O2 in reactions resembling reaction (57) may contribute to the ozone-induced oxidative burst in plant apoplast [8].
ACCEPTED MANUSCRIPT 62
ED
MA
NU
SC
RI P
T
(57)
AC CE
PT
The rate constant of the reaction of O3 with ethylene at 30 °C is 106 M-1 s-1 [490]. O3 can react with isoprene and others terpenes to give corresponding carbonyls [491]. It was also shown that ozonolysis of terpenes as -terpineol, limonene, and -pinene is associated with formation of HO•, reaction (58) [492].
(58)
The reaction of O3 with such amino acids as cysteine, tryptophan, methionine, tyrosine, histidine and other biologically important compounds like ascorbic acid and NADPH produces 1O2. The rate constants and yields of 1O2 formation for many relevant compounds are given by [493]. The yield of 1 O2 in the reaction of O3 with ascorbic acid was found to be pH dependent, indicating that 1O2 is produced in a reaction between O3 and the ascorbate anion, reaction (59) [494]. O3 is able to oxidize GSH to a sulfonic acid [495]. AsCH + O3 DHA + 1O2 + OH,
(59)
ACCEPTED MANUSCRIPT 63
where DHA is dehydroascorbate.
T
In addition to the classic ozonide forming reactions, electron transfer to ozone can lead to the formation of an ozonide anion radical, reaction (60). The ozonide anion radical is protonated very fast (k 1010 M-1 s-1), reaction (61) and then decomposed to hydroxyl radical and 3O2 [487,496].
RI P
IrCl63 + O3 IrCl62 + O3 O3 + H+ HO3
(60) (61)
SC
In the atmosphere, ozone promotes the appearance of nitrogen dioxide (NO2•) via reaction (62) with nitric oxide (NO•).
NU
O3 + NO• NO2• + O2
MA
8. Non-specific methods
(62)
8.1. Non-specific methods indicating the presence of ROS
PT
ED
Most direct ROS detection methods aim at detection and measurement of a specific type of ROS. Specificity is highly important because a method that detects several different ROS, each with different detection efficiency, has little predictive value for ROS-induced damage or ROS signaling. Several non-specific methods are, however, available in the market and they are sometimes used in the absence of specific ones. Non-specific reagents may also have advantages like low cytotoxicity or ability to localize to a specific cell compartment.
AC CE
H2DCFDA.The dyes 2′ ,7′ -dichlorodihydrofluorescein diacetate (H2DCFDA) and its derivatives 5-(and6)-chloromethyl-2′ ,7′ -dichlorodihydrofluorescein diacetate (CM-H2DCFDA) diffuse through the cell membrane and cell wall, but within the cell esterases remove the acetate groups and the chloromethyl group reacts with thiols. Therefore the dye cannot diffuse out of the cell, and a fluorescent oxidation product (DCF) can be detected by excitation at 492‒495 nm and emission at 517‒527 nm. CM-H2DCFDA has been used with Arabidopsis thaliana leaves; peeling off the epidermis is required to get the dye into leaf cells, and the dye responds to H2O2 treatment of the cells [23]. The chemical has also been used with cyanobacteria [130]. CM-H2DCFDA has been used as H2O2 detector (e.g. [497]), but because oxidation of CM-H2DCFDA depends on a Fenton-type reaction or on an unspecific enzymatic reaction with cytochrome c [498], a difference in the intensity of the fluorescence signal between two samples may indicate difference in H2O2, iron, copper or cytochrome c concentration. Also, UV and to a smaller degree visible light causes oxidation of DCFH [499], and DCF can produce O2• in light [500]. CellROX reagents. Molecular Probes Inc. produces several unspecific ROS detection reagents: CellROX® Deep Red Reagent with excitation and emission at 640 and 665 nm, respectively, CellROX® Orange Reagent (545/565 nm), CellROX® Green Reagent (485/520 nm) and OxyBurst (R) Green H2DCFDA, SE (2',7'-dichlorodihydrofluorescein diacetate, succinoimidyl ester). According to the website of the company (https://tools.lifetechnologies.com/content/sfs/manuals/mp10422.pdf; cited Feb 24, 2015), the Green Reagent binds to DNA and localizes to the nucleus in animal cells, and
ACCEPTED MANUSCRIPT 64
RI P
T
the OxyBurst Green H2DCFDA, SE penetrates to cells. These reagents have not been tested with photosynthetic material but might be useful, as the emission wavelengths of the Orange and Green reagents do not interfere with Chl a emission. The rate constants of these reactants with different ROS have not been published. The cell-impermeable OxyBurst Green H2HFF BSA reagent has been used for detection of apoplastic ROS production during the generation of systemic signals related to wounding [501].
NU
SC
HPF and APF. 3'-(p-hydroxyphenyl) fluorescein (HPF) and 3'-(p-aminophenyl) fluorescein (APF) were originally developed for the detection of HO• and peroxynitrite ion (ONOO-) [227]. In both cases a dearylation reaction converts a non-fluorescent dye to a fluorescent one, with excitation at 490 nm and emission at 515 nm. The dyes are cell permeable and resistant to autooxidation [227] and react with most ROS, including 1O2 and ROO• [24,227]. Data about their reactivity against O2• varies [24,227]. The reaction between APF and pure H2O2 is slow but becomes significant in the presence of a peroxidase [24].
ED
MA
TEMPO-9-AC. 4-((9-acridinecarbonyl)amino)-2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO-9-AC) is a free radical that is converted to a fluorescent form (with excitation at 366 nm and emission at 424 nm) in a reaction with another free radical like O2• [502]. Strong illumination of tobacco leaves after pinhole administration of TEMPO-9-AC did not cause detectable fluorescence [109]. In leaves, this probe remains mostly in the apoplast [109].
AC CE
PT
Ultra-weak luminescence, emitted by triplet excited carbonyls, pigments excited by excited carbonyls and by 1O2 has been shown to be associated with oxidative stress in plants, and mechanisms of production of light have been elucidated [503]. Long exposure time is required for quantitative imaging of ultra-weak luminescence.
8.2. Non-specific methods indicating reactions of ROS with biomolecules Proteins, DNA, lipids and various small biomolecules are known to react with ROS in vitro, but conclusive evidence for a reaction occurring in vivo or in an isolated system like thylakoid membrane preparation is difficult to obtain. Oxidation of proteins by ROS typically produces carbonyl groups in the amino acid side chains, and these can be detected with an antibody against 2,4-dinitrophenyl hydrazine (DNP) [504]. DNP-antibodies “Oxyblot™” can be used for general oxidative stress analysis. Reactive carbonyl species (RCS), defined as α,β-unsaturated aldehydes and ketones, are products of further reactions of lipid peroxidation [505]. Lipid peroxidation is initiated by reactive oxygen species, and therefore an increase in RCS, detected by the TBARS test, indirectly indicates an action of ROS. Examples of RCS occurring in plant material include malondialdehyde, 4-hydroxy-(E)-2nonenal, formaldehyde and acrolein. Unfortunately, the TBARS test fails to detect some physiologically relevant RCS like 4-hydroxy-(E)-2-nonenal [22].
9. Concluding remarks ROS are hot topics in plant biology because they function both as agents of damage and mediators of cellular signals. Understanding of both functions needs knowledge about the properties and reactions of each ROS, and the present review attempts to give an account of the biologically most
ACCEPTED MANUSCRIPT 65
relevant features of ROS, including redox potentials, key reactions and their rate constants, and lifetime and diffusion distance in biological material.
SC
RI P
T
The second aim of this review was to summarize the most relevant detection methods that have been used, or that are potentially usable, for experimental material deriving from plants, algae or cyanobacteria. The main focus is in photosynthetic tissues. We try to provide a listing of the properties of the method, including the possibilities to locate the ROS in a cell, toxicity of the detector substance and stability of both the detector substance and the signal. Unfortunately, the list often remains incomplete, because some topics have not yet been studied. We hope that the introduction to the methods may prompt future research of the missing topics and future development of better detection methods.
NU
To minimize artifacts, the following principles should be always be followed when chemical probes are used for ROS measurements.
MA
1. Ensure the specificity of the probe.
2. Check the toxicity of the probe to the study material and especially to photosynthesis.
ED
3. Ensure that the wavelengths and intensities of light used in the experiments do not cause artefacts like generation of ROS. 4. Investigate the cellular localization if the probe is used in vivo.
PT
5. Do not draw quantitative conclusions if the probe is unspecific.
AC CE
10. Acknowledgements
This study was financially supported by Academy of Finland (grants 259075 and 271832) and by Nordic Energy Research (AquaFEED project).
11. References
[1] B. Halliwell, J. Gutteridge, Free Radicals in Biology and Medicine, Oxford science publications, New York, 1999. [2] P. Pos íšil, Production of reactive oxygen s ecies by Photosystem II, Biochim. Biophys. Acta 1787 (2009) 1151‒1160. [3] B.B. Fischer, É. Hideg, A. Krieger-Liszkay, Production, detection, and signaling of singlet oxygen in photosynthetic organisms, Antioxid. Redox Signal. 18 (2013) 2145– 2162. [4] E. Tyystjärvi, Phototoxicity, in: L.D. Noodén (Ed.) Plant Cell Death Processes, Academic Press, San Diego, pp. 271-283, 2004. [5] E. Tyystjärvi, Photoinhibition of Photosystem II, Int. Rev. Cell Mol. Biol. 300 (2013) 243–303.
ACCEPTED MANUSCRIPT 66
[6] S.M. Gallejo, L.B. Pena, R.A. Barcia, C.E. Azpilicueta, M.F. Iannone, E.P. Rosales, M.S. Zawoznik, M.D. Groppa, M.P. Benavides, Unravelling cadmium toxicity and tolerance in plants: Insight into regulatory mechanisms, Env. Exp. Bot. 83 (2012) 33‒46.
RI P
T
[7] Y. Nishiyama, S.I. Allakhverdiev, N. Murata, Protein synthesis is the primary target of reactive oxygen species in the photoinhibition of photosystem II, Physiol. Plant. 142 (2011) 35‒46.
SC
[8] J. Kangasjärvi, P. Jaspers, H. Kollist, Signalling and cell death in ozone-exposed plants, Plant Cell Env. 28 (2005) 1021‒1036. [9] G. Miller, N. Suzuki, S. Ciftci-Yilmaz, R. Mittler, Reactive oxygen species homeostasis and signalling during drought and salinity stresses, Plant Cell Env. 33 (2010) 453‒467.
PT
ED
MA
NU
[10] C. Laloi, M. Havaux, Key players of singlet oxygen-induced cell death in plants. Front. Plant. Sci. 6 (2015) Art. 39. [11] S.S. Gill, N. Tuteja, Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants, Plant Physiol. Biochem. 48 (2010) 909‒930. [12] C.I. Cazzonelli, Carotenoids in nature: insights from plants and beyond, Funct. Plant Biol. 38 (2011) 833‒847. [13] K.J. Dietz, Peroxiredoxins in plants and cyanobacteria, Antioxid. Redox Signal. 15 (2011) 1129–1159. [14] P. Pos íšil, Molecular mechanisms of roduction and scavenging of reactive oxygen species by photosystem II, BBA-Bioenergetics 1817 (2012) 218–231. [15] L.A. del Río, ROS and RNS in plant physiology: an overview, J. Exp. Bot. 66 (2015) 2827–2837.
AC CE
[16] B. D'Autréaux, M.B. Toledano, ROS as signalling molecules: mechanisms that generate specificity in ROS homeostasis, Nature Rev. Mol. Cell Biol. 8 (2007) 813–824. [17] C.H. Foyer, S. Shigeoka, Understanding oxidative stress and antioxidant functions to enhance photosynthesis, Plant Physiol. 155 (2011) 93–100. [18] F.J. Schmitt, G. Renger, T. Friedrich, V.D. Kreslavski, S.K. Zharmukhamedov, D.A. Los, V.V. Kuznetsov, S.I. Allakhverdiev, Reactive oxygen species: Re-evaluation of generation, monitoring and role in stress-signaling in phototrophic organisms, Biochim. Biophys. Acta 1837 (2014) 835–848. [19] Chemspider, http://www.chemspider.com, referred to on August 20th to 30th, 2015. [20] I.V. Tetko, J. Gasteiner, R. Todeschini, A. Mauri, D. Livingstone, P. Ertl, V.A. Palyulin, E.V. Radchenko, N.S. Zerirov, A.S. Makarenko, V.Y. Tanchuk, V.V. Prokopenko, Virtual computational chemistry laboratory – design and description, J. Comput. Aid. Mol. Des. 19 (2005) 453–463. [21] V.V. Belousov, A.F. Fradkov, K.A. Lukyanov, D.B. Staroverov, K.S. Shakhbazov, A.V. Terskikh, S. Lukyanov, Genetically encoded fluorescent indicator for intracellular hydrogen peroxide, Nat. Methods 3 (2006) 281–286. [22] W.M. Nauseef, Detection of superoxide anion and hydrogen peroxide production by cellular NADPH oxidases, Biochim. Biophys. Acta 1840 (2014) 757–767.
ACCEPTED MANUSCRIPT 67
T
[23] K. A. Kristiansen, P.E. Jensen, I.M. Møller, A. Schulz, Monitoring reactive oxygen species formation and localisation in living cells by use of the fluorescent probe CMH2DCFDA and confocal laser microscopy, Physiol. Plant. 136 (2009) 369–383.
RI P
[24] M. Price, D. Kessel, On the use of fluorescence probes for detecting reactive oxygen and nitrogen species associated with photodynamic therapy, J. Biomed. Opt. 15 (2010) 051605.
SC
[25] C. Schweitzer, R. Schmidt, Physical mechanisms of generation and deactivation of singlet oxygen, Chem. Rev. 103 (2003) 1685–1757. [26] A.A. Krasnovskii, Photosensitized luminescence of singlet oxygen in solution, Biofizika 21 (1976) 748–749.
NU
[27] J.M. Wessels, M.A.J. Rodgers, Effect of solvent polarizability on the forbidden 1Δg → 3Σg- transition in molecular oxygen: A Fourier transform near-infrared luminescence study, J. Phys. Chem. 99 (1995) 17586–17592.
MA
[28] A.M. Falick, B.H. Mahan, R.J. Myers, Paramagnetic resonance spectrum of 1Δg oxygen molecule, J. Chem. Phys. 42 (1965) 1837‒1838.
ED
[29] M. Ruzzi, E. Sartori, A. Moscatelli, I.V. Khudyakov, N.J. Turro, Time-resolved EPR study of singlet oxygen in the gas phase, J. Phys. Chem. A 117 (2013) 5232‒5240.
AC CE
PT
[30] A. Krieger-Liszkay, Singlet oxygen production in photosynthesis, J. Exp. Bot. 56 (2005) 337–346. [31] C.J. Hoytink, Intermolecular electron exchange, Acc. Chem. Res. 2 (1969) 114– 120. [32] S. Santabarbara, E. Bordignon, R.C. Jennings, D. Carbonera, Chlorophyll triplet states associated with Photosystem II of thylakoids, Biochemistry 41 (2002) 8184– 8194. [33] É. Hideg, I. Vass, Singlet oxygen is not produced in Photosystem-I under photoinhibitory conditions, Photochem. Photobiol. 62 (1995) 949–952. [34] S. Cazzaniga, Z. Li, K.K. Niyogi, R. Bassi, L. Dall'Osto, The Arabidopsis szl1 Mutant Reveals a Critical Role of β-Carotene in Photosystem I Photoprotection, Plant Physiol. 159 (2012) 1745–1758. [35] A.U. Rehman, K. Cser, L. Sass, I. Vass, Characterization of singlet oxygen production and its involvement in photodamage of Photosystem II in the cyanobacterium Synechocystis PCC 6803 by histidine-mediated chemical trapping, Biochim. Biophys. Acta 1827 (2013) 689–698. [36] C. Flors, S. Nonell, Light and singlet oxygen in plant defense against pathogens: Phototoxic phenalenone phytoalexins, Acc. Chem. Res. 39 (2006) 293–300. [37] P. Pos íšil, A. Prasad, Formation of singlet oxygen and rotection against its oxidative damage in Photosystem II under abiotic stress, J. Photochem. Photobiol. B: Biol. 137 (2014) 39–48. [38] J.R. Kanofsky, B. Axelrod, Singlet oxygen production by soybean lipoxygenase isozymes, J. Biol. Chem. 261 (1986) 1099–1104.
ACCEPTED MANUSCRIPT 68
[39] D.K. Yadav, P. Pos íšil, Evidence on the formation of singlet oxygen in the donor side photoinhibition of Photosystem II: EPR spin-trapping study, PLoS ONE 7 (2012) e45883.
RI P
T
[40] M. Havaux, B. Ksas, A. Szewczyk, D. Rumeau, F. Franck, S. Caffarri, C. Triantaphylidés, Vitamin B6 deficient plants display increased sensitivity to high light and photo-oxidative stress, BMC Plant Biol. 9 (2009) 130. [41] C. Triantaphylidès, M. Havaux, Singlet oxygen in plants: production, detoxification and signaling, Trends Plant. Sci. 14 (2009) 219–228.
SC
[42] R.G.L. op den Camp, D. Przybyla, C. Ochsenbein, C. Laloi, C. Kim, A. Danon, D. Wagner, É. Hideg, C. Göbel, I. Feussner, M. Nater, K. Apel, Rapid induction of distinct stress responses after the release of singlet oxygen in Arabidopsis, Plant Cell 15 (2003) 2320‒2332.
NU
[43] A. Danon, N.S. Coll, K. Apel, Cryptochrome-1-dependent execution of programmed cell death induced by singlet oxygen in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 103 (2006) 17036‒17041.
MA
[44] K.P. Lee, C. Kim, F. Landgraf, K. Apel, EXECUTER1- and EXECUTER2-dependent transfer of stress-related signals from the plastid to the nucleus of Arabidopsis thaliana, Proc. Nat. Acad. Sci. USA 104 (2007) 10270–10275.
ED
[45] P.R. Ogilby, C.S. Foote, Chemistry of singlet oxygen. 42. Effect of solvent, solvent isotopic substitution and temperature on the lifetime of singlet molecular oxygen (1Δg), J. Am. Chem. Soc. 105 (1983) 3423–3430.
PT
[46] J.B. Arellano, Y.A. Yousef, T.B. Melø, S.B. Mohamad, R.J. Cogdell, K.R. Naqvi KR, Formation and geminate quenching of singlet oxygen in purple bacterial reaction center, J. Photochem. Photobiol. B: Biol 87 (2007) 105–112.
AC CE
[47] R. Schmidt, Deactivation of O2(1Δg) singlet oxygen by carotenoids: Internal conversion of excited encounter complexes, J. Phys. Chem. 108 (2004) 5509‒5513. [48] P.F. Conn, W. Schalch, T.G. Truscott, The singlet oxygen and carotenoid interaction, J. Photochem. Photobiol. B: Biol. 11 (1991) 41–47. [49] B.J. Pogson, H.M. Rissler, H.A. Frank, The role of carotenoids in energy quenching, in: T. Wydrzaynski, K. Satoh (Eds.), Photosystem II: The Light-Driven Water-Plastoquinone Oxidoreductase, Springer, The Netherlands , 2005, pp. 515–537. [50] M.J. Thomas, C.S. Foote, Chemistry of singlet oxygen – XXVI. Photooxygenation of phenols, Photochem. Photobiol. 27 (1978) 683–693. [51] E. Baciocchi, T. Del Giacco, O. Lanzalunga, A. Lapi, Singlet oxygen promoted carbon-heteroatom bond cleavage in dibenzyl sulfides and tertiary dibenzylamines. Structural effects and the role of exciplexes, J. Org. Chem. 72 (2007) 9582–9589. [52] G. Jiang, J. Chen, J.S. Huang, C.M. Che, Highly efficient oxidation of amines to imines by singlet oxygen and its application in Ugi-type reactions, Org. Lett. 11 (2009) 4568–4571. [53] Y. Lion, E. Gandin, A. van de Vorst, On the production of nitroxide radicals by singlet oxygen reaction: an EPR study, Photochem. Photobiol. 31 (1980) 305‒309.
ACCEPTED MANUSCRIPT 69
[54] S. Hayashi, H. Yasui, H. Sakurai, Essential role of singlet oxygen species in cytochrome P450-dependent substrate oxygenation by rat liver microsomes, Drug Metab. Pharmacokinet. 20 (2005) 14–23.
RI P
T
[55] C. Triantaphylidès, M. Krischke, F.A. Hoeberichts, B. Ksas, G. Gresser, M. Havaux, F. Van Breusegem, M.J. Mueller, Singlet oxygen is the major reactive oxygen species involved in photooxidative damage to plants, Plant Phys. 148 (2008) 960–968.
SC
[56] M.J. Davies, Reactive species formed on proteins exposed to singlet oxygen. Photochem. Photobiol. Sci. 3 (2004) 17–25. [57] T.P. Devasagayam, A.R. Sundquist, P. Di Mascio, S. Kaiser, H. Sies, Activity of thiols as singlet molecular oxygen quenchers, J. Photochem. Photobiol. B. 9 (1991) 105–116
NU
[58] M. Ehrenshaft, P. Bilski, M.Y. Li, C.F. Chignell, M.E. Daub, A highly conserved sequence is a novel gene involved in the de novo vitamin B6 biosynthesis, Proc. Natl. Acad. Sci. USA 96 (1999) 9374–9378.
MA
[59] G.G. Kramarenko, S.G. Hummel, S.M. Martin, G.R. Buettner, Ascorbate reacts with singlet oxygen to produce hydrogen peroxide, Photochem. Photobiol. 82 (2006) 1634–1637.
ED
[60] S.A. Khorobrykh, M. Karonen, E. Tyystjärvi, Experimental evidence suggesting that H2O2 is produced within the thylakoid membrane in a reaction between plastoquinol and singlet oxygen, FEBS Lett. 589 (2015) 779–86.
PT
[61] P.R. Ogilby, Singlet oxygen: there is indeed something new under the sun, Chem. Soc. Rev. 39 (2010) 3181–3209.
AC CE
[62] M.S. Patterson, S.J. Madsen, B.C. Wilson, Experimental tests of the feasibility of singlet oxygen luminescence monitoring in vivo during photodynamic therapy, J. Photochem. Photobiol. B: Biol. 5 (1990) 69–84. [63] J. Moan, E.O. Pettersen, T. Christensen, The mechanism of photodynamic inactivation of human cells in vitro in the presence of haematoporphyrin, Br. J. Cancer 39 (1979) 398–407. [64] A.A. Krasnovsky Jr., Singlet molecular oxygen in photobiochemical systems: IR phosphorescence studies, Membr. Cell Biol. 12 (1998) 665‒690. [65] H. Li, T.B. Melø, J.B. Arellano, K. Razi Naqvi, Temporal profile of the singlet oxygen emission endogenously produced by photosystem II reaction centre in an aqueous buffer, Photosynth. Res. 112 (2012) 75–79. [66] X. Ragás, X. He, M. Agut, M. Roxo-Rosa, A. Rocha Gonsalves, A.C. Serra, S. Nonell, Singlet oxygen in antimicrobial photodynamic therapy: photosensitizer-dependent production and decay in E. coli, Molecules 18 (2013) 2712–2725. [67] M.K. Kuimova, G. Yahioglu, P.R. Ogilby, Singlet oxygen in a cell: Spatially dependent lifetimes and quenching rate constants, J. Am. Chem. Soc. 131 (2009) 332‒340. [68] P.C. Lee, M.A.J. Rodgers, Singlet molecular oxygen in micellar systems. 1. Distribution equilibria between hydrophobic and hydrophilic compartments, J. Phys. Chem. 87 (1983) 4894–4898.
ACCEPTED MANUSCRIPT 70
[69] B. Ehrenberg, J.L. Anderson, C.S. Foote, Kinetics and yield of singlet oxygen photosensitized by hypericin in organic and biological media, Photochem. Photobiol. 68 (1998) 135–140.
RI P
T
[70] S. Hackbarth, J. Schlothauer, A. Preuss, B. Röder, New insights to primary photodynamic effects – Singlet oxygen kinetics in living cells, J. Photochem. Photobiol. B: Biol. 98 (2010) 173–179.
SC
[71] A. Telfer, T.C. Oldham, D. Phillips, J. Barber, Singlet oxygen formation detected by near-infrared emission from isolated photosystem II reaction centres: direct correlation between P680 triplet decay and luminescence rise kinetics and its consequences for photoinhibition, J. Photochem. Photobiol. B: Biol. 48 (1999) 89–96.
NU
[72] R. Dĕdic, A. Svoboda, J. Pšenčík, L. Lu ínková, J. Komenda, J. Hála, Time and spectral resolved phosphorescence of singlet oxygen and pigments in Photosystem II particles, J. Lumin. 102–103 (2003) 313–317.
MA
[73] T. Tomo, H. Kusakabe, R. Nagao, H. Ito, A. Tanaka, S. Akimoto, M. Mimuro, S. Okazaki, Luminescence of singlet oxygen in photosystem II complexes isolated from cyanobacterium Synechocystis sp. PCC6803 containing monovinyl or divinyl chlorophyll a, Biochim. Biophys. Acta 1817 (2012) 1299–1305.
ED
[74] T.K. Antal, I.B. Kovalenko, A.B. Rubin, E. Tyystjärvi, Photosynthesis-related quantities for education and modeling, Photosynth. Res. 117 (2013) 1‒30.
PT
[75] J. Moan, T. Christensen, Photodynamic effects on human cells exposed to light in the presence of hematoporphyrin. Localization of the active dye, Cancer Lett. 11 (1981) 209–214.
AC CE
[76] J. Moan, On the diffusion length of singlet oxygen in cells and tissues, J. Photochem. Photobiol. B. 6 (1990) 343–347. [77] G.N. Bosio, T. Breitenbach, J. Parisi, M. Reigosa, F.H. Blaikie, B.W. Pedersen, E.F.F. Silva, D.O. Mártire, P.R. Ogilby, Antioxidant β-carotene does not quench singlet oxygen in mammalian cells. J. Am. Chem. Soc. 135 (2013) 272–279. [78] P. Han, D.M. Bartels, Temperature dependence of oxygen diffusion in H2O and D2O, J. Phys. Chem. 100 (1996) 5597–5602. [79] J. Baier, M. Maier, R. Engl, M. Landthaler, W. Bäumler, Time-resolved investigations of singlet oxygen luminescence in water, in phosphatidylcholine, and in aqueous suspensions of phosphatidylcholine or HT29 cells, J. Phys. Chem. B 109 (2005) 3041–3046. [80] E. Skovsen, J.W. Snyder, J.D.C. Lambert, P.R. Ogilby, Lifetime and diffusion of singlet oxygen in a cell, J. Physical Chem. B 109 (2005) 8570–8573. [81] J.W. Snyder, E. Skovsen, J.D.C. Lambert, L. Poulsen, P.R. Ogilby, Optical detection of singlet oxygen from single cells, Phys. Chem. Chem. Phys. 8 (2006) 4280–4293. [82] M. Niedre, M.S. Patterson, B.C. Wilson, Direct near-infrared luminescence detection of singlet oxygen generated by photodynamic therapy in cells in vitro and tissues in vivo. Photochem. Photobiol. 75 (2002) 382–391. [83] I.S. Zulfugarov, A. Tovuu, J.H. Kim, C.H. Lee, Detection of reactive oxygen species in higher plants, J. Plant Biol. 54 (2011) 351–357.
ACCEPTED MANUSCRIPT 71
RI P
T
[84] J.R. Kanofsky, Measurement of singlet-oxygen in vivo: Progress and pitfalls, Photochem. Photobiol. 87 (2011) 14–17. [85] A. Telfer, Singlet oxygen production by PSII under light stress: Mechanism, -carotene, Plant Cell Physiol. 55 (2014) 1216– 1223. [86] W.M. Irvine, J.B. Pollack, Infrared optical properties of water and ice spheres, Icarus 8 (1968) 324–360.
SC
[87] J.G. Bayly, V.B. Kartha, W.H. Stevens, The absorption spectra of liquid phase H2O, HDO and D2O from 0.7 μm to 10 μm, Infrared Phys. 3 (1963) 211–223.
NU
[88] A.N. Macpherson, A. Telfer, J. Barber, T.G. Truscott, Direct detection of singlet oxygen from isolated Photosystem II reaction centres, Biochim. Biophys. Acta, 1143 (1993) 301–309.
MA
[89] Y. Fu, A.A. Krasnovsky Jr., C.S. Foote, Singlet oxygen dimol-sensitized luminescence from thermally generated singlet oxygen, J. Am. Chem. Soc. 115 (1993) 10282‒10285. [90] R.P. Wayne, Singlet oxygen in the environmental sciences, Res. Chem. Intermed. 20 (1994) 395‒422.
PT
ED
[91] L.Y. Zang, F.J.G.M. van Kuijk, B.R. Misra, H.P. Misra, The specificity and product of quenching singlet oxygen by 2, 2, 6, 6-tetramethylpiperidine, Biochem. Mol. Biol. Int. 37 (1995) 283–293.
AC CE
[92] M. Hakala-Yatkin, E. Tyystjärvi, Inhibition of Photosystem II by the singlet oxygen sensor compounds TEMP and TEMPD, Biochim. Biophys. Acta 1807 (2011) 243–250. [93] É. Hideg, Z. Deák, M. Hakala-Yatkin, M. Karonen, AW. Rutherford, E. Tyystjärvi, I. Vass, A. Krieger-Liszkay, Pure forms of the singlet oxygen sensors TEMP and TEMPD do not inhibit Photosystem II, Biochim. Biophys. Acta 1807 (2011) 1658–1661. [94] E. Blumwald, R.J. Mehlhorn, L. Packer, Studies of osmoregulation in salt adaptation of cyanobacteria with ESR spin-probe techniques, Proc. Nat. Acad. Sci. USA 80 (1983) 2599–2602. [95] Y. Nishiyama, S.I. Allakhverdiev, H. Yamamoto, H. Hayashi, N. Murata, Singlet oxygen inhibits the repair of Photosystem II by suppressing the translation elongation of the D1 protein in Synechocystis sp. PCC 6803, Biochem. 43 (2004) 11321–11330. [96] É. Hideg, C. Spetea, I. Vass, Singlet oxygen and free radical production during acceptor- and donor-side-induced photoinhibition. Studies with spin trapping EPR s ectrosco y, Biochim. Bio hys. Acta 1186 (1994) 143−152. [97] É, Hideg, C. Spetea, I. Vass, Singlet oxygen production in thylakoid membranes during photoinhibition as detected by EPR spectroscopy, Photosynth. Res. 39 (1994) 191–199. [98] M. Karonen, H. Mattila, P. Huang, F. Mamedov, S. Styring, E. Tyystjärvi, A tandem mass spectrometric method for singlet oxygen measurement, Photochem. Photobiol. 90 (2014) 965–971.
ACCEPTED MANUSCRIPT 72
[99] J. Tandori, É. Hideg, L. Nagy, P. Maróti, I. Vass, Photoinhibition of carotenoidless reaction centers from Rhodobacter sphaeroides by visible light. Effects on protein structure and electron transport, Photosynth. Res. 70 (2001) 175–184.
RI P
T
[100] B.B. Fischer, A. Krieger-Liszkay, É. Hideg, I. Snyrychova, M. Wiesendanger, R.I. Eggen, Role of singlet oxygen in chloroplast to nucleus retrograde signaling in Chlamydomonas reinhardtii, FEBS Lett. 581 (2007) 5555–5560.
MA
NU
SC
[101] É. Hideg, T. Kálai, K. Hideg, Direct detection of free radicals and reactive oxygen species in thylakoids, in: R. Carpentier (Ed.), Photosynthesis Research Protocols, Second Edition, Humana Press, New York, 2011, pp. 187–200. [102] H.S. Ryang, C.S. Foote, Chemistry of singlet oxygen. 31. Low-temperature nuclear magnetic resonance studies of dye-sensitized photooxygenation of imidazoles: Direct observation of unstable 2,5-endoperoxide intermediates, J. Am. Chem. Soc. 101 (1979) 6683‒6687. [103] R.V. Bensasson, J. Frederiksen, M. Rougée, D. Lexa, N. Harrit, Correlations between the rate constant of singlet oxygen quenching by imidazole derivatives and anti-inflammatory activity in rats, Mol. Pharmacol. 42 (1992) 718‒722.
ED
[104] I. Kraljid, S. El Moshni, New method for detection of singlet oxygen in aqueous solutions, Photochem. Photobiol. 28 (1978) 577‒581.
AC CE
PT
[105] S.N. Roudyk, A. Moxhet, R.F. Matagne, J. Aghion, Evidence of singlet oxygen evolution by whole living cells of Chlamydomonas reinhardtii, Photosynth. Res. 47 (1996) 99–102. [106] D. Pastore, D. Trono, L. Padalino, N. Di Fonzo, S. Passarella, pNitrosodimethylaniline (RNO) bleaching by soybean lipoxygenase-1. Biochemical characterization and coupling with oxodiene formation, Plant Physiol. Biochem. 38 (2000) 845‒852. [107] J. Jung, H.S. Kim, The chromophores as endogenous sensitizers involved in the photogeneration of singlet oxygen in spinach thylakoids, Photochem. Photobiol. 52 (1990) 1003–1009. [108] A. Stemler, P. Jursinic, The Effects of Carbonic Anhydrase Inhibitors Formate, Bicarbonate, Acetazolamide, and imidazole on Photosystem II in maize chloroplasts, Arch. Biochem. Biophys. 221 (1983) 227–237. [109] É. Hideg, A comparative study of fluorescent singlet oxygen probes in plant leaves, Cent. Eur. J. Biol. 3 (2008) 273–284. [110] A. Gollmer, J. Arnbjerg, F.H. Blaikie, B.W. Pedersen, T. Breitenbach, K. Daasbjerg, M. Glasius, P.R. Ogilby, Singlet Oxygen Sensor Green®: Photochemical behavior in solution and in a mammalian cell, Photochem. Photobiol. 87 (2011) 6711– 679.
[111] M. Rác, M. Sedlářová, P. Pos íšil, The formation of electronically excited s ecies in the human multiple myeloma cell suspension, Sci. Rep. 5 (2015) 8882. [112] X. Ragás, A. Jiménez-Banzo, D. Sánchez-García, X. Batllori, S. Nonell, Singlet oxygen photosensitisation by the fluorescent probe Singlet Oxygen Sensor Green®, Chem. Commun. 20 (2009) 2920–2922.
ACCEPTED MANUSCRIPT 73
[113] L. Dall’Osto, A. Fiore, S. Cazzaniga, G. Giuliano, R. Bassi, Different roles of α- and β-branch xanthophylls in photosystem assembly and photoprotection, J. Biol. Chem. 282 (2007) 35056–35068.
RI P
T
[114] R.K. Sinha, J. Komenda, J. Kno ová, M. Sedlářová, P. Pos íšil, Small CAB-like proteins prevent formation of singlet oxygen in the damaged Photosystem II complex of the cyanobacterium Synechocystis sp. PCC 6803, Plant, Cell Environ. 35 (2012) 806– 818.
SC
[115] L. Dall’Osto, N.E. Holt, S. Kaligotla, M. Fuciman, S. Cazzaniga, D. Carbonera, H.A. Frank, J. Alric, R. Bassi, Zeaxanthin protects plant photosynthesis by modulating chlorophyll triplet yield in specific light-harvesting antenna subunits, J .Biol. Chem. 287 (2012) 41820–41834.
NU
[116] T. Kálai, É. Hideg, I. Vass, K. Hideg, Double (fluorescent and spin) sensors for detection of reactive oxygen species in the thylakoid membrane, Free Radic. Biol. Med. 24 (1998) 649–652.
MA
[117] É Hideg, T. Kálai, P.B. Kós, K. Asada, K. Hideg, Singlet oxygen in plants—its significance and possible detection with double (fluorescent and spin) indicator reagents, Photochem. Photobiol. 82 (2006) 1211–1218.
ED
[118] É. Hideg, T. Kálai, K. Hideg, I. Vass, Photoinhibition of photosynthesis in vivo results in singlet oxygen production detection via nitroxide-induced fluorescence quenching in broad bean leaves, Biochemistry 37 (1998) 11405–11411.
PT
[119] É. Hideg, K. Ogawa, T. Kálai, K. Hideg, Singlet oxygen imaging in Arabidopsis thaliana leaves under photoinhibition by excess photosynthetically active radiation, Physiol. Plant. 112 (2001) 10–14.
AC CE
[120] É. Hideg, C. Barta, T. Kálai, I. Vass, K. Hideg, K. Asada, Detection of singlet oxygen and superoxide with fluorescent sensors in leaves under stress by photoinhibition or UV radiation, Plant Cell Physiol. 43 (2002) 1154‒1164. [121] G. Agati, P. Matteini, A. Goti, M. Tattini, Chloroplast-located flavonoids can scavenge singlet oxygen, New Phytol. 174 (2007) 77‒89. [122] J. Glaeser, G. Klug, Photo-oxidative stress in Rhodobacter sphaeroides: protective role of carotenoids and expression of selected genes, Microbiology (2005) 151, 1927–1938. [123] N. Umezawa, K. Tanaka, Y. Urano, K. Kikuchi, T. Higuchi, T. Nagano, Novel fluorescent probes for singlet oxygen, Angew. Chem. Int. Ed. Engl. 38 (1999) 2899– 2901. [124] K. Tanaka, T. Miura, N. Umezawa, Y. Urano, K. Kikuchi, T. Higuchi, T. Nagano, Rational design of fluorescein-based fluorescence probes. Mechanism-based design of a maximum fluorescence probe for singlet oxygen, J. Am. Chem. Soc. 123 (2001) 2530–2536. [125] E. Yamamoto, S. Munne-Bosch, Y. Urano, K. Asada, Detection of singlet oxygen by using fluorescence probe DPAX in Photosystem, Plant Cell Physiol. 45 (2004) S194– S194. [126] R.H. Bisby, C.G. Morgan, I. Hamblett, A.A. Gorman, Quenching of singlet oxygen by Trolox C, ascorbate, and amino acids: Effects of pH and temperature, J. Phys. Chem. A 103 (1999) 7454‒7459.
ACCEPTED MANUSCRIPT 74
[127] A.M. Wade, H.N. Tucker, Antioxidant characteristics of L-histidine, J. Nutr. Biochem. 9 (1998) 308‒315.
RI P
T
[128] A. Telfer, S. M. Bishop, D. Phillips, J. Barber, Isolated photosynthetic reaction center of Photosystem II as a sensitizer for the formation of singlet oxygen. Detection and quantum yield determination using a chemical trapping technique, J. Biol. Chem. 269 (1994) 13244–13253.
MA
NU
SC
[129] L. Bersanini, N. Battchikova, M. Jokel, A. Rehman, I. Vass, Y. Allahverdiyeva, E.M. Aro, Flavodiiron protein Flv2/Flv4-related photoprotective mechanism dissipates excitation pressure of PSII in cooperation with phycobilisomes in cyanobacteria, Plant Physiol. 164 (2014) 805‒818. [130] K, Hakkila, T. Antal, A.U. Rehman, J. Kurkela, H. Wada, I. Vass, E. Tyystjärvi, T. Tyystjärvi, Oxidative stress and photoinhibition can be separated in the cyanobacterium Synechocystis sp. PCC 6803, Biochim. Biophys. Acta 1837 (2014) 217– 225. [131] X. Shu, A. Royant, M.Z. Lin, T.A. Aguilera, V. Lev-Ram, P.A. Steinbach, R.Y. Tsien, Mammalian expression of infrared fluorescent proteins engineered from a bacterial phytochrome, Science 324 (2009) 804–807.
ED
[132] T.-L. To, M.J. Fadul, X. Shu, Singlet oxygen triplet energy transfer-based imaging technology for mapping protein-protein procimity in intact cells, Nature Comm. 5 (2014) 4072.
AC CE
PT
[133] F.M. Pimenta, J.K. Jensen, M. Etzerodt, P.R. Ogilby, Protein-encapsulated bilirubin: paving the way to a useful probe for singlet oxygen, Photochem. Photobiol. Sci. 14 (2015) 665–677.
[134] C. Kim, K. Apel, Singlet oxygen-mediated signaling in plants: moving from flu to wild type reveals an increasing complexity, Photosynth. Res. 116 (2013) 455–464. [135] J.R. Anthony, K.L. Warczak, T.J. Donohue, A transcriptional response to singlet oxygen, a toxic byproduct of photosynthesis, Proc. Natl. Acad. Sci. USA 102 (2005) 6502–6507. [136] I. Gadjev, S. Vanderauwera, T.S. Gechev, C. Laloi, I.N. Minkov, V. Shulaev, K. Apel, D. Inzé, R. Mittler, F. Van Breusegem, Transcriptomic footprints disclose specificity of reactive oxygen species signaling in Arabidopsis, Plant Physiol. 141 (2006) 436–445. [137] F. Ramel, S. Birtic, S. Cuiné, C. Triantaphylidès, J.L. Ravanat, M. Havaux, Chemical quenching of singlet oxygen by carotenoids in plants, Plant Physiology 158 (2012) 1267–1278. [138] U. Leisinger, K. Rüfenacht, B. Fischer, M. Pesaro, A. Spengler, A.J.B. Zehnder, R.I.L Eggen, The glutathione peroxidase homologous gene from Chlamydomonas reinhardtii is transcriptionally upregulated by singlet oxygen, Plant Mol. Biol. 46 (2001) 395–408.
ACCEPTED MANUSCRIPT 75
[139] A. Baruah, K. Šimkova, K. A el, C. Laloi, Arabido sis mutants reveal multi le singlet oxygen signaling pathways involved in stress response and development, Plant Mol. Biol. 70 (2009) 547–563.
T
[140] M. Havaux, Carotenoid oxidation products as stress signals in plants, Plant J. 79 (2014) 597–606.
RI P
[141] B. Nowicka, J. Kruk, Plastoquinol is more active than α-tocopherol in singlet oxygen scavenging during high light stress of Chlamydomonas reinhardtii, Biochim. Biophys. Acta 1817 (2012) 389–394.
SC
[142] N. Kobayashi, D. DellaPenna, Tocopherol metabolism, oxidation and recycling under high light stress in Arabidopsis, Plant J. 55 (2008) 607–618.
NU
[143] R.H. Young, D. Brewer, R.A. Keller, The determination of rate constants of reaction and lifetimes of singlet oxygen in solution by a flash photolysis technique, J. Am. Chem. Soc. 95 (1973) 375–379. [144] P. Cioni, G.B. Strambini, Effect of heavy water on protein flexibility, Biophys. J. 82 (2002) 3246–3253.
MA
[145] G. Renger, H.J. Eckert, A. Bergmann, J. Bernarding, B. Liu, A. Napiwotzki, F. Reifarth, H.J. Eichler, Fluorescence and spectroscopic studies of exciton trapping and electron transfer in Photosystem II of higher plants, Aust. J. Plant Physiol. 22 (1995) 167–181.
PT
ED
[146] R. Pratt, F.N. Craig, S.F. Trelease, Influence of deuterium oxide on photochemical and dark reactions of photosynthesis, Science 12 (1937) 271–273. [147] É. Hideg, P.B. Kós, I. Vass, Photosystem II damage induced by chemically generated singlet oxygen in tobacco leaves, Physiol. Plant 131 (2007) 33–40.
AC CE
[148] L. Kovács, F. Ayaydin, T. Kálai, J. Tandori, P.B. Kós, É. Hideg, Assessing the applicability of singlet oxygen photosensitizers in leaf studies, Photochem. Photobiol. 90 (2014) 129–136. [149] B. Barényi, G.H. Krause, Inhibition of photosynthetic reactions by light, A study with isolated spinach chloroplasts, Planta 163 (1985) 218–226. [150] M. Hakala-Yatkin, P. Sarvikas, P. Paturi, M. Mäntysaari, H. Mattila, T. Tyystjärvi, L. Nedbal, E. Tyystjärvi, Magnetic field protects plants against high light by slowing down production of singlet oxygen, Physiol. Plant. 142 (2011) 26‒34. [151] S. Inoue, K. Ejima, E. Iwai, H. Hayashi, J. Appel, E. Tyystjärvi, N. Murata, Y. Nishiyama, Protection by -tocopherol of the repair of photosystem II during photoinhibition in Synechocystis sp. PCC 6803, Biochim. Biophys. Acta 1807 (2011) 236‒241. [152] J.M. Burns, W.J. Cooper, J.L. Ferry, D.W. King, B.P. DiMento, K. McNeill, C.J. Miller, W.L. Miller, B.M. Peake, S.A. Rusak, A.L. Rose, T.D. Waite, Methods for reactive oxygen species (ROS) detection in aqueous environments, Aquat. Sci. 74 (2012) 683– 734. [153] T. Kálai, E. Hideg, F. Ayaydin, Synthesis and potential use of 1,8-naphthalimide type 1O2 sensor molecules, Photochem. Photobiol. Sci. 12 (2013) 432–438. [154] D.R. Lide (Ed.) Handbook of Chemistry and Physics, 84th Edition, CRC Press, Boca Raton, 2003.
ACCEPTED MANUSCRIPT 76
[155] A. Dinçer, T. Aydemir, Purification and characterization of catalase from chard (Beta vulgaris var. cicla), J. Enzyme Inhib. 16 (2001) 165–175.
T
[156] E.A. Havir, N.A. McHale, Purification and characterization of an isozyme of catalase with enhanced-peroxidatic activity from leaves of Nicotiana sylvestris, Arch. Biochem. Biophys. 283 (1990) 491–495.
RI P
[157] G. Arabaci, A. Usluoglu, Catalytic properties and immobilization studies of catalase from Malva sylvestris L., J. Chem. 2013 (2013) Art. 686185.
SC
[158] M.A. Hoque, M. Uraji, A. Torii, M.N. Banu, I.C. Mori, Y. Nakamura, Y. Murata, Methylglyoxal inhibition of cytosolic ascorbate peroxidase from Nicotiana tabacum, J. Biochem. Mol. Toxicol. 26 (2012) 315–321.
NU
[159] M. Drábková, H.C.P. Matthijs, W. Admiraal, B. Maršálek, Selective effects of H2O2 on cyanobacterial photosynthesis, Photosynthetica 45 (2007) 363‒369.
MA
[160] H.C. Matthijs, P.M. Visser, B. Reeze, J. Meeuse, P.C. Slot, G. Wijn, R. Talens, J. Huisman, Selective suppression of harmful cyanobacteria in an entire lake with hydrogen peroxide, Water Res. 46 (2012) 1460‒1472. [161] T.S. Gechev, F. Van Breugesem, J.M. Stone, I. Denev, C. Laloi, Reactive oxygen species as signals that modulate plant stress responses and programmed cell death, Bioessays 28 (2006) 1091‒1101.
ED
[162] C.C. Winterbourn, D. Metodiewa, Reactivity of biologically important thiol compounds with superoxide and hydrogen peroxide, Free Radic. Biol. Med, 27 (1999) 322–328.
PT
[163] W.M. Kaiser, Reversible inhibition of the Calvin cycle and activation of oxidative pentose phosphate cycle in isolated intact chloroplasts by hydrogen peroxide, Planta, 145 (1979) 377–382.
AC CE
[164] F. Verniquet, J. Gaillard, M. Neuburger, R. Douce, Rapid inactivation of plant aconitase by hydrogen peroxide, Biochem J. 276 (1991) 643–648. [165] M. Meinhard, P.L. Rodriguez, E. Grill, The sensitivity of AB12 to hydrogen peroxide links the abscisic acid-response regulator to redox signaling, Planta 214 (2002) 775–782. [166] C. Waszczak, S. Akter, D. Eeckhout, G. Persiau, K. Wahni, N. Bodra, I. Van Molle, B. De Smet, D. Vertommen, K. Gevaert, G. De Jaeger, M. Van Montagu, J. Messens, F. Van Breusegem, Sulfenome mining in Arabidopsis thaliana, Proc. Natl. Acad. Sci. USA 111 (2014) 11545–11550. [167] M. Zheng, F. Åslund, G. Storz, Activation of the OxyR transcription factor by reversible disulfide bond formation, Science 279 (1998) 1718–1721. [168] O. Augusto, S. Miyamoto, Oxygen radicals and related species. In: K. Pantopoulos, H.M. Schipper (Eds.), Principles of Free Radical Biomedicine, Nova Science Publishers, Montreal, 2012, pp. 19–42. [169] J. Weinstein, B.H.J. Bielski, Kinetics of the interaction of HO2 and O2 radicals with hydrogen peroxide; the Haber-Weiss reaction, J. Am. Chem. Soc. 101 (1979) 58–62. [170] W.H. Melhuish, H.C. Sutton, Study of the Haber-Weiss reaction using a sensitive method for detection of OH radicals, J. Chem. Soc. Chem. Commun. 22 (1978) 970– 971.
ACCEPTED MANUSCRIPT 77
[171] Y. Mizuta, T. Masumizu, M. Kohno, A. Mori, L. Packer, Kinetic analysis of the Fenton reaction by ESR-spin trapping, Biochem. Mol. Biol. Int. 43 (1997) 1107–1120.
RI P
T
[172] M. Masarwa, H. Cohen, D. Meyerstein, D.L. Hickman, A. Bakac, J.H. Espenson, Reactions of low valent transition metal complexes with hydrogen peroxide. Are they "Fenton-like" or not? 1. The case of Cu+aq and Cr2+aq, J. Am. Chem. Soc. 110 (1988) 4293–4297.
SC
[173] J.W. Moffett, R.G. Zika, Reaction kinetics of hydrogen peroxide with copper and iron in seawater, Environ. Sci. Technol. 21 (1987) 804–810. [174] P. Salgado, V. Melin, D. Contreras, Y. Moreno, H.D. Mansilla, Fenton reaction driven by iron ligands, J. Chil. Chem. Soc. 58 (2013) 2096–2101.
NU
[175] F. Deyhimi, F. Nami, Peroxidase-catalyzed electrochemical assay of hydrogen peroxide: A ping–pong mechanism, Int. J. Chem. Kinet. 44 (2012) 699–704.
MA
[176] C. Miayke, F. Michihata, K. Asada, Scavenging of hydrogen peroxide in prokaryotic and eukaryotic algae: Acquisition of ascorbate peroxidase during the evolution of cyanobacteria, Plant Cell Physiol. 32 (1991) 33–43.
ED
[177] K. Asada, The water-water cycle as alternative photon and electron sinks, Phil. Trans. R. Soc. Lond. B. Biol. Sci. 355 (2000) 1419–1431.
PT
[178] K. Asada, K. Yoshikawa, M. Takahashi, Y. Maeda, K. Enmanji, Superoxide dismutase from a blue-green alga, Plectoneme boryanum, J. Biol. Chem. 250 (1975) 2801–2807.
AC CE
[179] J. Weller, K.M. Kizina, K. Can, G. Bao, M. Müller, Response properties of the genetically encoded optical H2O2 sensor HyPer, Free Rad. Biol. Med. 76 (2014) 227– 241. [180] A. Costa, I. Drago, S. Behera, M. Zottini, P. Pizzo, J.I. Schroeder, T. Pozzan, F. Lo Schiavo, H2O2 in plant peroxisomes: an in vivo analysis uncovers a Ca2+-dependent scavenging system, Plant J. 62 (2010) 760–772. [181] H. Thordal-Christensen, Z. Zhang, Y. Wei, D.B. Collinge, Subcellular localization of H2O2 in plants. H2O2 accumulation in papillae and hypersensitive response during the barley-powdery mildew interaction, Plant J. 11 (1997) 1187–1194. [182] A. Daudi, Z. Cheng, J.A. O'Brien, N. Mammarella, S. Khan, F.M. Ausubel, G.P. Bolwell, The apoplastic oxidative burst peroxidase in Arabidopsis is a major component of pattern-triggered immunity, Plant Cell 24 (2012) 275–287. [183] A. Daudi, J.A. O’Brien, Detection of hydrogen peroxide by DAB staining in Arabidopsis leaves, Bio-protocol 2 (2012) e263. http://www.bio-protocol.org/e263 [184] I. Šnyrychová, F. Ayaydin, É. Hideg, Detecting hydrogen eroxide in leaves in vivo – a comparison of methods, Physiol. Plant. 135 (2009) 1–18. [185] M.J. Fryer, L. Ball, K. Oxborough, S. Karpinski, P.M. Mullineaux, N.R. Baker, Control of Ascorbate Peroxidase 2 expression by hydrogen peroxide and leaf water status during excess light stress reveals a functional organization of Arabidopsis leaves, Plant J. 33 (2003) 691‒705.
ACCEPTED MANUSCRIPT 78
T
[186] C.S. Bestwick, I.R. Brown, M.H.R. Bennet, J.M. Mansfield, Localization of hydrogen peroxide accumulation during the hypersensitive reaction of lettuce cells to Pseudomonas syringae pv. phaseolicola, Plant Cell 9 (1997) 209–221.
RI P
[187] O.B. Blokhina, T.V. Chirkova, K.V. Fagerstedt, Anoxic stress leads to hydrogen peroxide formation in plant cells, J. Exp. Bot. 52 (2001) 1179–1190.
NU
SC
[188] L. Liu, K.E.L. Eriksson, J.F.D. Dean, Localization of hydrogen peroxide production in Zinnia elegans L. stems, Phytochemistry 52 (1999) 545–554.
MA
[189] T.T. Ngo, H.M. Lenhoff, A sensitive and versatile chromogenic assay for peroxidase and peroxidase-coupled reactions, Anal. Biochem. 105 (1980) 389–397.
ED
[190] S. Veljovid-Jovanovid, G. Noctor, C.H. Foyer, Are leaf hydrogen eroxide concentrations commonly overestimated? The potential influence of artefactual interference by tissue phenolics and ascorbate, Plant Physiol. Biochem. 40 (2002) 501–507.
PT
[191] I. Garrido, F. Espinosa, M. Carmen Álvarez-Tinaut, Oxidative defence reactions in sunflower roots induced by methyl-jasmonate and methyl-salicylate and their relation with calcium signalling, Protoplasma 237 (2009) 27‒39.
AC CE
[192] J. Frew, P. Jones, G. Scholes, Spectrophotometric determination of hydrogen peroxide and organic hydroperoxides at low concentrations in aqueous solution, Anal. Chim. Acta 155 (1983) 139–150.
[193] J.A. Hernández, E. Olmos, F.J. Corpas, F. Sevilla, L.A. del Río, Salt-induced oxidative stress in chloroplasts of pea plants, Plant Sci. 105 (1995) 151–167.
[194] J. Manai, H. Gouia, F.J. Corpas, Redox and nitric oxide homeostasis are affected in tomato (Solanum lycopersicum) roots under salinity-induced oxidative stress, J. Plant Physiol. 171 (2014) 1028‒1035.
[195] P.D. Olson, J.E. Varner, Hydrogen peroxide and lignification, Plant J. 4 (1993) 887–892.
[196] R. Kaur-Sawhney, H.E. Flores, A.W. Galston, Polyamine oxidase in oat leaves: a cell wall-localized enzyme, Plant Physiol. 68 (1981) 494–498.
ACCEPTED MANUSCRIPT 79
T
[197] S. Maruthasalam, Y.L. Liu, C.M. Sun, P.Y. Chen, C.W. Yu, P.F. Lee, C.H. Lin, Constitutive expression of a fungal glucose oxidase gene in transgenic tobacco confers chilling tolerance through the activation of antioxidative defence system, Plant Cell Rep. 29 (2010) 1035–1048.
RI P
[198] M. Hicks, J.M. Gebicki, Spectrophotometric method for the determination of lipid hydroperoxides, Anal. Biochem. 99 (1979) 249‒253.
SC
[199] P. Van Gestelen, P. Ledeganck, I. Wynant, R.J. Caubergs, H. Asard, The cantharidin-induced oxidative burst in tobacco BY-2 cell suspension cultures, Protoplasma 205 (1998) 83–92.
NU
[200] J.T. Corbett, The scopoletin assay for hydrogen peroxide. A review and a better method, J. Biochem. Biophys. Methods 18 (1989) 297–308.
MA
[201] M. Hakala-Yatkin, M. Mäntysaari, H. Mattila, E. Tyystjärvi, Contributions of visible and ultraviolet parts of sunlight to photoinhibition, Plant Cell Physiol. 51 (2010) 1745‒1753.
ED
[202] N. Liu, Z. Lin, H. Mo, Metal (Pb, Cd, and Cu)-induced reactive oxygen species accumulations in aerial root cells of the Chinese banyan (Ficus microcarpa), Ecotoxicology 21 (2012) 2004–2011.
PT
[203] K. Bóka, N. Orbán, Z. Kristóf, Dynamics and localization of H2O2 production in elicited plant cells, Protoplasma 230 (2007) 89–97.
AC CE
[204] K.S. Gould, J. McKelvie, K.R. Markham, Do anthocyanins function as antioxidants in leaves? Imaging of H2O2 in red and green leaves after mechanical injury, Plant Cell Env. 25 (2002) 1261–1269. [205] O.S. Wolfbeis, A. Dürkop, M. Wu, Z. Lin, A europium-ion-based luminescent sensing probe for hydrogen peroxide, Angew. Chem. Int. Ed. Engl. 41 (2002) 4495– 4498. [206] Y.R. Abdrakhimova, I.M. Andreev, A.G. Shugaev, Determination of ROS generation rates in plant mitochondria in vitro using fluorescent indicators: nonspecific effects of inhibitors of terminal oxidases, Russ. J. Plant Physiol. 62 (2015) 136– 141. [207] T.K. Antal, P. Sarvikas, E. Tyystjärvi, Two-electron reactions S2QB → S0QB and S3QB → S1QB are involved in deactivation of higher S states of the oxygen-evolving complex of Photosystem II, Biophys. J. 96 (2009) 4672–4680. [208] J. Yuan, A.M. Shiller, Determination of subnanomolar levels of hydrogen peroxide in seawater by reagent-injection chemiluminescence detection, Anal. Chem. 71 (1999) 1975–1980. [209] A. Daiber, M. Oelze, M. August, M. Wendt, K. Sydow, H. Wieboldt, A.L.Klschyov, T. Munzel, Detection of superoxide and peroxynitrite in model systems and mitochondria by the luminol analogue L-012, Free Radic. Res. 38 (2004) 259‒269. [210] F.J. Pérez, S. Rubio, An improved chemiluminescence method for hydrogen peroxide determination in plant tissues, Plant Growth Regul. 48 (2006) 89–95.
ACCEPTED MANUSCRIPT 80
T
[211] M.O. Vogel, M. Moore, K. König, P. Pecher, K. Alsharafa, J. Lee, K.J. Dietz, Fast retrograde signaling in response to high light involves metabolite export, MITOGENACTIVATED PROTEIN KINASE6, and AP2/ERF transcription factors in Arabidopsis, Plant Cell 26 (2014) 1151‒1165.
RI P
[212] R. Vergara, F. Parada, S. Rubio, F.J. Pérez, Hypoxia induces H2O2 production and activates antioxidant defence systems in grapevine buds through mediation of H2O2 and ethylene, J. Exp. Bot. 63 (2012) 4123‒4131.
SC
[213] L. Zhang, J. Wang, Y. Tian, Electrochemical in-vivo sensors using nanomaterials made from carbon species, noble metals, or semiconductors, Microchim. Acta 181 (2014) 1471–1484.
NU
[214] Q.Q. Ren, X.J. Yuan, X.R. Huang, W. Wen, Y.D. Zhao, W. Chen, In vivo monitoring of oxidative burst on aloe under salinity stress using hemoglobin and single-walled carbon nanotubes modified carbon fiber ultramicroelectrode, Biosens. Bioelectron. 50 (2013) 318–324.
MA
[215] Q. Xu, F. Wei, Z. Wang, Q. Yang, Y.D. Zhao, H. Chen, In vivo monitor oxidative burst induced by Cd2+ stress for the oilseed rape (Brassica napus L.) based on electrochemical microbiosensor, Phytochem. Anal. 21 (2010) 192–196.
ED
[216] M.I. González-Sánchez, L. González-Macia, M.T. Pérez-Prior, E. Valero, J. Hancock A. J. Killard, Electrochemical detection of extracellular hydrogen peroxide in Arabidopsis thaliana: a real-time marker of oxidative stress, Plant Cell Environ. 36 (2013) 869–878.
AC CE
PT
[217] G.J. Kremers, S.G. Gilbert, P.J. Cranfill, M.W. Davidson, D.W. Piston, Fluorescent proteins at a glance, J. Cell. Sci. 124 (2011) 157–160. [218] S.K. Gjetting, A. Schulz, A.T. Fuglsang, Perspectives for using genetically encoded fluorescent biosensors in plants, Front. Plant Sci. 4 (2013) 234. [219] R. Dixit, R. Cyr, Cell damage and reactive oxygen species production induced by fluorescence microscopy: effect on mitosis and guidelines for non-invasive fluorescence microscopy, Plant J. 36 (2003) 280–290. [220] A.M. Bogdanov, A.S. Mishin, I.V. Yampolsky, V.V. Belousov, D.M. Chudakov, F.V. Subach, V.V. Verkhusha, S. Lukyanov, K.A. Lukyanov, Green fluorescent proteins are light-induced electron donors, Nat. Chem. Biol. 5 (2009) 459–461. [221] A. Hernández-Barrera, A. Velarde-Buendía, I. Zepeda, F. Sanchez, C. Quinto, R. Sánchez-Lopez, A.Y. Cheung, H.-M. Wu, L. Cardenas, Hyper, a hydrogen peroxide sensor, indicates the sensitivity of the arabidopsis root elongation zone to aluminum treatment, Sensors 15 (2015) 855–867. [222] S. Bieker, L. Riester, M. Stahl, J. Franzaring, U. Zentgraf, Senescence-specific alteration of hydrogen peroxide levels in Arabidopsis thaliana and oilseed rape spring variety Brassica napus L. cv. Mozart, J. Integr. Plant Biol. 54 (2012) 540–554. [223] B. Enyedi, M. Zana, Á. Donkó, M. Geiszt, Spatial and temporal analysis of NADPH oxidase-generated hydrogen peroxide signals by novel fluorescent reporter proteins, Antioxid. Redox Signal. 19 (2013) 523–534. [224] P. Lariguet, P. Ranocha, M. De Meyer, O. Barbier, C. Penel, C. Dunand, Identification of a hydrogen peroxide signalling pathway in the control of lightdependent germination in Arabidopsis, Planta 238 (2013) 381–395.
ACCEPTED MANUSCRIPT 81
[225] M. Kwasniewski, K. Chwialkowska, J. Kwasniewska, J. Kusak, K. Siwinski, I. Szarejko, Accumulation of peroxidase-related reactive oxygen species in trichoblasts correlates with root hair initiation in barley, J. Plant Physiol. 170 (2013) 185–195.
RI P
T
[226] K.R. Menon, R. Balan, G.K. Suraishkumar, Stress induced lipid production in Chlorella vulgaris: Relationship with specific intracellular reactive species levels, Biotechnol. Bioeng. 110 (2013) 1627–1636.
SC
[227] K. Setsukinai, Y. Urano, K. Kakinuma, H.J. Majima, T. Nagano, Development of novel fluorescence probes that can reliably detect reactive oxygen species and distinguish specific species, J. Biol. Chem. 278 (2003) 3170–3175.
NU
[228] R. Desikan, S.A.H. Mackerness, J.T. Hancock, S.J. Neill, Regulation of the Arabidopsis transcriptome by oxidative stress, Plant Physiol. 127 (2001) 159–172.
MA
[229] Y. Kanesaki, H. Yamamoto, K. Paithoonrangsarid , M. Shoumskaya, I. Suzuki, H. Hayashi, N. Murata, Histidine kinases play important roles in the perception and signal transduction of hydrogen peroxide in the cyanobacterium, Synechocystis sp. PCC 6803, Plant J. 49 (2007) 313–324. [230] H. Cheng, Q. Zhang, D. Guo, Genes that respond to H2O2 are also evoked under light in Arabidopsis, Mol. Plant 6 (2013) 226–228.
ED
[231] C.J. Baker, N.M. Mock, A method to detect oxidative stress by monitoring changes in the extracellular antioxidant capacity in plant suspension cells, Physiol. Mol. Plant Path. 64 (2004) 255–261.
AC CE
PT
[232] D.T. Sawyer, M.J. Gibian, The chemistry of superoxide ion, Tetrahedron, 35 (1979) 1471–1481. [233] I.B. Afanas’ev, Su eroxide Ion: Chemistry And Biological Im lications, CRC Press, Boca Raton, Florida, USA, 1989. [234] A.A. Frimer, Superoxide chemistry in non-aqueous media, In: M.G Simic, K.A. Taylor, J.F. Ward, C. von Sonntag (eds.), Oxygen Radicals In Biology And Medicine, Plenum Press, New York, 1988, pp. 29–38. [235] Z.V. Todres, Ion-radical organic chemistry: principles and applications, Second Ed. CRC Press,Taylor & Francis Group, Boca Raton, Florida, USA, 2008. [236] K. Asada, The water-water cycle in chloroplasts: scavenging of active oxygens and [237] B.H.J. Bielski, D.E. Cabelli, R.L. Arudi, A.B. Ross, Reactivity of HO2/ O2- radicals in aqueous solution, J. Phys. Chem. Ref. Data 14 (1985) 1041–1100. [238] W.H. Koppenol, Solvation of the superoxide anion, In: G. Gohen, R.A. Greenwald (eds.), Oxy Radicals And Their Scavenger System, Elsevier Biomedical, New York, 1983, pp. 274‒277. [239] P. Pos íšil, The role of metals in roduction and scavenging of reactive oxygen species in Photosystem II, Plant Cell Physiol. 55 (2014) 1224–1232. [240] A. Bafana, S. Dutt, A. Kumar, S. Kumar, P.S Ahuja, The basic and applied aspects of superoxide dismutase, J. Mol. Catal. B: Enzymatic 68 (2011) 129–138. [241] W.C. Danen, R.J. Warner, The remarkable nucleophilicity of superoxide anion radical. Rate constants for reaction of superoxide ion with aliphatic bromides, Tetrahedron Lett. 18 (1977) 989–992.
ACCEPTED MANUSCRIPT 82
[242] A.A. Frimer, The organic chemistry of superoxide anion radical, in: S. Patai (ed.), The Chemistry Of Functional Groups: Peroxides, Wiley, Chichester, 1983, pp. 429–461.
RI P
T
[243] H. Yamamoto, T. Mashino, T. Nagano, M. Hirobe, (1986) Epoxidations of olefins by peroxy intermediate generated in situ from carbon tetrachloride and superoxide, J. Am. Chem. Soc. 108 (1986) 539–541.
SC
[244] S. Dikalov, J. Jiang, R.P. Mason, Characterization of the high-resolution ESR spectra of superoxide radical adducts of 5-(diethoxyphosphoryl)-5-methyl-1-pyrroline N-oxide (DEPMPO) and 5,5-dimethyl-1-pyrroline N-oxide (DMPO), Analysis of conformational exchange, Free Radic. Res. 39 (2005) 825–836.
NU
[245] K. Barnese, E.B. Gralla, J.S. Valentine, D.E. Cabelli, Biologically relevant mechanism for catalytic superoxide removal by simple manganese compounds. Proc. Natl. Acad. Sci. USA 109 (2012) 6892–6897.
MA
[246] B. Gray, A.J. Carmichael, Kinetics of superoxide scavenging by dismutase enzymes and manganese mimics determined by electron spin resonance, Biochem. J. 281 (1992) 795–802. [247] J.P. Kehrer, The Haber–Weiss reaction and mechanisms of toxicity, Toxicology 149 (2000) 43–50.
ED
[248] S.J. Blanksby, V.M. Bierbaum, G.B. Ellison, S. Kato, Superoxide does react with peroxides: direct observation of the Haber-Weiss reaction in the gas phase, Angew. Chem. Int. Ed. Engl. 46 (2007) 4948–4950.
PT
[249] L.A. MacManus-Spencer, B.L. Edhlund, K. McNeill, Singlet oxygen production in the reaction of superoxide with organic peroxides, J. Org. Chem. 71 (2006) 796–799.
AC CE
[250] I.B. Afanas'ev, V.V. Grabovetskii, N.S. Kuprianova, Kinetics and mechanism of the reactions of superoxide ion in solution. Part 5. Kinetics and mechanism of the interaction of superoxide ion with vitamin E and ascorbic acid, J. Chem. Soc., Perkin Trans. 2 (1987) 281–285. [251] H. Wefers, H. Sies, Oxidation of glutathione by the superoxide radical to the disulfide and the sulfonate yielding singlet oxygen, Eur. J. Biochem. 137 (1983) 29–36. [252] C.C. Winterbourn, D. Metodiewa, The reaction of superoxide with reduced glutathione, Arch. Biochem. Biophys. 314 (1994) 284–290. [253] N. Gotoh, E. Niki, Rates of interactions of superoxide with vitamin E, vitamin C and related compounds as measured by chemiluminescence, Biochim. Biophys. Acta 1115 (1992) 201–207. [254] B. Mikami, S. Ida, Reversible inactivation of ferredoxin-nitrate reductase from the cyanobacterium Plectonema boryanum. The role of superoxide anion and cyanide, Plant Cell Physiol. 27 (1986) 1013–1021. [255] N. Shimizu, K. Kobayashi, K. Hayashi, The reaction of superoxide radical with catalase. J. Biol. Chem. 259 (1984) 4414–4418. [256] M.W. Sutherland, The generation of oxygen radicals during host plant responses to infection, Physiol. Mol. Plant Pathol. 39 (1991) 79–93 [257] I. Fridovich, Superoxide radical: an endogenous toxicant, Ann. Rev. Pharmacol. Toxicol. 23 (1983) 239–257.
ACCEPTED MANUSCRIPT 83
[258] A.L. Rose, A. Godrant, M. Furnas, T.D. Waite, Dynamics of nonphotochemical superoxide production and decay in the Great Barrier Reef lagoon, Limnol. Oceanogr. 55 (2010) 1521–1536.
T
[259] M.I. Heller, P.L. Croot, Superoxide decay kinetics in the southern ocean, Environ. Sci. Technol. 44 (2010) 191–196.
RI P
[260] M. Saran, W. Bors, Signalling by O2-• and NO•: how far can either radical, or any specific reaction product, transmit a message under in vivo conditions? Chem. Biol. Interact. 90 (1994) 35–45.
SC
[261] M. Kavdia, A computational model for free radicals transport in the microcirculation, Antioxid. Redox Signal. 8 (2006) 1103–1111.
NU
[262] H.M. Hassan, I. Fridovich, Impermeability of the E. coli cell envelope to O2radicals, in: W.H. Bannister, J.V. Bannister (Eds.), Biological and clinical aspects of superoxide and superoxide dismutase, Elsevier, New York, 1980, pp. 57–61.
MA
[263] B.K. Semin, L.N. Davletshina, K.N. Timofeev, I.I. Ivanov, A.B. Rubin, M. Seibert, Production of reactive oxygen species in decoupled, Ca2+-depleted PSII and their use in assigning a function to chloride on both sides of PSII, Photosynth. Res. 117 (2013) 385–399.
ED
[264] Q. Wang, Y. Hou, J. Miao, G. Li, Effect of UV-B radiation on the growth and antioxidant enzymes of Antarctic sea ice microalgae Chlamydomonas sp. ICE-L, Acta. Physiol. Plant. 31 (2009) 1097–1102.
PT
[265] B.H.J. Bielski, Re-evaluation of spectral and kinetic properties of HO2 and O2free radicals, Photochem. Photobiol. 28 (1978) 645–649.
AC CE
[266] E. Saito, B.H.J. Bielski, Electron paramagnetic resonance spectrum of HO2 radical in aqueous solution, J. Am. Chem. Soc. 83 (1961) 4467–4468. [267] V.C. Runeckles, M. Vaartnou, EPR evidence for superoxide anion formation in leaves during exposure to low levels of ozone, Plant Cell Environ. 20 (1997) 306‒314. [268] B.A. Goodman, T.G. Reichenauer, Formation of paramagnetic products in leaves of wheat (Triticum aestivum) as a result of ozone-induced stress, J. Sci. Food Agric. 83 (2003) 1248–1255. [269] G. Bačid, M. Mojovid, EPR s in tra ing of oxygen radicals in methodological overview, Ann. NY Acad. Sci. 1048 (2005) 230–243.
lants a
[270] A. Steffen-Heins, B. Steffens, EPR spectroscopy and its use in planta—a promising technique to disentangle the origin of specific ROS, Front. Environ. Sci. 3 (2015) 1–6. [271] T. Oda, T. Akaike, K. Sato, A Ishimatsu, S. Takeshita, T. Muramatsu, H. Maeda, Hydroxyl radical generation by red tide algae, Arch. Biochem. Biophys. 294 (1992) 38– 43. [272] D.K. Yadav, A. Prasad, J. Kruk, P. Pospísil, Evidence for the involvement of loosely bound plastosemiquinones in superoxide anion radical production in Photosystem II, PLoS ONE 9 (2014) e115466. [273] A. Krieger-Liszkay, P.B. Kós, É. Hideg, Superoxide anion radicals generated by methylviologen in photosystem I damage photosystem II, Physiol. Plant. 142 (2011) 17‒25.
ACCEPTED MANUSCRIPT 84
[274] K. Liu, J. Sun, Y.G. Song, B. Liu, Y.K. Xu, S.X. Zhang, Q. Tian, Y. Liu, Superoxide, hydrogen peroxide and hydroxyl radical in D1/D2/cytochrome b-559 Photosystem II reaction center complex, Photosynth. Res. 81 (2004) 41–47.
T
[275] M. Mojovid, M. Vuletid, G.G. Bačid, Ž. Vučinid, Oxygen radicals roduced by lant plasma membranes: an EPR spin-trap study, J. Exp. Bot. 55 (2004) 2523–2531.
RI P
[276] J. Bogdanovic Pristov, S. Veljovic Jovanovic, A. Mitrovic, I. Spasojevic, UVirradiation provokes generation of superoxide on cell wall polygalacturonic acid, Physiol. Plant. 148 (2013) 574–581.
SC
[277] L. Chen, H. Jia, Q. Tian, L. Du, Y. Gao, X. Miao, Y. Liu, Protecting effect of phosphorylation on oxidative damage of D1 protein by down-regulating the production of superoxide anion in photosystem II membranes under high light, Photosynth. Res. 112 (2012) 141–148.
NU
[278] M. Kozuleva, I. Klenina, I. Proskuryakov, I. Kirilyuk, B. Ivanov, Production of superoxide in chloroplast thylakoid membranes. ESR study with cyclic hydroxylamines of different lipophilicity, FEBS Lett. 585 (2011) 1067‒1071.
MA
[279] S.I. Dikalov, I.A. Kirilyuk, M. Voinov, I.A. Grigor’ev, EPR detection of cellular and mitochondrial superoxide using cyclic hydroxylamines, Free Radic. Res. 45 (2011) 417–430.
ED
[280] N.N. Vylegzhanina, L.K. Gordon, F.V. Minibayeva, O.P. Kolesnikov, Superoxide production as a stress response of wounded root cells: ESR spin-trap and acceptor methods, Appl. Magn. Reson. 21 (2001) 63‒70.
PT
[281] F. Minibayeva, R.P. Beckett, High rates of extracellular superoxide production in bryophytes and lichens, and an oxidative burst in response to rehydration following desiccation, New Phyt. 152 (2001) 333–341.
AC CE
[282] G.M. Rosen, E. Finkelstein, E.J. Rauckman, A method for the detection of superoxide in biological systems, Arc. Biochem. Biophys. 215 (1982) 367–378. [283] R.A.J. Hodgson, J.K. Raison, Superoxide production by thylakoids during chilling and its implication in the susceptibility of plants to chilling-induced photoinhibition, Planta 183 (1991) 222–228. [284] N. Warwar, A. Mor, R. Fluhr, R.P. Pandian, P. Kuppusamy, A. Blank, Detection and imaging of superoxide in roots by an electron spin resonance spin-probe method, Biophys. J. 101 (2011) 1529–1538. [285] S. Rinalducci, J.Z. Pedersen, L. Zolla, Formation of radicals from singlet oxygen produced during photoinhibition of isolated light-harvesting proteins of photosystem II, Biochim. Biophys. Acta 1608 (2004) 63‒73. [286] C. Frejaville, H. Karoui, B. Tuccio, F. Le Moigne, M. Culcasi, S. Pietri, R. Lauricella, P. Tordo, 5-(Diethoxyphosphoryl)-5-methyl-l-pyrroline N-Oxide: A new efficient phosphorylated nitrone for the in vitro and in vivo spin trapping of oxygen-centered radicals, J. Med. Chem. 38 (1995) 258−265. [287] G. Olive, A. Mercier, F. Le Moigne, A. Rockenbauer, P. Tordo, 2-ethoxycarbonyl2-methyl-3,4-dihydro-2h-pyrrole-1-oxide: Evaluation of the spin trapping properties, Free Radic. Biol. Med. 28 (2000) 403–408. [288] V.K. Kutala, F.A. Villamena, G. Ilangovan, D. Maspoch, N. Roques, J. Veciana, C. Rovira, P. Kuppusamy, Reactivity of superoxide anion radical with a perchlorotriphenylmethyl (trityl) radical, J. Phys. Chem. B 112 (2008) 158–167.
ACCEPTED MANUSCRIPT 85
[289] K.J. Liu, M. Miyake, T. Panz, H. Swartz, Evaluation of DEPMPO as a spin trapping agent in biological systems, Free Radical Biol. Med. 26 (1999) 714–721.
T
[290] I. Šnyrychová, Im rovement of the sensitivity of EPR s in tra ing in biological systems by cyclodextrins: A model study with thylakoids and photosystem II particles, Free Radic. Biol. Med. 48 (2010) 264–274.
RI P
[291] G.R. Buettner, L.W. Oberley, Considerations in the spin trapping of superoxide and hydroxyl radical in aqueous systems using 5,5-dimethyl-1-pyrroline-1-oxide, Biochem. Biophys. Res. Commun. 83 (1978) 69‒74.
SC
[292] H. Zhao, J. Joseph, H. Zhang, H. Karoui, B. Kalyanaraman, Synthesis and biochemical applications of a solid cyclic nitrone spin trap: a relatively superior trap for detecting superoxide anions and glutathiyl radicals, Free Radic. Biol. Med. 31 (2001) 599–606.
NU
[293] K. Stolze, N. Rohr-Udilova, T. Rosenau, R. Stadtmüller, H. Nohl, Very stable superoxide radical adducts of 5-ethoxycarbonyl- 3,5-dimethyl-pyrroline N-oxide (3,5EDPO) and its derivatives, Biochem. Pharmacol. 69 (2005) 1351–1361.
MA
[294] E. Finkelstein, G.M. Rosen, E.J. Rauckman, Production of hydroxyl radical by decomposition of superoxide spin-trapped products, Mol. Pharmacol. 21 (1982) 262‒265.
ED
[295] C.F. Chignell, A.G. Motten, R.H. Sik, C.E. Parker, K. Reszka, A spin-trapping study of the photochemistry of 5,5-dimethyl-1-pyrroline n-oxide (DMPO), Photochem. Photobiol. 59 (1994) 5–11.
PT
[296] M. Akita, Y. Ohta, The effect of Tiron, a water soluble radical scavenger, on growth, morphology and alkaloid content of adventitious roots in Atropa belladonna, Plant Cell Rep. 19 (2000) 705–709.
AC CE
[297] N. Khan, C.M. Wilmot, G.M. Rosen, E. Demidenko, J. Sun, J. Jose h, J. O’hara, B. Kalyanaraman, H.M. Swartz, Spin traps: in vitro toxicity and stability of radical adducts, Free Radic. Biol. Med. 34 (2003) 1473–1481. [298] K. Anzai, T. Aikawa, Y Furukawa, Y. Matsushima, S. Urano, T. Ozawa, ESR measurement of rapid penetration of DMPO and DEPMPO spin traps through lipid bilayer membranes, Arch. Biochem. Biophys. 415 (2003) 251–256. [299] S.K. Ramu, H.M. Peng, D.R. Cook, Nod factor induction of reactive oxygen species production is correlated with expression of the early nodulin gene rip1 in Medicago truncatula, Mol. Plant Microbe Interact. 15 (2002) 522–528. [300] V.C. Mai, W. Bednarski, B. Borowiak-Sobkowiak, B. Wilkaniec, S. Samardakiewicz, I. Morkunas, Oxidative stress in pea seedling leaves in response to Acyrthosiphon pisum infestation, Phytochemistry 93 (2013) 49–62. [301] J. Waring, M. Klenell, U. Bechtold, G.J.C., Underwood, N.R. Baker, Light-induced responses of oxygen photoreduction, reactive oxygen species production and scavenging in two diatom species, J. Phycol. 46 (2010) 1206–1217. [302] E. Omoto, H. Nagao, M. Taniguchi, H. Miyake, Localization of reactive oxygen species and change of antioxidant capacities in mesophyll and bundle sheath chloroplasts of maize under salinity, Physiol. Plant. 149 (2013) 1–12. [303] C.F.G. Bournonville, J.C. Díaz-Ricci, Quantitative determination of superoxide in plant leaves using a modified NBT staining method, Phytochem. Anal. 22 (2011) 268– 271.
ACCEPTED MANUSCRIPT 86
[304] N.W. Roehm, G.H. Rodgers, S.M. Hatfield, A.L. Glasebrook, An improved colorimetric assay for cell proliferation and viability utilizing the tetrazolium salt XTT, J. lmmunol. Methods 142 (1991) 257–265.
RI P
T
[305] M.W. Sutherland, B.A. Learmonth, The tetrazolium dyes MTS and XTT provide new quantitative assays for superoxide and superoxide dismutase, Free Radical Res. 27 (1997) 283–289.
SC
[306] A.J. Able, D.I. Guest, M.W. Sutherland, Use of a new tetrazolium-based assay to study the production of superoxide radicals by tobacco cell cultures challenged with avirulent zoospores of Phytophthora parasitica var nicotianae, Plant Physiol. 117 (1998) 491–499.
NU
[307] Z.S. Zhang, C. Yang, H.Y. Gao, L.T. Zhang, X.L. Fan, M.J. Liu, The higher sensitivity of PSI to ROS results in lower chilling–light tolerance of photosystems in young leaves of cucumber, J. Photochem. Photobiol. B: Biol. 137 (2014) 127–134.
MA
[308] B.H.J. Bielski, G.G. Shiue, S. Bajuk, Reduction of nitro blue tetrazolium by CO2and O2- radicals, J. Phys. Chem. 84 (1980) 830–833. [309] S.I. Liochev, I. Batinic-Haberle, I. Fridovich, The effect of detergents on the reduction of tetrazolium salts, Arch. Biochem. Biophys. 324 (1995) 48–52.
ED
[310] R.O. Olojo, R.H. Xia, J.J. Abramson, Spectrophotometric and fluorometric assay of superoxide ion using 4-chloro-7-nitrobenzo-2-oxa-1,3-diazole, Anal.l Biochem. 339 (2005) 338–344.
PT
[311] L. Benova, I. Fridovich, Is reduction of the sulfonated tetrazolium 2,3-bis (2methoxy-4-nitro-5-sulfophenyl)-2-tetrazolium 5-carboxanilide a reliable measure of intracellular superoxide production? Anal. Biochem. 310 (2002) 186–190.
AC CE
[312] D. Vyas, R. Sahoo, S. Kumar, Possible mechanism and implications of phenolicsmediated reduction of XTT (sodium, 3’-[1-[phenylaminocarbonyl]-3,4-tetrazolium]bis(4-methoxy-6-nitro) benzene-sulfonic acid hydrate), Current Science 83 (2002) 1588–1592. [313] K. Asada, K. Kiso, K. Yoshikawa, Univalent reduction of molecular oxygen by spinach chloroplasts on illumination, J. Biol. Chem. 249 (1974) 2175–2181. [314] C. Sgherri, M.F. Quartacci, F. Navari-Izzo, Early production of activated oxygen species in root apoplast of wheat following copper excess, J. Plant Physiol. 164 (2007) 1152‒1160. [315] S.A. Khorobrykh, B.N. Ivanov, Oxygen reduction in a plastoquinone pool of isolated pea thylakoids, Photosynth. Res. 71 (2002) 209–219. [316] J. Kruk, M. Jemiola-Rzemioska, K. Strzalka, Cytochrome c is reduced mainly by plastoquinol and not by superoxide in thylakoid membranes at low and medium light intensities: its specific interaction with thylakoid membrane lipids, Biochem. J. 375 (2003) 215–220. [317] L. Thomson, M. Trujillo, R. Telleri, R. Radi, Kinetics of cytochrome c2+ oxidation by peroxynitrite: implications for superoxide measurements in nitric oxide-producing biological systems, Arch. Biochem. Biophys. 319 (1995) 491–497. [318] A. Azzi, C. Montecucco, C. Richter, The use of acetylated ferricytochrome c for the detection of superoxide radicals produced in biological membranes, Biochem. Biophys. Res. Commun. 65 (1975) 597–603.
ACCEPTED MANUSCRIPT 87
[319] H.P. Misra, I. Fridovich, The univalent reduction of oxygen by reduced flavins and quinones, J. Biol. Chem. 247 (1972) 188–192.
RI P
T
[320] T. Roach, M. Ivanova, R.P. Beckett, F.V. Minibayeva, I. Green, H.W. Pritchard, I. Kranner, An oxidative burst of superoxide in embryonic axes of recalcitrant sweet chestnut seeds as induced by excision and desiccation, Physiol. Plant. 133 (2008) 131– 139.
SC
[321] B. Szal, M. Drozd, A.M. Rychter, Factors affecting determination of superoxide anion generated by mitochondria from barley roots after anaerobiosis, J. Plant Physiol. 161 (2004) 1339–1346. [322] Y. Mathieu, M.A. Rouet-Mayer, H. Barbier-Brygoo, C. Laurière, Activation by fatty acids of the production of active oxygen species by tobacco cells, Plant Physiol. Biochem. 40 (2002) 313–324.
NU
[323] C. Chen, D.R. Thakker, The fallacy of using adrenochrome reaction for measurement of reactive oxygen species formed during Cytochrome P450-mediated metabolism of xenobiotics, J. Pharmacol. Exp. Ther. 300 (2002) 417–420.
MA
[324] M.L. Genova, N.M. Abd-Elsalam, E.S.M.E. Mahdy, A. Bernacchia, M. Lucarini, G.F. Pedulli, G. Lena, Redox cycling of adrenaline and adrenochrome catalyzed by mitochondrial Complex I, Arch. Biochem. Biophys. 447 (2006) 167–173.
ED
[325] E.F. Elstner, A. Heupel, Inhibition of nitrite formation from bydroxylammonium chloride: a simple assay for superoxide dismutase, Anal. Biochem. 70 (1976) 616–620.
PT
[326] S. Erdal, Androsterone-induced molecular and physiological changes in maize seedlings in response to chilling stress, Plant Physiol. Biochem. 57 (2012) 1‒7.
AC CE
[327] F.E. Callahan, D.W. Becker, G.M. Cheniae, Studies on the photoinactivation of the water-oxidizing enzyme. II. Characterization of weak light photoinhibition of PSII and its light-induced recovery, Plant Physiol. 82 (1986) 261–269. [328] M. Hakala, I. Tuominen, M. Keränen, T. Tyystjärvi, E. Tyystjärvi, Evidence for the role of the oxygen-evolving manganese complex in photoinhibition of Photosystem II, Biochim. Biophys. Acta 1706 (2005) 68‒80. [329] R. Michalski, J. Zielonka, M. Hardy, J. Joseph, B. Kalyanaraman, Hydropropidine: A novel, cell-impermeant fluorogenic probe for detecting extracellular superoxide, Free Radic. Biol. Med. 54 (2013) 135–147. [330] L. Garnier, F. Simon-Plas, P. Thuleau, J.-P. Agnel, J.-P. Blein, R. Ranjeva, J.-L. Montillet, Cadmium affects tobacco cells by a series of three waves of reactive oxygen species that contribute to cytotoxicity, Plant Cell Environ. 29 (2006) 1956–1969. [331] K. M. Robinson, M.S. Janes, M. Pehar, J.S. Monette, M.F. Ross, T.M. Hagen, M.P. Murphy, J.S. Beckman, Selective fluorescent imaging of superoxide in vivo using ethidium-based probes, Proc. Natl. Acad. Sci. USA 103 (2006) 15038–15043. [332] L. Benov, L. Sztejnberg, I. Fridovich, Critical evaluation of the use of hydroethidine as a measure of superoxide anion radical, Free Radic. Biol. Med. 25 (1998) 826–831.
ACCEPTED MANUSCRIPT 88
[333] I. Papapostolou, N. Patsoukis, C.D. Georgiou, The fluorescence detection of superoxide radical using hydroethidine could be complicated by the presence of heme-proteins, Anal. Biochem. 332 (2004) 290–298.
RI P
T
[334] N. Patsoukis, I. Papapostolou, C.D. Georgiou, Interference of nonspecific peroxidases in the fluorescence detection of superoxide radical by hydroethidine oxidation: A new assay for H2O2, Anal. Bioanal. Chem. 381 (2005) 1065–1072.
SC
[335] H. Zhao, J. Joseph, H.M. Fales, E.A. Sokoloski, R.L. Levine, J. Vasquez-Vivar, B. Kalyanaraman, Detection and characterization of the product of hydroethidine and intracellular superoxide by HPLC and limitations of fluorescence, Proc. Natl. Acad. Sci. USA 102 (2005) 5727–5732.
NU
[336] J. Zielonka, B. Kalyanaraman, Hydroethidine- and MitoSOX-derived red fluorescence is not a reliable indicator of intracellular superoxide formation: Another inconvenient truth, Free Radic. Biol. Med. 48 (2010) 983‒1001.
MA
[337] J. Zielonka, J. Vasquez-Vivar, B. Kalyanaraman, The confounding effects of light, sonication, and Mn(III)TBAP on quantitation of superoxide using hydroethidine, Free Radic. Biol. Med. 41 (2006) 1050–1057.
ED
[338] Y. Yamamoto, Y. Kobayashi, S.R. Devi, S. Rikiishi, H. Matsumoto, Aluminum toxicity is associated with mitochondrial dysfunction and the production of reactive oxygen species in plant cells, Plant Physiol. 128 (2002) 63–72.
PT
[339] C.D. Georgiou, I. Papapostolou, N. Patsoukis, T. Tsegenidis, T. Sideris, An ultrasensitive fluorescent assay for the in vivo quantification of superoxide radical in organisms, Anal. Biochem. 347 (2005) 144‒151.
AC CE
[340] P. Back, F. Matthijssens, J.R. Vanfleteren, B.P. Braeckman, A simplified hydroethidine method for fast and accurate detection of superoxide production in isolated mitochondria, Anal. Biochem. 423 (2012) 147–151. [341] M.V. Martin, D.F. Fiol, V. Sundaresan, E.J. Zabaleta, G.C. Pagnussat, oiwa, a female gametophytic mutant impaired in a mitochondrial manganese-superoxide dismutase, reveals crucial roles for reactive oxygen species during embryo sac development and fertilization in Arabidopsis, Plant Cell 25 (2013) 1573–1591.
[342] C.K. Auh, T.M. Murphy, Plasma membrane redox enzyme is involved in the synthesis of O2- and H2O2 by Phytophthora elicitor-stimulated rose cells, Plant Physiol. 107 (1995) 1241–1247.
[343] S.I. Liochev, I. Fridovich, Lucigenin as a mediator of superoxide production: Revisited, Free Radic. Biol. Med. 25 (1998) 926–928.
ACCEPTED MANUSCRIPT 89
RI P
T
[344] M. Janiszewski, H.P. Souza,X. Liu, M.A. Pedro, J.L. Zweier, F.R.M. Laurindo, Overestimation of NADH-driven vascular oxidase activity due to lucigenin artifacts, Free Radic. Biol. Med. 32 (2002) 446‒453.
SC
[345] I.A. Schepetkin, Lucigenin as a substrate of microsomal NAD(P)Hoxidoreductases, Biochemistry-Moscow 64 (1999) 25‒32.
MA
NU
[346] J. Zielonka, J.D. Lambeth, B. Kalyanaraman, On the use of L-012, a luminol-based chemiluminescence probe, for detecting superoxide and identifying inhibitors of NADPH oxidase: a reevaluation, Free Radic. Biol. Med. 65 (2013) 1310‒1314.
ED
[347] Y. Miura, H. Yoshioka, N. Doke, An autophotographic determination of the active oxygen generation in potato tuber discs during hypersensitive response to fungal infection or elicitor, Plant Sci. 105 (1995) 45–52.
AC CE
PT
[348] B. Pourrut, G. Perchet, J. Silvestre, M. Cecchi, M. Guiresse, E. Pinelli, Potential role of NADPH-oxidase in early steps of lead-induced oxidative burst in Vicia faba roots, J. Plant Physiol. 165 (2008) 571‒579.
[349] M. Kobayashi, I. Ohura, K. Kawakita, N. Yokota, M. Fujiwara, K. Shimamoto, N. Doke, H. Yoshioka, Calcium-dependent protein kinases regulate the production of reactive oxygen species by potato NADPH oxidase, Plant Cell 19 (2007) 1065–1080.
[350] E.K. Miller, I. Fridovich, A demonstration that O2- is a crucial intermediate in the high quantum yield luminescence of luminol, J. Free Radic. Biol. Med. 2 (1986) 107– 110.
[351] Y. Nishinaka, Y. Aramaki, H. Yoshida, H. Masuya, T. Sugawara, Y. Ichimori, A new sensitive chemiluminescence probe, L-012, for measuring the production of superoxide anion by cells, Biochem. Biophys. Res. Commun. 193 (1993) 554‒559.
[352] T. Kawano, N. Sahashi, K. Takahashi, N. Uozumi, S. Muto, Salicylic acid induces extracellular superoxide generation followed by an increase in cytosolic calcium ion in tobacco suspension culture: the earliest events in salicylic acid signal transduction, Plant Cell Physiol. 39 (1998) 721–730.
ACCEPTED MANUSCRIPT 90
RI P
T
[353] I.C. Mori, R. Pinontoan, T. Kawano, S. Muto, Involvement of superoxide generation in salicylic acid-induced stomatal closure in Vicia faba, Plant Cell Physiol. 42 (2001) 1383–1388.
SC
[354] S. Sun, X. Li, G. Zhang, H. Ma, D. Zhang, Z. Bao, Determination of H2O2dependent generation of singlet oxygen from human saliva with a novel chemiluminescence probe, Biochim. Biophys. Acta 1760 (2006) 440–444.
MA
NU
[355] J. Zheng, S.R. Springston, J. Weinstein-Lloyd, Quantitative analysis of hydroperoxyl radical using flow injection analysis with chemiluminescence detection, Anal. Chem. 75 (2003) 4696‒4700.
ED
[356] A. Godrant, A.L. Rose, G. Sarthou, T.D. Waite, New method for the determination of extracellular production of superoxide by marine phytoplankton using the chemiluminescence probes MCLA and red-CLA, Limnol. Oceanogr. Methods 7 (2009) 682–692.
AC CE
PT
[357] A. Milne, M.S. Davey, P. J. Worsfold, E.P. Achterberg, A.R. Taylor, Real-time detection of reactive oxygen species generation by marine phytoplankton using flow injection–chemiluminescence, Limnol. Oceanogr. Methods 7 (2009)706–715.
[358] Z.F. Deng, Q. Rui, X. Yin, HQ. Liu, Y. Tian, In vivo detection of superoxide anion in bean sprout based on ZnO nanodisks with facilitated activity for direct electron transfer of superoxide dismutase, Anal. Chem. 80 (2008) 5839–5846.
[359] O. Kocabay, E. Emregul, S. Aras, K.C. Emregul, Carboxymethylcellulose-gelatin superoxidase dismutase electrode for amperometric superoxide radical sensing, Bioprocess Biosyst. Eng. 35 (2012) 923–930.
[360] O. Kocabay, E. Emregul, S.S. Aydin, S. Aras, Detection of superoxide radicals in tomato plants exposed to salinity, drought, cold and heavy metal stress using CMC-GSOD biosensor, Artif. Cells Nanomed. Biotechnol. 41 (2013) 352–358.
[361] W.Wang, H. Fang, L. Groom, A. Cheng, W. Zhang, J. Liu, X. Wang, K. Li, P. Han, M. Zheng, J. Yin, W. Wang, M.P. Mattson, J.P.Y. Kao, E.G. Lakatta, S.-S. Sheu, K. Ouyang, J. Chen, R.T. Dirksen, H. Cheng, Superoxide flashes in mitochondria, Cell 134 (2008) 279‒290.
ACCEPTED MANUSCRIPT 91
RI P
T
[362] J. He, Y. Duan, D. Hua, G. Fan, L. Wang, Y. Liu, Z. Chen, L. Han, L.J. Qu, Z. Gong, DEXH Box RNA Helicase-mediated mitochondrial reactive oxygen species production in Arabidopsis mediates crosstalk between abscisic acid and auxin signaling, Plant Cell 24 (2012) 1815–1833.
SC
[363] M. Schwarzländer, D.C. Logan, M.D. Fricker, L.J. Sweetlove, The circularly permuted yellow fluorescent protein cpYFP that has been used as a superoxide probe is highly responsive to pH but not superoxide in mitochondria: implications for the existence of su eroxide ‘flashes’, Biochem. J. 437 (2011) 381–387.
MA
NU
[364] M. Schwarzländer, S. Wagner, Y.G. Ermakova, V.V. Belousov, R. Radi, J.S. Beckman, G.R. Buettner, N. Demaurex, M.R. Duchen, H.J. Forman, M.D. Fricker, D. Gems, A.P. Halestrap, B. Halliwell, U. Jakob, I.G. Johnston, N.S. Jones, D.C. Logan, B. Morgan, F.L. Müller, D.G. Nicholls, S.J. Remington, P.T. Schumacker, C.C. Winterbourn, L.J. Sweetlove, A.J. Meyer, T.P. Dick, M.P. Mur hy, The ‘mitoflash’ robe c YFP does not respond to superoxide, Nature 514 (2014) E12–E14.
ED
[365] O.K. Okamoto, J.W. Hastings, Genome-wide analysis of redox-regulated genes in a dinoflagellate, Gene 321 (2003) 73–81.
PT
[366] T.M. Murphy, H. Vu, T. Nguyen, The superoxide synthases of rose cells, comparison of assays, Plant Physiol. 117 (1998) 1301–1305.
AC CE
[367] M.M. Tarpey, I. Fridovich, Methods of detection of vascular reactive species nitric oxide, superoxide, hydrogen peroxide, and peroxynitrite, Circ. Res. 89 (2001) 224–236. [368] G.R. Buettner, The pecking order of free radicals and antioxidants: lipid peroxidation, α-tocopherol, and ascorbate, Arch. Biochem. Biophys. 300 (1993) 535– 543. [369] G.V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross, Critical review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (•OH/•O-) in aqueous solution, J. Phys. Chem. 17(1988) 513‒886. [370] W.R. Haag, C.C.D. Yao Rate constants for reaction of hydroxyl radicals with several drinking water contaminants, Environ. Sci. Technol. 26 (1992) 1005–1013. [371] E. Codorniu-Hernández, P.G. Kusalik, Mobility mechanism of hydroxyl radicals in aqueous solution via hydrogen transfer, J. Am. Chem.Soc. 134 (2012) 532−538. [372] S.X. Chen, P. Schopfer, Hydroxyl-radical production in physiological reactions. A novel function of eroxidase, Eur. J. Biochem. 260 (1999) 726−735. [373] M.L. McCormick, T.L. Roeder, M.A. Railsback, B.E. Britigan, Eosinophil peroxidase-dependent hydroxyl radical generation by human eosinophils, J. Biol. Chem. 269 (1994) 27914−27919.
ACCEPTED MANUSCRIPT 92
T
[374] L. Zang, H. He, Q. Xu, Y. Yu, N. Zheng, W. Liu, T. Hayashi, S. Tashiro, S. Onodera, T. Ikejima, Reactive oxygen species H2O2 and •OH, but not O2•− promote oridonininduced phagocytosis of apoptotic cells by human histocytic lymphoma U937 cells, Int. Immunopharmacol. 15 (2013) 414–423.
RI P
[375] S.L. Richards, K.A. Wilkins, S.M. Swarbreck, A.A. Anderson, N.Habib, A.G. Smith, M. McAinsh, J.M. Davies, The hydroxyl radical in plants: from seed to seed, J. Exp. Bot. 66 (2014) 37−46.
SC
[376] S.C. Fry, Oxidative scission of plant cell wall polysaccharides by ascorbateinduced hydroxyl radicals, Biochem. J. 332 (1998) 507−515.
NU
[377] C. Whitaker, R.P. Beckett, F.V. Minibayeva, I. Kranner, Alleviation of dormancy by reactive oxygen species in Bidens pilosa L. seeds, S. Afr. J. Bot. 76 (2010) 601–605.
MA
[378] V. Hadži-Taškovid Šukalovid, M. Vuletid, S. Veljovid-Jovanovid, Ž. Vučinid, The effects of manganese and copper in vitro and in vivo on peroxidase catalytic cycles, J. Plant Physiol. 167 (2010) 1550‒1557.
ED
[379] B. Kukavica, M. Mojovid, Ž. Vučinid, V. Maksimovid, U.Takahama, S. VeljovidJovanovid, Generation of hydroxyl radical in isolated ea root cell wall, and the role of cell wall-bound peroxidase, Mn-SOD and phenolics in their production, Plant Cell Physiol. 50 (2009) 304–317.
PT
[380] E. Heyno, V. Mary, P. Schopfer, A. Krieger-Liszkay, Oxygen activation at the plasma membrane: relation between superoxide and hydroxyl radical production by isolated membranes, Planta 234 (2011) 35–45. [381] S.C. Fry, J.G. Miller, J.C. Dumville, A proposed role for copper ions in cell wall loosening, Plant and Soil 247 (2002) 57–67.
AC CE
[382] I. Yruela, Transition metals in plant photosynthesis, Metallomics 5 (2013) 10901109. [383] S.E. Tjus, H.V. Scheller, B. Andersson, B.L. Møller, Active oxygen produced during selective excitation of Photosystem I is damaging not only to Photosystem I, but also to Photosystem II, Plant Physiol. 125 (2001) 2007‒2015. [384] K. Asada, Production and scavenging of reactive oxygen species in chloroplasts and their functions, Plant Physiol. 141 (2006) 391–396. [385] B. Jakob, U. Heber, Photoproduction and detoxification of hydroxyl radicals in chloroplasts and leaves and relation to photoinactivation of Photosystems I and II, Plant Cell Physiol. 37 (1996) 629−635. [386] P. Pos íšil, A. Arató, A. Krieger-Liszkay, A.W. Rutherford, Hydroxyl radical generation by Photosystem II, Biochemistry 43 (2004) 6783−6792. [387] A. Arató, N. Bondarava, A. Krieger-Liszkay, Production of reactive oxygen species in chloride- and calcium-depleted Photosystem II and their involvement in photoinhibition, Biochim. Biophys. Acta 1608 (2004) 171‒180. [388] A.Yamashita, N. Nijo, P. Pos íšil, N. Morita, D. Takenaka, R. Aminaka, Y. Yamamoto, Y. Yamamoto, Quality control of Photosystem II - Reactive oxygen species are responsible for the damage to Photosystem II under moderate heat stress, J. Biol. Chem. 283 (2008) 28380–28391.
ACCEPTED MANUSCRIPT 93
[389] E. Keunen, T. Remans , S. Bohler , J. Vangronsveld, A. Cuypers, Metal-induced oxidative stress and lant mitochondria, Int. J. Mol. Sci. 12 (2011) 6894−6918.
RI P
T
[390] G.N.R. Tripathi, Electron-transfer component in hydroxyl radical reactions observed by time resolved resonance Raman spectroscopy, J. Am. Chem. Soc. 120 (1998) 4161–4166. [391] S.C. Fisher, M.A. Schoonen, B.J. Brownawell, Phenylalanine as a hydroxyl radical-specific probe in pyrite slurries, Geochem Trans. 13 (2012).
SC
[392] C. Leeuwenburgh, P. Hansen, A. Shaish, J.O. Holloszy, J.W. Heinecke, Markers of protein oxidation by hydroxyl radical and reactive nitrogen species in tissues of aging rats, Am. J. Physiol. 274 (1998) 453‒461.
NU
[393] W.H. Koppenol, J. Butler, Energetics of interconversion reactions of oxyradicals, Adv. Free Radical Biol. Med. 1 (1985) 91–131.
MA
[394] H. Sies, Strategies of antioxidant defense, Eur. J. Biochem. 215 (1993) 213−219. [395] W.A. Pryor, Oxy-radicals and related species: Their formation, lifetimes, and reactions, Ann. Rev. Physiol. 48 (1986) 657–67.
ED
[396] C.F. Babbs, M.J. Gale, Colorimetric assay for methanesulfinic acid in biological sam les, Anal. Biochem. 163 (1987) 67−73.
PT
[397] M. Becana, R.V. Klucas, Transition metals in legume root nodules: Irondependent free radical production increases during nodule senescence, Proc. Natl. Acad. Sci. USA. 89 (1992) 8958−8962.
AC CE
[398] C.F. Babbs, J.A. Pham, R.C. Coolbaugh, Lethal hydroxyl radical production in paraquat-treated lants, Plant Physiol. 90 (1989) 1267−1270. [399] P.L. Popham, A. Novacky, Use of dimethyl sulfoxide to detect hydroxyl adical during bacteria-induced hypersensitive reaction, Plant Physiol. 96 (1991) 1157−1160. [400] H. Tschiersch, E. Ohmann, Photoinhibition in Euglena gracilis: Involvement of reactive oxygen species, Planta 191 (1993) 316‒323. [401] K.H. Cheeseman, A. Beavis, H. Esterbauer, Hydroxyl-radical-induced ironcatalysed degradation of 2-deoxyribose, Biochem. J. 252 (1988) 649−653. [402] X. Li, W. Mai, L. Wang, W. Han, A hydroxyl-scavenging assay based on DNA damage in vitro, Anal. Biochem. 438 (2013) 29–31. [403] D. Malenčid, S. Kevrešan, M. Po ovid, D. Štajner, B. Po ovid, B. Ki rovski, S. Djurid, Cholic acid changes defense res onse to oxidative stress in soybean induced by As ergillus niger, Cent. Eur. J. Biol. 7 (2012) 132−137. [404] V.M. Achary, N.L. Parinandi, B.B. Panda, Aluminum induces oxidative burst, cell wall NADH peroxidase activity, and DNA damage in root cells of Allium cepa L., Environ. Mol. Mutagen. 53 (2012) 550−560. [405] G.R. Buettner, R.P. Mason, in: R.G. Cutler, H. Rodriguez (Eds.), Spin-trapping methods for detecting superoxide and hydroxyl free radicals in vitro and in vivo, World Scientific, New Jersey, London, Singapore, Hong Kong, 2003, pp 27‒38.
ACCEPTED MANUSCRIPT 94
[406] B. Halliwell, J.M. Gutteridge, Biologically relevant metal ion-dependent hydroxyl radical generation. An u date, FEBS Lett. 307 (1992) 108−12.
SC
RI P
T
[407] A. Liszkay, E. van der Zalm, P. Schopfer, Production of reactive oxygen intermediates (O2•-, H2O2, and •OH) by maize roots and their role in wall loosening and elongation growth, Plant Physiol. 136 (2004) 3114–3123. [408] M.D. Engelmann, R.T. Bobier, T. Hiatt, I.F. Cheng, Variability of the Fenton reaction characteristics of the EDTA, DTPA, and citrate complexes of iron, BioMetals 16 (2003) 519–527. [409] I. Snyrychova, P. Pospisil, J. Naus, The effect of metal chelators on the production of hydroxyl radicals in thylakoids, Photosynth. Res. 88 (2006) 323–329.
NU
[410] M.L. McCormick, G.R. Buettner, B.E. Britigan, The spin trap α-(4-Pyridyl-1oxide)-N-tert-butylnitrone stimulates peroxidase-mediated oxidation of deferoxamine, J. Biol. Chem 270 (1995) 29265–29269
MA
[411] E.G. Janzen, Y.Y. Wang, R.V. Shetty, Spin trapping with α-Pyridyl 1-Oxide N-tertButyl nitrones in aqueous solutions. A unique electron spin resonance spectrum for the hydroxyl radical adduct, J. Am. Chem. Soc. 100 (1978) 2923−2925.
ED
[412] B.Z Zhu, H.T. Zhao, B. Kalyanaraman, B. Frei, Metal-independent production of hydroxyl radicals by halogenated quinones and hydrogen peroxide: An ESR spin trapping study, Free Radic. Biol. Med. 32 (2002) 465–473.
AC CE
PT
[413] S. Pou, C.L. Ramos, T. Gladwell, E. Renks, M. Centra, D. Young, M.S. Cohen, G.M. Rosen, A kinetic approach to the selection of a sensitive spin trapping system for the detection of hydroxyl radical, Anal. Biochem. 217 (1994) 76−83. [414] S. Van Doorslaer, A. Dedonder, M. de Block, F. Callens, Oxidative stress in plants: EPR monitoring in DMPO-DMSO based extracts, J. Plant Physiol. 154 (1999) 132−136. [415] G.S Timmins, K.J. Liu, E.J.H. Bechara, Y. Kotake, H.M. Swartz, Trapping of free radicals with direct in vivo EPR detection: A comparison of 5,5-dimethyl-1-pyrrolineN-oxide and 5-diethoxyphosphoryl-5-methyl-1-pyrroline-N-oxide as s in tra s for HO• and SO4•-, Free Radic. Biol. Med. 27 (1999) 329–333. [416] P. Schopfer, A. Liszkay, M. Bechtold, G. Frahry, A. Wagner, Evidence that hydroxyl radicals mediate auxin-induced extension growth, Planta 214 (2002) 821−828. [417] A. Liszkay, B. Kenk, P. Schopfer, Evidence for the involvement of cell wall peroxidase in the generation of hydroxyl radicals mediating extension growth, Planta 217 (2003) 658–667. [418] E. Heyno, C. Klose, A. Krieger-Liszkay, Origin of cadmium-induced reactive oxygen species production: mitochondrial electron transfer versus plasma membrane NADPH oxidase, New Phyt. 179 (2008) 687−699.
ACCEPTED MANUSCRIPT 95
T
[419] A. Rastogi, P. Pos íšil, Production of hydrogen eroxide and hydroxyl radical in potato tuber during the necrotrophic phase of hemibiotrophic pathogen Phytophthora infestans infection, J. Photochem. Photobiol. 117 (2012) 202–206.
RI P
[420] S. Deng, M. Yu, Y. Wang, Q. Jia, L. Lin, H. Dong, The antagonistic effect of hydroxyl radical on the development of a hypersensitive response in tobacco, FEBS J. 277 (2010) 5097–5111.
NU
SC
[421] S. Pou, Y.I. Huang, A. Bhan, V.S. Bhadti, R.S. Hosmane, S.Y. Wu, G.L. Cao, G.M. Rosen, A fluorophore-containing nitroxide as a probe to detect superoxide and hydroxyl radical generated by stimulated neutrophils, Anal. Biochem. 212 (1993) 85−90.
MA
[422] S. Renew, E. Heyno, P. Schopfer, A. Liszkay, Sensitive detection and localization of hydroxyl radical production in cucumber roots and Arabidopsis seedlings by spin trapping electron paramagnetic resonance spectroscopy, Plant J. 44 (2005) 342–347.
AC CE
PT
ED
[423] K. Müller, A. Linkies, R.A.M. Vreeburg, S.C. Fry, A. Krieger-Liszkay, G. LeubnerMetzger, In vivo cell wall loosening by hydroxyl radicals during cress seed germination and elongation growth, Plant Physiol. 150 (2009) 1855−1865. [424] V. Demidchik, T.A. Cuin, D. Svistunenko, S.J. Smith, A.J. Miller, S. Shabala, A. Sokolik, V. Yurin, Arabidopsis root K+-efflux conductance activated by hydroxyl radicals: single-channel properties, genetic basis and involvement in stress-induced cell death, J. Cell Sci. 123 (2010) 1468−1479. [425] I. Yruela, J.J. Pueyo, P.J. Alonso, R. Picorel, Photoinhibition of Photosystem II from higher plants - Effect of copper inhibition, J. Biol. Chem. 271 (1996) 27408−27415. [426] F. Navari-Izzo, C. Pinzino, M.F. Quartacci, C.L. Sgherri, Superoxide and hydroxyl radical generation, and superoxide dismutase in PSII membrane fragments from wheat, Free Radic. Res. 31 (1999) S3–S9. [427] S. Zhang, J. Weng, J. Pan, T. Tu, S. Yao, C. Xu, Study on the photo-generation of superoxide radicals in Photosystem II with EPR spin trapping techniques, Photosynth. Res. 75 (2003) 41–48. [428] S. Chen, C. Yin, S. Qiang, F. Zhou, X. Dai, Chloroplastic oxidative burst induced by tenuazonic acid, a natural photosynthesis inhibitor, triggers cell necrosis in Eupatorium adenophorum Spreng, Biochim. Biophys. Acta 1797 (2010) 391–405. [429] C. Fufezan, A.W. Rutherford, A. Krieger-Liszkay, Singlet oxygen production in herbicide-treated hotosystem II, FEBS Lett. 532 (2002) 407−410. [430] P. Pos íšil, I. Šnyrychová, J. Nauš, Dark roduction of reactive oxygen s ecies in photosystem II membrane particles at elevated temperature: EPR spin-trapping study, Biochim. Biophys. Acta 1767 (2007) 854–859. [431] L. Zolla, S. Rinalducci, Involvement of active oxygen species in degradation of light-harvesting proteins under light stresses, Biochemistry 41 (2002) 14391−14402.
ACCEPTED MANUSCRIPT 96
[432] C. Spetea, É. Hideg, I. Vass, Low pH accelerates light-induced damage of Photosystem II by enhancing the probability of the donor-side mechanism of photoinhibition, Biochim. Biophys. Acta 1318 (1997) 275–283.
RI P
T
[433] S. Veljovid-Jovanovid, B. Kukavica, T. Cvetid, M. Mojovid, Ž. Vučinid, Ascorbic acid and the oxidative processes in pea root cell wall isolates - Characterization by fluorescence and EPR spectroscopy, Ann. N.Y. Acad. Sci. 1048 (2005) 500–504.
SC
[434] B. Kukavica, A. Mitrovid, M. Mojovid, S. Veljovid-Jovanovid, Effect of Indole-3acetic acid on pea root growth, peroxidase profiles and hydroxyl radical formation, Arch. Biol. Sci. 59 (2007) 319−326. [435] J.D. Maksimovid, M. Mojovid, V. Maksimovid, V. Römheld, M. Nikolic, Silicon ameliorates manganese toxicity in cucumber by decreasing hydroxyl radical accumulation in the leaf apoplast, J. Exp. Bot. 63 (2012) 2411–2420.
NU
[436] L. Tamás, K. Valentovičová, Ľ. Halušková, J. Huttová, I. Mistrík, Effect of cadmium on the distribution of hydroxyl radical, superoxide and hydrogen peroxide in barley root tip, Protoplasma 236 (2009) 67–72.
MA
[437] N. Liu, Z. Lin, L. Guan, G. Gaughan, G. Lin, Antioxidant enzymes regulate reactive oxygen species during pod elongation in Pisum sativum and Brassica chinensis, PLoS ONE 9 (2014) e87588.
ED
[438] X.F. Yang, X.Q. Guo, Fe(II)–EDTA chelate-induced aromatic hydroxylation of terephthalate as a new method for the evaluation of hydroxyl radical-scavenging ability, Analyst 126 (2001) 928–932.
AC CE
PT
[439] J.C. Barreto, G.S. Smith, N.H. Strobel, P.A. McQuillin, T.A. Miller, Terepthalic acid: A dosimeter for the detection of hydroxyl radicals in vitro, Life Sci. 56 (1995) 89−96. [440] J.Y. Fang, C.-H. Chu, I. Sarang, K.-C. Fang, C.-P. Hsu, Y.-F. Huang, C.-H. Hsu, C.-C. Chen, S.-S. Li, J.A. Yeh, D.-J. Yao, Y.-L. Wang, Electronic hydroxyl radical microsensors based on the conductivity change of polyaniline, Sens. Actuators B: Chem. 208 (2015) 99–105. [441] V. Demidchik, S.N. Shabala, K.B. Coutts, M.A. Tester, J.M. Davies, Free oxygen radicals regulate plasma membrane Ca2+- and K+-permeable channels in plant root cells, J. Cell Sci. 116 (2003) 81−88. [442] R. Mittler, G. Berkowitz, Hydrogen peroxide, a messenger with too many roles? Redox Report 6 (2001) 69–72. [443] H.W. Gardner, in: H.W.S. Chan (ed.) Autoxidation of unsaturated lipids, Academic Press Inc., London, 1987, pp. 51‒93. [444] V.V. Azatyan, E.T. Denisov, Inhibition of Chain Reactions, Gordon & Breach, Amsterdam, 2000. [445] J. A. Howard, K.U. Ingold, Self-reaction of sec-butylperoxy radicals. Confirmation of the Russell mechanism, J. Am. Chem. Soc. 90 (1968) 1056–1058. [446] M. Havaux, Spontaneous and thermoinduced photon emission: new methods to detect and quantify oxidative stress in plants, Trends Plant Sci. 8 (2003) 409‒413.
ACCEPTED MANUSCRIPT 97
T
[447] H.W. Gardner, Oxygen radical chemistry of polyunsaturated fatty acids, Free Rad. Biol. Med. 7 (1989) 65‒86.
RI P
[448] M. Havaux, C. Triantaphylidès, B. Genty, Autoluminescence imaging: a noninvasive tool for mapping oxidative stress, Trends Plant Sci. 11 (2006) 480‒484.
SC
[449] H. Esterbauer, H. Zollner, R.J. Schaur, Aldehydes formed by lipid peroxidation: mechanisms of formation, occurrence and determination, in: C. Vigo-Pelfrey (Ed.), Membrane Lipid Peroxidation, Vol 1, CRC Press, Florida, 1990, pp. 239‒283.
PT
ED
MA
NU
[450] J. Mano, S. Khorobrykh, K. Matsui, Y. Iijima, N. Sakurai, H. Suzuki, D. Shibata, Acrolein is formed from trienoic fatty acids in chloroplast: A targeted metabolomics approach, Plant Biotechnol. 31 (2014) 535–543. [451] D.M. Hodges, J.M. DeLong, C.F. Forney, R.K. Prange, Improving the thiobarbituric acid-reactive-substances assay for estimating lipid peroxidation in plant tissues containing anthocyanin and other interfering compounds, Planta 207 (1999) 604‒611. [452] K. Matsui, K. Sugimoto, P. Kakumyan, S.A. Khorobrykh, J. Mano. Volatile oxylipins formed under stress in plants. In: D. Armstrong (Ed.) Lipidomics, Methods in Molecular Biology, Vol. 580, Humana Press, Totowa, 2009, pp 17‒27. [453] M. Trujillo, G. Ferrer-Sueta, L. Thomson, L. Flohé, R. Radi, Kinetics of peroxiredoxins and their role in the decomposition of peroxynitrite, Subcell. Biochem. 44 (2007) 83‒113. -carotene: An unusual type of lipid antioxidant,
AC CE
[454] Science 224 (1984) 569–573.
[455] A. El-Agamey, G.M. Lowe, D.J. McGarvey, A. Mortensen, D.M. Phillip, T.G. Truscott, A.J. Young, Carotenoid radical chemistry and antioxidant/pro-oxidant properties, Arch. Biochem. Biophys. 430 (2004) 37–48. [456] F.P. Corongiu, A. Milla, An improved and simple method for determining diene conjugation in autoxidized fatty acids, Chem. Biol. Interact. 44 (1983) 289–297. [457] Y. Yamamoto, E. Niki, R. Tanimura, Y. Kamiya, Study of oxidation by chemiluminescence. IV. Detection of low levels of lipid hydroperoxides by chemiluminescence. J. Am. Oil Chem. Soc. 62 (1985) 1248‒1250. [458] A. Szterk, P.P. Lewicki, A new chemiluminescence method for detecting lipid peroxides in vegetable oils, J. Am. Oil Chem. Soc. 87 (2010) 361‒367. [459] D. V. Vavilin, J.-M. Ducruet, The origin of 115-130°C thermoluminescence bands in chlorophyll-containing material, Photochem. Photobiol. 68 (1998) 191-198. [460] M. Havaux, L. Dall'Osto, R. Bassi, Zeaxanthin has enhanced antioxidant capacity with respect to all other xanthophylls in Arabidopsis leaves and functions independent of binding to PSII antennae, Plant Physiol. 145 (2007) 1506-1520.
ACCEPTED MANUSCRIPT 98
[461] G. Levesque-Tremblay, M. Havaux, F. Ouellet, The chloroplastic lipocalin AtCHL prevents lipid peroxidation and protects Arabidopsis against oxidative stress, Plant J. 60 (2009) 691-702.
RI P
T
[462] N.C. Shantha, E.A. Decker, Rapid, sensitive, iron-based spectrophotometric methods for determination of peroxide values of food lipids, J. AOAC Int. 77 (1994) 421‒424.
NU
SC
[463] J.M. DeLong, R.K. Prange, D.M. Hodges, C.F. Forney, M.C. Bishop, M. Quilliam, Using a modified ferrous oxidation-xylenol orange (FOX) assay for detection of lipid hydroperoxides in plant tissue. J. Agric. Food Chem. 50 (2002) 248‒254. [464] R. Bou, R. Codony, A. Tres, E.A. Decker, F. Guardiola, Determination of hydroperoxides in foods and biological samples by the ferrous oxidation-xylenol orange method: a review of the factors that influence the method's performance, Anal. Biochem. 377 (2008) 1‒15.
MA
[465] C. Gay, J. Collins, J.M. Gebicki, Hydroperoxide assay with the ferric-xylenol orange complex, Anal. Biochem. 273 (1999) 149‒155. [466] J. Nouroozzadeh, J. Tajaddinisarmadi, S.P. Wolff, Measurement of plasma hydroperoxide concentrations by the ferrous oxidation xylenol orange assay in conjunction with triphenylphosphine, Anal. Biochem. 220 (1994) 403‒409.
ED
[467] D. Banerjee, U.K. Madhusoodanan, M. Sharanabasappa,S. Ghosh, J. Jacob, Measurement of plasma hydroperoxide concentration by FOX-1 assay in conjunction with triphenylphosphine, Clin Chim Acta 337 (2003) 147‒152.
PT
[468] D. Hornero-Méndez, A. Pérez-Gálvez, M.-I. Mínguez-Mosquera, A rapid spectrophotometric method for the determination of peroxide value in food lipids with high carotenoid content, J. Am. Oil Chem. Soc 78 (2001) 1151‒1155.
AC CE
[469] W. Jessup, R.T. Dean, J.M. Gebicki, Iodometric determination of hydroperoxides in lipids and proteins, in: J.N. Abelson, M.I. Simon, H. Sies (Eds.), Oxygen Radicals in Biological Systems, Pt C., 233 (Methods in Enzymology) (1994), pp. 289‒303. [470] Y. Watanabe, K. Hashimoto, A. Omori, Y. Uda, M. Nomura, Suppressive ability of defatted rice bran against lipid oxidation in cookies containing iron, Biosci. Biotech. Bioch. 74 (2010) 262‒265. [471] G. Schmitz, The oxidation of iodine to iodate by hydrogen peroxide, Phys. Chem. Chem. Phys. 3 (2001) 4741‒4746. [472] Y. Okimoto, A. Watanabe A., E. Niki, T. Yamashita, N. Noguchi, A novel fluorescent probe diphenyl-1-pyrenylphosphine to follow lipid peroxidation in cell membranes, FEBS Lett. 474 (2000) 137‒140. [473] K. Shioji, Y. Oyama, K. Okuma, H. Nakagawa Synthesis and properties of fluorescence probe for detection of peroxides in mitochondria, Bioorg. Med. Chem. Lett. 20 (2010) 3911-3915. [474] N. Soh, T. Ariyoshi, T. Fukaminato, K. Nakano, M. Irie, T. Imato, Novel fluorescent probe for detecting hydroperoxides with strong emission in the visible range, Bioorg. Med. Chem. Lett. 16 (2006) 2943−2946.
ACCEPTED MANUSCRIPT 99
SC
RI P
T
[475] S.A. Khorobrykh, A.A. Khorobrykh, D.V. Yanykin, B.N. Ivanov, V.V. Klimov, J. Mano, Photoproduction of catalase-insensitive peroxides on the donor side of manganese-depleted photosystem II: evidence with a specific fluorescent probe, Biochemistry 50 (2011) 10658‒10665. [476] K. Yamanaka, Y. Saito, J. Sakiyama, Y. Ohuchi,F. Oseto,N. Noguchi, A novel fluorescent probe with high sensitivity and selective detection of lipid hydroperoxides in cells, RSC Advances 2 (2012) 7894–7900. [477] K. Krumova, L.E. Greene, G. Cosa, Fluorogenic α-tocopherol analogue for monitoring the antioxidant status within the inner mitochondrial membrane of live cells, J. Am. Chem. Soc. 135 (2013) 17135–17143.
MA
NU
[478] D.R. Janero, Malondialdehyde and thiobarbituric acid-reactivity as diagnostic indexes of lipid-peroxidation and peroxidative tissue-injury, Free Radic. Biol. Med. 9 (1990) 515‒540. [479] A. Baryla, C. Laborde, J.L. Montillet, C. Triantaphylidès, P. Chagvardieff, Evaluation of lipid peroxidation as a toxicity bioassay for plants exposed to copper, Environ Pollut. 109 (2000) 131‒135.
ED
[480] C. Jo, D.U. Ahn, Fluorometric analysis of 2-thiobarbituric acid reactive substances in turkey, Poult. Sci. 77 (1998) 475‒480.
AC CE
PT
[481] H. Kosugi, T. Kato, K. Kikugawa, Formation of yellow, orange, and red pigments in the reaction of alk-2-enals with 2-thiobarbituric acid, Anal. Biochem. 165 (1987) 456‒464. [482] D. Lapenna, F. Cuccurullo, TBA test and free MDA assay in evaluation of lipidperoxidation and oxidative stress in tissue systems, Am. J. Physiol. 265 (1993) H1030‒H031. [483] B. Halliwell, S. Chirico, Lipid-peroxidation - its mechanism, measurement, and significance, Am. J. Clin. Nutr. 57 (1993) S715‒S725. [484] J. Tsaknis, S. Lalas, M. Hole, G. Smith, V. Tychopoulos, Rapid high-performance liquid chromatographic method of determining malondialdehyde for evaluation of rancidity in edible oils, Analyst 123 (1998) 325–327. [485] Y.S. Wang, M.D. Ding, X.G. Gu, J.L. Wang, Y. Pang, L.P. Gao, T. Xia, Analysis of interfering substances in the measurement of malondialdehyde content in plant leaves, Am. J. Biochem. Biotechnol. 9 (2013) 235‒242. [486] J. Mano, M. Nagata, S. Okamura, T. Shiraya, T. Mitsui, Identification of oxidatively modified proteins in salt-stressed Arabidopsis: A carbonyl-targeted proteomics approach, Plant Cell Physiol. 55 (2014) 1233–1244. [487] W.H. Ko enol, The reduction otential of the cou le O3/O3•-. Consequences for mechanisms of ozone toxicity, FEBS Lett. 140 (1982) 169‒172. [488] A. Laisk, O. Kull, H. Moldau, Ozone concentration in leaf intercellular air spaces is close to zero, Plant Physiol. 90 (1989) 1163–1167. [489] R. Criegee, Mechanism of ozonolysis, Angew. Chem. Int. Ed. 14 (1975) 745–752. [490] J.T. Herron, R.E. Huie, Rate constants for the reactions of ozone with ethene and propene, from 235.0 to 362.0.deg.K, J. Phys. Chem. 78 (1974) 2085–2088.
ACCEPTED MANUSCRIPT 100
[491] A. Calogirou, B.R. Larsen, D. Kotzias, Gas-phase terpene oxidation products: a review. Atmos. Environ. 33 (1999) 1423–1439.
T
[492] C.D. Forester, J.R. Wells, Hydroxyl radical yields from reactions of terpene mixtures with ozone, Indoor Air 21 (2011) 400‒409.
RI P
[493] J.R. Kanofsky, P. Sima, Singlet oxygen production from the reactions of ozone with biological molecules, J. Biol. Chem. 266 (1991) 9039‒9042.
SC
[494] S. Enami, M.R. Hoffmann, A.J. Colussi, Acidity enhances the formation of a persistent ozonide at aqueous ascorbate/ozone gas interfaces, Proc. Natl. Acad. Sci. USA 105 (2008) 7365–7369.
NU
[495] S. Enami, M.R. Hoffmann, A.J. Colussi, Ozone oxidizes glutathione to a sulfonic acid, Chem. Res. Toxicol. 22 (2009) 35–40.
MA
[496] L.E. Bennett, P. Warlop, Electron transfer to ozone: outer-sphere reactivities of the ozone/ozonide and related non-metal redox couples, Inorg. Chem 29 (1990) 1975–1981. [497] Y. Wang, A. Ries, K.T. Wu, A. Yang, N.M. Crawford, The Arabidopsis prohibitin gene phb3 functions in nitric oxide-mediated responses and in hydrogen peroxideinduced nitric oxide accumulation, Plant Cell, 22 (2010) 249–259.
PT
ED
[498] M. Karlsson, T. Kurz, U.T. Brunk, S.E. Nilsson, C.I. Frennesson, What does the commonly used DCF test for oxidative stress really show? Biochem. J. 428 (2010) 183– 190.
AC CE
[499] S. Burow, G. Valet, Flow-cytometric characterization of stimulation, free radical formation, peroxidase-activity and phagocytosis of human-granulocytes with 2’,7’dichloroflorescein (DCF), Eur. J. Cell Biol. 43 (1987) 128–133. [500] Marchesi, C. Rota, Y.C. Fann, C.F. Chignell, R.P. Mason, Photoreduction of the fluorescent dye 2’-7’-dichlorofluorescein: A spin trapping and direct electron spin resonance study with implications for oxidative stress measurements, Free Rad. Biol. Med. 26 (1999) 148–161. [501] G. Miller, K. Schlaugh, R. Tam, D. Cortes, M.A. Torres, V. Shulaev, J.L. Dangl, R. Mittler, The plant NADPH oxidase RBOHD mediates rapid systemic signaling in response to diverse stimuli, Sci. Signal. 84 (2009) r45. [502] F. Tabet, C. Savoia, E.L. Schiffrin, R.M. Touyz, Differential calcium regulation by hydrogen peroxide and superoxide in vascular smooth muscle cells from spontaneously hypersensitive rats, J. Cardiovasc. Pharmacol. 44 (2004) 200–208. [503] P. Pos íšil, A. Prasad, M. Rác, Role of reactive oxygen s ecies in ultra-weak photon emission in biological systems, J. Photochem. Photobiol. B Biol. 139 (2014) 11–23. [504] A. Nakamura, S. Goto, Analysis of protein carbonyls with 2,4-dinitrophenyl hydrazine and its antibodies by immunoblot in two-dimensional gel electrophoresis, J. Biochem. 119 (1996) 768–774. [505] J. Mano, Reactive carbonyl species: Their production from lipid peroxides, action in environmental stress, and the detoxification mechanism, Plant Physiol. Biochem. 59 (2012) 90–97.
ACCEPTED MANUSCRIPT 101
Figure legends
RI P
T
Fig. 1. A plot of temporally and spectrally resolved luminescence obtained for Chl a in acetone. The peak at 1270 nm, decaying in 40 µs originates from 1O2. Reprinted from [72], Copyright (2003), with permission from Elsevier.
NU
SC
Fig. 2. Effect of TEMP on PSII in intact cells of the diatom Phaeodactylum tricornutum. The decay of Chl a fluorescence yield after a single turnover flash was measured with the FL-200 fluorometer (Photon Systems Instruments, Brno, Czech Republic) from a P. tricornutum culture before (black dashed line) and after 3 min incubation with 10 mM TEMP that had been purified by vacuum distillation (solid red line).
MA
Fig. 3. Demonstration that 30 min illumination of tobacco leaves with low light (LL, PPFD 50 µmol m2 -1 s ) induces hardly detectable quenching of DanePy fluorescence, whereas increase in SOSG fluorescence is obvious after the same treatment. Reprinted from [109] with kind permission from Springer Science and Business Media.
ED
Fig. 4. Stoichiometric production of H2O2 in a reaction between 1O2 and plastoquinol. Reprinted from [60], Copyright (2015), with permission from Elsevier.
PT
Fig. 5. Formation of H2O2 and O2• in cucumber cotyledons in the light, visualized by staining with DAB and with NBT, respectively. Reprinted from [307], Copyright (2014), with permission from Elsevier.
AC CE
Fig. 6. EPR spectra of BMPO/•OOH and BMPO/•OH adducts, as indicated. The dotted lines indicate computer simulations of the respective spectra. Reprinted from [292], Copyright (2001), with permission from Elsevier.
ACCEPTED MANUSCRIPT
SC
RI P
T
102
AC CE
PT
ED
MA
NU
Figure 1
ACCEPTED MANUSCRIPT
NU
SC
RI P
T
103
AC CE
PT
ED
MA
Figure 2
ACCEPTED MANUSCRIPT
NU
SC
RI P
T
104
AC CE
PT
ED
MA
Figure 3
ACCEPTED MANUSCRIPT
PT AC CE
Figure 4
ED
MA
NU
SC
RI P
T
105
ACCEPTED MANUSCRIPT
NU
SC
RI P
T
106
AC CE
PT
ED
MA
Figure 5
ACCEPTED MANUSCRIPT
MA
NU
SC
RI P
T
107
AC CE
PT
ED
Figure 6
ACCEPTED MANUSCRIPT 108
Highlights
AC CE
PT
ED
MA
NU
SC
RI P
T
The review discusses the reactions of Reactive Oxygen Species (ROS) Detection methods of ROS have been critically reviewed Advantages and disadvantages of a large number of chemical detection methods are described The main focus of the review is in the detection or ROS from photosynthetic organisms and tissues