Next Article in Journal
(–)-Loliolide Isolated from Sargassum horneri Suppressed Oxidative Stress and Inflammation by Activating Nrf2/HO-1 Signaling in IFN-γ/TNF-α-Stimulated HaCaT Keratinocytes
Previous Article in Journal
Antioxidant-Based Therapy Reduces Early-Stage Intestinal Ischemia-Reperfusion Injury in Rats
Previous Article in Special Issue
Isoprene: An Antioxidant Itself or a Molecule with Multiple Regulatory Functions in Plants?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

Antioxidant Defenses in Plants: A Dated Topic of Current Interest

by
Lucia Guidi
1 and
Massimiliano Tattini
2,*
1
Department of Agriculture, Food and Environment, University of Pisa, Via del Borghetto 80, I-56124 Pisa, Italy
2
Institute for Sustainable Plant Protection, National Research Council of Italy (CNR), Via Madonna del Piano 10, Sesto F.no, I-50019 Florence, Italy
*
Author to whom correspondence should be addressed.
Antioxidants 2021, 10(6), 855; https://doi.org/10.3390/antiox10060855
Submission received: 20 May 2021 / Accepted: 24 May 2021 / Published: 27 May 2021
(This article belongs to the Special Issue Antioxidant Defenses in Plants)
Plants have been challenged against oxidative stress since their appearance on land [1], and their ability not only to survive but to profitably grow in an ever-changing environment allowed for the conquest of land [2,3]. The evolution of an effective antioxidant defense network (sensu Mittler [4]) has been of key significance for increasing the complexity of land plants [5], in turn allowing for conquest habitats of increasing stress severity [6]. Indeed, current-day land plants are equipped with an arsenal of “primary” and “secondary” antioxidant defenses to finely tune the levels of different reactive oxygen species (ROS) and to effectively acclimate/adapt to a wide array of environmental stressors of different origins [7]. Plants, as a consequence of their sessile nature, have been capable not only of developing an integrated and modular network to effectively reduce ROS within a sub-lethal concentration range, but also of using ROS as well as changes in redox homeostasis as signaling events to activate further defenses against unpredictable oxidative stress events [8,9,10]. This also conforms to most recent views that “ROS are good” [4]. This is the case, for instance, of low molecular weight antioxidants such as ascorbate and glutathione: they provide essential information on the cellular redox state, and they influence gene expression associated with biotic and abiotic stress responses to maximize defense [11]. Further, plants have used and still use some classes of secondary metabolites, the biosynthesis of which arose during the water-to-land transition of plants, to regulate/modulate ABA and auxin signaling, acting on downstream agents (e.g., H2O2 and MAPK) [12,13] of the hormone signaling networks [14,15,16,17,18]. There is intriguing, compelling evidence that transcription factors regulating the biosynthesis of these specialized metabolites are under redox control [19,20]. Consequently, changes in cellular redox homeostasis may activate the biosynthesis of specialized metabolites, which, in turn, offer additional redox modulation, therefore constituting a regulatory circuit of hormone/ROS signaling [14,15]. This is consistent with a model for redox homeostasis in which ROS–antioxidant interactions represent metabolic interfaces for signals derived from metabolism and from the environment, as suggested by Foyer and Noctor [11]. It is not surprising that how plants adjust the suite of antioxidant defenses in response to the wide range of environmental stressors has been and still is one of the timely issues in plant science [21,22,23,24,25,26,27]. Indeed, global change imposes the well-known acclimation/adaptation syndrome of living organisms, which may be particularly severe for plants, which do not possess the ‘flight’ strategy usually displayed by other organisms [28].
The special issue “Antioxidant defenses in plants” is therefore timely and offers contributions that may help readers to focus on most recent aspects of how plants cope with oxidative stress of different origins. As outlined above, our special issue is not centered merely in exploring the antioxidant that arsenal plants activate to limit oxidative stress and damage, but it also focuses on the signaling/functional roles of ROS (and reactive nitrogen species, RNS as well) in response to “environmental stressors” of different origins.
This is exactly the case of the review article by Hasanuzzaman and colleagues [29]. The authors provide a comprehensive review of the chemistry of ROS, their generation and subcellular compartmentation. The authors offer a critical review of both the oxidative stress generated and antioxidant defense activated in response to a wide array of stress agents, including soil salinity, water deficit, metal ions excess, and high and low temperatures. Hasanuzzaman and colleagues [29] also critically explore the role of ROS, especially H2O2 as a signaling molecule, which leads to acclimation processes and confers tolerance under environmental stressors of different origins [27,30,31]. This is consistent with a recent observation that the exogenous application of H2O2 increases tolerance to abiotic stresses [32]. The authors also focus their attention on the crosstalk between ROS and a range of other reactive species, namely reactive nitrogen (RNS), sulphur (RSS) and carbonyl species (RCS), which might act in concert in signal transduction pathways under stressful conditions [33].
The research article by Van Rensburg and Van den Ende [34] is centered on the role of ROS in plants exposed to biotic stresses, and they specifically examine the role of ROS in the responses of Arabidopsis thaliana to pathogen infection, namely to Botrytis cinerea. Actually, the authors explore how γ-aminobutyric acid (GABA) treatment affects plants challenged against pathogens by modulating ROS homeostasis, a timely issue in plant response to biotic stressors [35]. In a very low concentration range, GABA effectively primed resistance against B. cinerea, thereby supporting the author’s idea that GABA signaling was at play, rather than metabolic effects. The view of GABA behaving as a signaling molecule is consistent with its putative role in cellular ROS dynamics. This has a consequential direct impact on plant resistance, since ROS are crucial components of the infection strategy of B. cinerea. Of interest is the finding that pretreatment with GABA keeps low H2O2 levels upon infection, because of enhanced activities of both catalase and glutathione peroxidase, therefore slowing down the infection process [36]. Finally, the authors offer the intriguing hypothesis that GABA signaling might be translated to distal tissues and organs through its constant secretion/uptake between cells (as also occurs in animals) [37,38]: the issue merits further investigation, as it might be involved in systemic acquired resistance [39].
The role of reactive species, namely of RNS, in plant responses to biotic stressors is also the focus of the research paper by Labudda and colleagues [40]. The authors explore the interplay between RNS involved in signal transduction and plant defense against the cyst nematode Heterodera schachtii Schmidt’s infection on Arabidopsis thaliana roots. Labudda and colleagues [40] provide conclusive evidence for RNS production in root cells as a response to cyst nematode attack and parasitism. In turn, RNS, mostly NO, participate in the regulation of the MAPKs (mitogen activated protein kinase) cascade [41]. The authors also hypothesize that H2O2, which is known to activate MAPKs [42,43] (including MAPK6), may have been responsible for the intense NO burst and the downstream signaling pathway in A. thaliana roots upon H. schachtii infestation. The crosstalk between ROS and RNS is crucial in activating defense responses [44], and Lubudda and colleagues [40] report on the potential interplay between the superoxide anion, generated by plasma membrane NADPH oxidase, and NO. They suggest that the formation of peroxinitrite (ONOO-) through the reaction of O2- and NO is central to the response of A. thaliana roots to infection, since ONOO- may freely pass cell membranes and interacts with bio-targets [45]. Therefore, the findings of Labudda et al. [40] offer further support about the central role of NO in root-pathogen interaction, which may be significant to developing novel plant-protection strategies [46].
The roles of reactive oxygen and nitrogen species in plants suffering from abiotic stress agents, ranging from salt/drought stress to high light and nutrient deficiency, have also been indirectly investigated in the review article by Pardo–Hernández and colleagues [47]. In detail, the authors critically review the pertinent literature on the application of melatonin (MET), a tryptophan derivative that has great success, in stress mitigation [48,49]. MET is a key molecule that is functional in regulating the growth, development and reproduction of plants [50]. This mainly results from MET being involved in the regulation of oxidative stress through its effect not only on the network of enzymatic and nonenzymatic defenses but also on the oxidative signaling driven by both ROS and NO [49,51]. There is indeed evidence of MET being involved in the NO-signaling pathway: for instance, NO-induced S-nitrosylation of antioxidant-related proteins [52] is crucial in regulating the antioxidant capacity of the ascorbic acid/glutathione cycle under nitro-oxidative conditions. Nonetheless, Pardo–Hernandez and colleagues [47] conclude that the mechanisms through which MET and NO crosstalk in ameliorating “plant performance” under stressful conditions is far from being conclusively addressed [41].
The ROS- and NO-induced regulation/modulation of gene expression is also the issue investigated by Pucciariello and Perata [53] in their review article. This article represents a curiosity in the context, since it explores the apparent contradiction of the presence and the significance of ROS as well as of RNS (i.e., NO) under low O2 availability [54,55,56]. The authors conduct a comprehensive analysis of studies aimed at unveiling how the availability of ROS and NO, and the consequential adjustments in the antioxidant system, allow for plant adaptation under different levels of O2 availability and water submergence. Pucciariello and Perata [53] conduct their analysis mainly, but not exclusively, by looking at the molecular events that are responsible for the ROS and NO modulation of the expression of specific transcription factors, such as the group VII ethylene responsive factors (ERF-VII), regulating key developmental processes under hypoxia [57]. They conclude that ROS might act as an additional signaling mechanism, acting in parallel to direct O2 sensing. Pucciariello and Perata [53] also highlight the role of ROS production upon post-submergence. In fact, during submergence, turbid water reduces the access of light to underwater organs, while a loss in root hydraulic conductivity may paradoxically lead to dehydration, both phenomena impeding photosynthesis, thereby leading to ROS generation [58]. There is evidence of a key involvement of primary antioxidants in conferring post-submergence tolerance, though the matter still needs investigation.
Our special issue includes, as expected, a number of contributions that explore how the antioxidant defense system is at work in plants suffering from a range of stress agents of abiotic origin. These articles are mostly, but not exclusively, concerned with the roles of secondary metabolites in mitigating stress-induced photooxidative damage in a range of plants.
For instance, in their comprehensive review article, Pollastri and colleagues [59] reason about the “antioxidant” functions of isoprene, a volatile isoprenoid mainly emitted by perennial plants [60,61], the biosynthesis of which is strongly activated both by high temperature and solar irradiance but also because of water deficit [62,63]. The authors report evidence of isoprene being capable of quenching a range of reactive oxygen and nitrogen species, as previously reported in a comprehensive review article by Vickers and colleagues [64]. This suggests a direct antioxidant role for isoprene, but whether isoprene scavenges ROS inside the leaf or in the leaf boundary layer, after exiting leaves through stomata, remains unclear. Pollastri and colleagues [59] report evidence that, with respect to nonemitting plants, isoprene-emitting plants are able to maintain high photosynthetic rates for longer periods when exposed to abiotic stresses, not only because of their ROS quenching capacity but also by improving the thylakoid membrane stability [65,66,67]. In other words, isoprene emitters are well suited to “avoiding”, and not only to countering, photooxidative stress. Further, the authors discuss relatively recent findings showing that isoprene may alter both the proteome and the metabolome in some species, resulting in an enhanced biosynthesis of a range of antioxidants [68,69]. Finally, Pollastri and colleagues [59] discuss recent reports suggesting that isoprene may control the antioxidant system, generally acting as a negative regulator of ROS production in a hormone-like manner. Overall, this is consistent with their hypothesis of isoprene as “a molecule with multiple regulatory functions”, which conforms to the general idea that so-called secondary metabolites, whose biosynthesis is highly consuming for plants in terms of energy and carbon skeletons, serve multiple functional roles in plant-environment interactions. In the vast majority of cases, these functions go beyond the mere ability to quench ROS.
Agati and colleagues [70] discuss the putative antioxidant functions of a relevant class of metabolites synthesized through a branch-pathway of the general phenylpropanoid biosynthetic pathway, specifically the flavonoids. The effective abilities of flavonoids to serve antioxidant functions in an in planta situation is a long-debated question. There are several reasons that the authors suggest are responsible for this continued discussion. First, early experiments provided erroneous beliefs about flavonoids being optimally suited to absorb UV-B wavelengths, whereas flavonoids have molar extinction coefficient maxima at >330 nm (for the review article, see [71,72]). The view about flavonoids being exclusively located in the vacuoles of epidermal cells [73,74] is equally erroneous: flavonoids accumulate in different leaf tissues and subcellular compartments, including the nucleus, the cytoplasm and the chloroplast envelope [75]. Consequently, flavonoids are optimally suited to encounter and scavenge a range of ROS, including H2O2 and singlet oxygen, as conclusively shown in recent experiments [16,17,76]. Second, the view about ROS, particularly H2O2, being “inherently” unable to freely move from its generation sites to other cell organs, where flavonoids mostly accumulate, is erroneous [77,78]. The authors also suggest that the picture representing the complexity of the dynamic cellular organization during the most severe stressful conditions [79,80] needs extensive re-drawing in order to delineate the actual scenario within which biological processes occur. Finally, Agati et al. [70] highlight that an unequivocal definition of antioxidants is mandatory in order to unveil the effective antioxidant capacities of different metabolites, including but not limited to flavonoids. The authors report a few examples of such highly contrasting definitions, ranging from the most wide and hence robust definition given by Gutteridge and Halliwell [81] to the most constraining one proposed by Hernandez et al. [82]. The authors propose to simply define a plant antioxidant in terms of its ability to “counter” oxidative stress and damage, the ability being “quantified” by measurements of “oxidative stress reduction”.
In their research article, Marchica and colleagues [83] also investigate the functional role of phenylpropanoids in the response mechanisms of plants facing abiotic stressors, specifically in response to ozone (O3), a tropospheric pollutant with a high oxidative potential [84,85]. The authors conducted an experiment on Salvia officinalis, looking at both the molecular events regulating the biosynthesis as well as the functional roles of phenylpropanoids in O3-stressed plants. The authors showed that a single pulse of O3 was capable of greatly affecting the phenylpropanoid biosynthetic pathway, leading to an enhanced biosynthesis of a range of phenylpropanoids, especially of rosmarinic acid, a caffeic acid ester synthesized from 3,4-dihydroxyphenyllactic acid through the action of rosmarinic acid synthase (RSA). As such, rosmarinic acid (which accounted for approx. 50% of the entire phenylpropanoid pool) may serve effective antioxidant functions in O3-treated plants, due to the presence of two catechol groups in the benzene rings. Marchica and colleagues [83] also report on the consistent relationship between RAS activity, which increased upon 5 h of O3 fumigation, and the consequential enhanced accumulation of rosmarinic acid, which was observed 24 h after the onset of the O3 treatment. The authors’ data are therefore consistent with the general view that the enhanced biosynthesis and accumulation of phenolic compounds may well constitute a “secondary” antioxidant defense system, complementing the activity of “primary” antioxidants to avoid irreversible oxidative injury in plants under O3 exposure.
In their research article, Kim and colleagues [86] provide evidence of carotenoids serving antioxidant functions in sweet potato plants exposed to severe heat stress (plants were indeed exposed for 12 h at 47 °C). The authors generated transgenic sweet potato plants using the Orange gene, encoding for a protein that behaves as a holdase chaperone, a post-transcriptional regulator of phytoene synthase (PSY), a rate-limiting enzyme in the carotenoid biosynthetic pathway [87]. Kim and colleagues [86] show that transgenic plants are more tolerant to heat stress compared to wild-type plants, based on a lower accumulation of both H2O2 and membrane permeability, which was particularly evident in the plant suffering from the most severe heat stress. The authors attribute the superior heat tolerance to the dramatically higher carotenoid concentration, especially ß-carotene and cryptoxanthin, detected in transgenic plants. This is consistent with the notion of ß-carotene being most suitable to quench singlet oxygen [88], whereas cryptoxanthin (which is generated through the hydroxylation of ß-carotene [89,90]) likely confers a greater stability to thylakoid membranes, in a similar fashion to zeaxanthin [91,92].
The increase of the overall antioxidant defense system is of crucial significance to the profitable growth of a range of annual crop species (which do not constitutively possess the antioxidant arsenal of the vast majority of wild plants) that are frequently exposed to severe environmental pressures imposed by global change [62]. Padilla and colleagues [93] explore the effectiveness of vegetable grafting (a sustainable technique commonly used to enhance plant tolerance to abiotic stressors) in enhancing plant tolerance to drought stress [94]. For this purpose, the authors grafted the pepper (Caspicuum annuum) landrace “Sueca” on two C. annuum hybrids, H90 and H92, which have been previously reported to be sensitive and tolerant to drought stress, respectively. Padilla and colleagues [93] performed measurements on an impressive amount of physiological and biochemical traits to assess the rootstock effect in conferring drought tolerance to a scion/rootstock combination. Actually “Sueca” grafted on H92 displayed a better photosynthetic performance under severe drought, which translated into a lower peroxidation of membrane lipids compared to the Sueca/H90 combination. The two scion/rootstock combinations greatly differed for the biosynthesis and metabolism of ascorbic acid, with Sueca/H92, but not Sueca/H90, showing an increased accumulation of ASA when the drought stress became severe. The authors hypothesize that in addition to well-known ROS scavenging properties of ASA, the ASA-zeaxanthin interaction and the consequential enhancement of nonphotochemical quenching [95,96] were likely responsible for the sustained photosynthesis in Sueca/92 compared to Sueca/90.
This Editorial is in the form of a brief review article. Our efforts have been to provide readers not only with a brief description of each original contribution, but with up-to-date literature concerning the role of oxidative signals as well as of the wide range of “antioxidants” that help plants to cope with environmental stressors of different origins. We hope our efforts will be useful to readers of Antioxidants.

Funding

This research received no external funding.

Acknowledgments

We thank the colleagues who contributed to the special issue as well as the colleagues who critically reviewed the MS. Our special thanks to the antioxidants editorial office.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. de Vries, J.; Archibald, J.M. Plant evolution: Landmarks on the path to terrestrial life. New Phytol. 2018, 217, 1428–1434. [Google Scholar] [CrossRef]
  2. Delwiche, C.F.; Cooper, E.D. The evolutionary origin of a terrestrial flora. Curr. Biol. 2005, 25, R899–R910. [Google Scholar] [CrossRef] [Green Version]
  3. Tena, G. From algae to land plants. Nat. Plants 2020, 6, 594. [Google Scholar] [CrossRef] [PubMed]
  4. Mittler, R. ROS are good. Trends Plant Sci. 2017, 22, 11–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Wagner, A. The molecular origins of evolutionary innovations. Trends Genet. 2011, 27, 397–410. [Google Scholar] [CrossRef]
  6. Fürst-Jansen, J.M.R.; de Vries, S.; de Vries, J. Evo-physio: On stress responses and the earliest land plants. J. Exp. Bot. 2020, 71, 3254–3269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Zandalinas, S.I.; Fritschi, F.B.; Mittler, R. Signal transduction networks during stress combination. J. Exp. Bot. 2020, 71, 1734–1741. [Google Scholar] [CrossRef] [PubMed]
  8. Foyer, C.-H.; Ruban, A.V.; Noctor, G. Viewing oxidative stress through the lens of oxidative signalling rather than damage. Biochem. J. 2017, 474, 877–883. [Google Scholar] [CrossRef] [Green Version]
  9. Noctor, G.; Reichheld, J.P.; Foyer, C.H. ROS-related redox regulation and signaling in plants. Semin. Cell Dev. Biol. 2018, 80, 3–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Zandalinas, S.A.; Mittler, R. ROS-induced ROS release in plant and animal cells. Free Radic. Biol. Med. 2018, 122, 21–27. [Google Scholar] [CrossRef]
  11. Foyer, C.H.; Noctor, C. Redox homeostasis and antioxidant Signaling: A metabolic interface between stress perception and physiological responses. Plant Cell 2005, 17, 1866–1875. [Google Scholar] [CrossRef] [Green Version]
  12. Bigeard, J.; Colcombet, J.; Hirt, H. Signaling mechanisms in pattern-triggered immunity (PTI). Mol. Plant 2015, 8, 521–539. [Google Scholar] [CrossRef] [Green Version]
  13. De Zelicourt, A.; Colcombet, J.; Hirt, H. The role of MAPK modules and ABA during abiotic stress signaling. Trends Plant Sci. 2016, 21, 677–685. [Google Scholar] [CrossRef] [PubMed]
  14. Brunetti, C.; Sebastiani, F.; Tattini, M. Review: ABA, flavonols and the evolvability of land plants. Plant Sci. 2019, 280, 448–454. [Google Scholar] [CrossRef] [PubMed]
  15. Brunetti, C.; Fini, A.; Sebastiani, F.; Gori, A.; Tattini, M. Modulation of phytohormone signaling: A primary function of flavonoids in plant-environment interactions. Front. Plant Sci. 2018, 9, 1042. [Google Scholar] [CrossRef] [Green Version]
  16. Watkins, J.M.; Chapman, J.M.; Muday, G.K. Abscisic acid-induced reactive oxygen species are modulated by flavonols to control stomata aperture. Plant Physiol. 2017, 175, 1807–1824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Watkins, J.M.; Hechler, P.J.; Muday, G.K. Ethylene-induced flavonol accumulation in guard cells suppresses reactive oxygen species and moderates stomatal aperture. Plant Physiol. 2014, 164, 1707–1717. [Google Scholar] [CrossRef] [Green Version]
  18. Peer, W.A.; Cheng, Y.; Murphy, A.S. Evidence of oxidative attenuation of auxin signalling. J. Exp. Bot. 2013, 64, 2629–2639. [Google Scholar] [CrossRef]
  19. Martins, L.; Trujillo-Hernandez, J.A.; Reichheld, J.P. Thiol based redox signaling in plant nucleus. Front. Plant Sci. 2018, 9, 705. [Google Scholar] [CrossRef] [PubMed]
  20. He, H.; van Breusegem, F.; Mahmdi, A. Redox-dependent control of nuclear transcription in plants. J. Exp. Bot. 2018, 69, 3359–3372. [Google Scholar] [CrossRef] [PubMed]
  21. Gill, S.S.; Tuteja, N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiol. Biochem. 2010, 48, 909–930. [Google Scholar] [CrossRef] [PubMed]
  22. Hossain, M.S.; Dietz, K.J. Tuning of redox regulatory mechanisms, reactive oxygen species and redox homeostasis under salinity stress. Front. Plant Sci. 2016, 7, 548. [Google Scholar] [CrossRef] [Green Version]
  23. Foyer, C.H. Reactive oxygen species, oxidative signaling and the regulation of photosynthesis. Environ. Exp. Bot. 2018, 154, 134–142. [Google Scholar] [CrossRef]
  24. Hasanuzzaman, M.; Buyan, M.H.M.B.; Anee, T.I.; Parvin, K.; Nahar, K.; Al Mahmud, J.; Fujita, M. Regulation of ascorbate-glutathione pathway in mitigating oxidative damage in plants under abiotic stress. Antioxidants 2019, 8, 384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Laxa, M.; Liebthal, M.; Welman, T.; Chibani, K.; Dietz, K.J. The role of the plant antioxidant system in drought tolerance. Antioxidants 2019, 8, 94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Foyer, C.H.; Baker, A.; Wright, M.; Sparkes, I.A.; Mhamdi, A.; Schippers, J.H.M.; Breusegem, F.V. On the move: Redox-dependent protein relocation in plants. J. Exp. Bot. 2020, 71, 620–631. [Google Scholar] [CrossRef]
  27. Dvorak, P.; Krasylenko, Y.; Zeiner, A.; Samaj, J.; Takac, T. Signaling toward reactive oxygen species-scavenging enzymes in plants. Front. Plant Sci. 2021, 11, 618835. [Google Scholar] [CrossRef]
  28. Fini, A.; Tattini, M.; Esteban, M. Plants’ responses to novel environmental pressures. Front. Plant Sci. 2017, 8, 2000. [Google Scholar] [CrossRef] [Green Version]
  29. Hasanuzzaman, M.; Buyan, M.H.M.B.; Zulfikar, F.; Raza, A.; Mohsin, S.M.; Al Mahmud, J.; Fujita, M.; Fotopoulos, V. Reactive oxygen species and antioxidant defense in plants under abiotic stress: Revisiting the crucial role of a universal defense regulator. Antioxidants 2020, 9, 681. [Google Scholar] [CrossRef]
  30. Neill, S.; Desikan, R.; Hancock, J. Hydrogen peroxide signalling. Curr. Opin. Plant Biol. 2002, 5, 388–395. [Google Scholar] [CrossRef]
  31. Sies, H. Findings in redox biology: From H2O2 to oxidative stress. J. Biol. Chem. 2020, 295, 13458–13473. [Google Scholar] [CrossRef]
  32. Savvides, A.; Ali, S.; Tester, M.; Fotopoulos, V. Chemical priming of plants against multiple abiotic stresses: Mission possible? Trends Plant Sci. 2016, 21, 329–340. [Google Scholar] [CrossRef] [Green Version]
  33. Biswas, M.S.; Fukaki, H.; Mori, I.C.; Nakahara, K.; Mano, J.I. Reactive oxygen species and reactive carbonyl species constitute a feed-forward loop in auxin signaling for lateral root formation. Plant J. 2019, 100, 536–548. [Google Scholar] [CrossRef] [PubMed]
  34. Van Rensburg, H.C.J.; Van den Ende, W. Priming with ϒ-aminobutyric acid against Botrytis cinerea reshuffles metabolism and reactive oxygen species: Dissecting signalling and metabolism. Antioxidants 2020, 9, 1174. [Google Scholar] [CrossRef]
  35. Tarkowski, L.P.; Signorelli, S.; Höfte, M. γ-Aminobutyric acid and related amino acids in plant immune responses: Emerging mechanisms of action. Plant Cell Environ. 2020, 43, 1103–1116. [Google Scholar] [CrossRef]
  36. Wang, Y.; Gu, W.; Meng, Y.; Xie, T.; Li, L.; Li, J.; Wei, S. γ-Aminobutyric acid imparts partial protection from salt stress injury to maize seedlings by improving photosynthesis and upregulating osmoprotectants and antioxidants. Sci. Rep. 2017, 7, 43609. [Google Scholar] [CrossRef]
  37. Li, L.; Dou, N.; Zhang, H.; Wu, C. The versatile GABA in plants. Plant Signal Behav. 2021, 16, 3. [Google Scholar] [CrossRef]
  38. Xu, B.; Long, Y.; Feng, X.; Zhu, X.; Sai, N.; Chircova, L.; Bets, A.; Herrmann, J.; Edwards, E.J.; Okamoto, M.; et al. GABA signalling modulates stomatal opening to enhance plant water use efficiency and drought resilience. Nat. Commun. 2021, 12, 1952. [Google Scholar] [CrossRef] [PubMed]
  39. Fichman, Y.; Mittler, R. Rapid systemic signaling during abiotic and biotic stresses: Is the ROS wave master of all trades? Plant J. 2020, 102, 887–896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Labudda, M.; Różańska, E.; Gietler, M.; Fidler, J.; Muszyńska, E.; Prabucka, B.; Morkunas, B. Cyst nematode infection elicits alteration in the level of reactive nitrogen species, protein S-nitrosylation and nitration, and nitrosoglutathione reductase in Arabidopsis thaliana roots. Antioxidants 2020, 9, 795. [Google Scholar] [CrossRef] [PubMed]
  41. Zhu, Y.; Gao, H.; Lu, M.X.; Hao, C.Y.; Pu, Z.Q.; Guo, M.J.; Hou, D.R.; Chen, L.Y.; Huang, X. Melatonin-nitric oxide crosstalk and their roles in the redox network in plants. Int. J. Mol. Sci. 2019, 20, 6200. [Google Scholar] [CrossRef] [Green Version]
  42. Jammes, F.; Song, C.; Shin, D.; Munemasa, S.; Takeda, K.; Gu, K.; Cho, D.; Lee, S.; Giordo, R.; Sritubtim, S.; et al. MAPK kinases MPK9 and MPK12 are preferentially expressed in guard cells and positively regulate ROS-mediated ABA signaling. Proc. Natl. Acad. Sci. USA 2009, 106, 20520–20525. [Google Scholar] [CrossRef] [Green Version]
  43. Danquah, A.; de Zelicourt, A.; Colcombet, J.; Hirt, H. The role of ABA and MAPK signaling pathways in plant abiotic stress responses. Biotechnol. Adv. 2013, 32, 40–52. [Google Scholar] [CrossRef]
  44. Zaninotto, F.; La Camera, S.; Polverari, A.; Delledonne, M. Cross talk between reactive nitrogen and oxygen species during the hypersensitive disease resistance response. Plant Physiol. 2006, 141, 379–383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Vandelle, E.; Delledonne, M. Peroxynitrite formation and function in plants. Plant Sci. 2011, 181, 534–539. [Google Scholar] [CrossRef] [PubMed]
  46. Ding, Y.; Gardiner, D.M.; Xiao, D.; Kazan, K. Regulators of nitric oxide signaling triggered by host perception in a plant pathogen. Proc. Natl. Acad. Sci. USA 2020, 117, 11147–11157. [Google Scholar] [CrossRef]
  47. Pardo-Hernandez, M.; Lopéz-Delacalle, M.; Rivero, R.M. ROS and NO regulation by melatonin under abiotic stress in plants. Antioxidants 2020, 9, 1078. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, N.; Sun, Q.; Zhang, H.; Cao, Y.; Weeda, S.; Ren, S.; Guo, Y.-D. Roles of melatonin in abiotic stress resistance in plants. J. Exp. Bot. 2014, 66, 647–656. [Google Scholar] [CrossRef] [Green Version]
  49. Back, K. Melatonin metabolism, signaling and possible roles in plants. Plant J. 2021, 105, 376–391. [Google Scholar] [CrossRef]
  50. Arnao, M.B.; Hernández-Ruiz, J. Is phytomelatonin a new plant hormone? Agronomy 2020, 10, 95. [Google Scholar] [CrossRef] [Green Version]
  51. Sun, C.; Liu, L.; Wang, L.; Li, B.; Jin, C.; Lin, X. Melatonin: A master regulator of plant development and stress responses. J. Integr. Plant Biol. 2020, 63, 126–141. [Google Scholar] [CrossRef] [PubMed]
  52. Begara-Morales, J.C.; Sánchez-Calvo, B.; Chaki, M.; Valderrama, R.; Mata-Pérez, C.; López-Jaramillo, J.; Padilla, M.N.; Carreras, A.; Corpas, F.J.; Barroso, J.B. Dual regulation of cytosolic ascorbate peroxidase (APX) by tyrosine nitration and S-nitrosylation. J. Exp. Bot. 2014, 65, 527–538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Pucciariello, C.; Perata, P. The oxidative paradox in low oxygen stress in plants. Antioxidants 2021, 10, 332. [Google Scholar] [CrossRef] [PubMed]
  54. Paradiso, A.; Caretto, S.; Leone, A.; Bove, A.; Nisi, R.; De Gara, L. ROS production and scavenging under anoxia and re-oxygenation in Arabidopsis cells: A balance between redox signaling and impairment. Front. Plant Sci. 2016, 7, 1803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Pucciariello, C.; Perata, P. New insights into reactive oxygen species and nitric oxide signalling under low oxygen in plants. Plant Cell Environ. 2017, 40, 473–482. [Google Scholar] [CrossRef] [PubMed]
  56. Patel, M.K.; Pandey, S.; Burritt, D.J.; Tran, L.S.P. Plant responses to low-oxygen stress: Interplay between ROS and NO signaling pathways. Environ. Exp. Bot. 2019, 161, 134–142. [Google Scholar] [CrossRef]
  57. Hartmann, S.; Sasidharan, R.; Voesenek, L.A.C.J. The role of ethylene in metabolic acclimations to low oxygen. New Phytol. 2021, 229, 64–70. [Google Scholar] [CrossRef]
  58. Fukao, T.; Yeung, E.; Baily-Serres, J. The submergence tolerance regulator SUB1A mediates crosstalk between submergence and drought tolerance in rice. Plant Cell 2011, 23, 412–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Pollastri, S.; Baccelli, I.; Loreto, F. Isoprene: An antioxidant itself or a molecule with multiple regulatory functions in plants? Antioxidants 2021, 10, 684. [Google Scholar] [CrossRef]
  60. Sharkey, T.D.; Yeh, S. Isoprene emission from plants. Annu. Rev. Plant Biol. 2001, 52, 407–436. [Google Scholar] [CrossRef]
  61. Loreto, F.; Fineschi, S. Reconciling functions and evolution of isoprene emission in higher plants. New Phytol. 2015, 206, 578–582. [Google Scholar] [CrossRef]
  62. Fini, A.; Brunetti, C.; Loreto, F.; Centritto, M.; Ferrini, F.; Tattini, M. Isoprene responses and functions in plants challenged by environmental pressures associated to climate change. Front. Plant Sci. 2017, 8, 1281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Tattini, M.; Loreto, F.; Fini, A.; Guidi, L.; Brunetti, C.; Velikova, V.; Gori, A.; Ferrini, F. Isoprenoids and phenylpropanoids are part of the antioxidant defense orchestrated daily by drought-stressed Platanus × acerifolia plants during Mediterranean summers. New Phytol. 2015, 207, 613–626. [Google Scholar] [CrossRef] [PubMed]
  64. Vickers, C.E.; Gershenzon, J.; Lerdau, M.T.; Loreto, F. A unified mechanism of action for volatile isoprenoids in plant abiotic stress. Nat. Chem. Biol. 2009, 5, 283–291. [Google Scholar] [CrossRef]
  65. Velikova, V.; Varkonyi, Z.; Szabo, M.; Maslenkova, L.; Nogues, I.; Kovacs, L.; Peeva, V.; Busheva, M.; Garab, G.; Sharkey, T.D.; et al. Increased thermostability of thylakoid membranes in isoprene-emitting leaves probed with three biophysical techniques. Plant Physiol. 2011, 157, 905–916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Velikova, V.; Müller, C.; Ghirardo, A.; Rock, T.M.; Aichler, M.; Walch, A.; Schmitt-Kopplin, P.; Schnitzler, J.P. Knocking down of isoprene emission modifies the lipid matrix of thylakoid membranes and influences the chloroplast ultrastructure in poplar. Plant Physiol. 2015, 168, 859–870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Pollastri, S.; Jorba, I.; Hawkins, T.J.; Llusià, J.; Michelozzi, M.; Navajas, D.; Peñuelas, J.; Hussey, P.J.; Knight, M.R.; Loreto, F. Leaves of isoprene-emitting tobacco plants maintain PSII stability at high temperatures. New Phytol. 2019, 223, 1307–1318. [Google Scholar] [CrossRef]
  68. Velikova, V.; Ghirardo, A.; Vanzo, E.; Merl, J.; Hauck, S.M.; Schnitzler, J.P. Genetic manipulation of isoprene emissions in poplar plants remodels the chloroplast proteome. J. Proteome Res. 2014, 13, 2005–2018. [Google Scholar] [CrossRef] [PubMed]
  69. Behnke, K.; Kaiser, A.; Zimmer, I.; Brüggemann, N.; Janz, D.; Polle, A.; Hampp, R.; Hänsch, R.; Popko, J.; Schmitt-Kopplin, P.; et al. RNAi-mediated suppression of isoprene emission in poplar transiently impacts phenolic metabolism under high temperature and high light intensities: A transcriptomic and metabolomic analysis. Plant Mol. Biol. 2010, 74, 61–75. [Google Scholar] [CrossRef] [Green Version]
  70. Agati, G.; Brunetti, C.; Fini, A.; Gori, A.; Guidi, L.; Landi, M.; Sebastiani, F.; Tattini, M. Are flavonoids effective antioxidants in plants? Twenty years of our investigation. Antioxidants 2020, 9, 1098. [Google Scholar] [CrossRef]
  71. Caldwell, M.M.; Robberecht, R.; Flint, S.D. Internal filters: Prospects for UV-acclimation in higher plants. Physiol. Plant. 1983, 58, 445–453. [Google Scholar] [CrossRef]
  72. Cockell, C.S.; Knowland, J. Ultraviolet radiation screening compounds. Biol. Rev. 1999, 74, 311–345. [Google Scholar] [CrossRef] [PubMed]
  73. Hrazdina, G.; Marx, G.A.; Hoch, H.C. Distribution of secondary plant metabolites and their biosynthetic enzymes in pea (Pisum sativum L.) leaves. Anthocyanins and flavonol glycosides. Plant Physiol. 1982, 70, 745–748. [Google Scholar] [CrossRef] [Green Version]
  74. Schmelzer, E.; Jahnen, W.; Hahlbrock, K. In situ localization of light-induced chalcone synthase mRNA, chalcone synthase, and flavonoid end products in epidermal cells of parsley leaves. Proc. Natl. Acad. Sci. USA 1988, 85, 2989–2993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Agati, G.; Azzarello, E.; Pollastri, S.; Tattini, M. Flavonoids as antioxidants in plants: Location and functional significance. Plant Sci. 2012, 196, 67–76. [Google Scholar] [CrossRef]
  76. Agati, G.; Matteini, P.; Goti, A.; Tattini, M. Chloroplast-located flavonoids can scavenge singlet oxygen. New Phytol. 2007, 174, 77–89. [Google Scholar] [CrossRef] [PubMed]
  77. Wang, H.; Schoebel, S.; Schmitz, F.; Dong, H.; Hedfalk, K. Characterization of aquaporin-driven hydrogen peroxide transport. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183065. [Google Scholar] [CrossRef]
  78. Saijo, Y.; Lu, E.P.-I. Plant immunity in signal integration between biotic and abiotic stress responses. New Phytol. 2020, 255, 87–104. [Google Scholar] [CrossRef] [Green Version]
  79. Liao, C.-Y.; Bassham, D.C. Combating stress: The interplay between hormone signaling and autophagy in plants. J. Exp. Bot. 2020, 71, 1723–1733. [Google Scholar] [CrossRef]
  80. Kikuchi, Y.; Nakamura, S.; Woodson, J.D.; Ishida, H.; Ling, Q.; Hidema, Y.; Jarvis, R.P.; Hagihara, S.; Izumi, M. Chloroplast autophagy and ubiquitination combine to manage oxidative damage and starvation responses. Plant Physiol. 2020, 183, 1531–1544. [Google Scholar] [CrossRef] [PubMed]
  81. Halliwell, B.; Gutteridge, J.M.C. Free Radicals in Biology and Medicine, 4th ed.; Oxford University Press: Oxford, UK, 2007. [Google Scholar]
  82. Hernández, I.; Alegre, L.; van Breusegem, F.; Munné-Bosch, S. How relevant are flavonoids as antioxidants in plants? Trends Plant Sci. 2009, 14, 125–132. [Google Scholar] [CrossRef] [PubMed]
  83. Marchica, A.; Cotrozzi, L.; Detti, R.; Lorenzini, G.; Pellegrini, E.; Petersen, M.; Nali, C. The biosynthesis of phenolic compounds is an integrated defence mechanism to prevent ozone injury in Salvia officinalis. Antioxidants 2020, 9, 1274. [Google Scholar] [CrossRef]
  84. Grulke, N.E.; Heath, R.L. Ozone effects on plants in natural ecosystems. Plant Biol. 2020, 22, 12–37. [Google Scholar] [CrossRef] [PubMed]
  85. Hoshika, Y.; Fares, S.; Pellegrini, E.; Conte, A.; Paoletti, E. Water use strategy affects avoidance of ozone stress by stomatal closure in Mediterranean trees—A modelling analysis. Plant Cell Environ. 2020, 43, 611–623. [Google Scholar] [CrossRef] [PubMed]
  86. Kim, S.E.; Lee, C.J.; Park, S.U.; Lim, Y.H.; Park, W.S.; Kim, H.J.; Ahn, M.J.; Kwak, S.S.; Kim, H.S. Overexpression of the golden SNP-carrying orange gene enhances carotenoid accumulation and heat stress tolerance in sweetpotato plants. Antioxidants 2021, 10, 51. [Google Scholar] [CrossRef] [PubMed]
  87. Yuan, H.; Owsiany, K.; Sheeja, T.; Zhou, X.; Rodriguez, C.; Li, Y.; Welsch, R.; Chayut, N.; Yang, Y.; Thannhauser, T.W.; et al. A single amino acid substitution in an ORANGE protein promotes carotenoid over accumulation in Arabidopsis. Plant Physiol. 2015, 169, 421–431. [Google Scholar] [CrossRef] [Green Version]
  88. D’Alessandro, S.; Havaux, M. Sensing β-carotene oxidation in photosystem II to master plant stress tolerance. New Phytol. 2019, 223, 1776–1783. [Google Scholar] [CrossRef]
  89. Li, Q.; Farre, G.; Naqvi, S.; Breitenbach, J.; Sanahuja, G.; Bai, C.; Sandmann, G.; Capell, T.; Christou, P.; Zhu, C. Cloning and functional characterization of the maize carotenoid isomerase and β-carotene hydroxylase genes and their regulation during endosperm maturation. Transgenic Res. 2010, 19, 1063–1068. [Google Scholar] [CrossRef]
  90. Yuan, H.; Zhang, J.; Nageswaran, D.; Li, L. Carotenoid metabolism and regulation in horticultural crops. Hortic. Res. 2015, 2, 15036. [Google Scholar] [CrossRef] [Green Version]
  91. Beckett, M.; Loreto, F.; Velikova, V.; Brunetti, C.; Di Ferdinando, M.; Tattini, M.; Calfapietra, C.; Farrant, J.M. Photosynthetic limitations and volatile and non-volatile isoprenoids in the poikilochlorophyllous resurrection plant Xerophyta humilis during dehydration and rehydration. Plant Cell Environ. 2012, 35, 20161–22074. [Google Scholar] [CrossRef] [PubMed]
  92. Dell’Aglio, E. Carotenoid composition affects thylakoid morphology and membrane fluidity. Plant Physiol. 2021, 185, 21–22. [Google Scholar] [CrossRef] [PubMed]
  93. Padilla, Y.G.; Gisbert-Mullor, R.; Lopez-Serrano, L.; Lopez-Galarza, S.; Calatayud, A. Grafting enhances pepper water stress tolerance by improving photosynthesis and antioxidant defense systems. Antioxidants 2021, 10, 576. [Google Scholar] [CrossRef] [PubMed]
  94. Goldschmidt, E.E. Plant grafting: New mechanisms, evolutionary implications. Front. Plant Sci. 2014, 5, 727. [Google Scholar] [CrossRef] [Green Version]
  95. Müller-Mulé, P.; Conklin, P.L.; Niyogi, K.K. Ascorbate deficiency can limit violaxanthin de-epoxidase activity in vivo. Plant Physiol. 2002, 128, 970–977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Müller-Mulé, P.; Golan, T.; Niyogi, K.K. Ascorbate-deficient mutants of Arabidopsis grow in high light despite chronic photooxidative stress. Plant Physiol. 2004, 134, 1163–1172. [Google Scholar] [CrossRef] [Green Version]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guidi, L.; Tattini, M. Antioxidant Defenses in Plants: A Dated Topic of Current Interest. Antioxidants 2021, 10, 855. https://doi.org/10.3390/antiox10060855

AMA Style

Guidi L, Tattini M. Antioxidant Defenses in Plants: A Dated Topic of Current Interest. Antioxidants. 2021; 10(6):855. https://doi.org/10.3390/antiox10060855

Chicago/Turabian Style

Guidi, Lucia, and Massimiliano Tattini. 2021. "Antioxidant Defenses in Plants: A Dated Topic of Current Interest" Antioxidants 10, no. 6: 855. https://doi.org/10.3390/antiox10060855

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop