Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2178 2023-05-24 08:13:59 |
2 format correct Meta information modification 2178 2023-05-24 08:26:40 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Hernández-Ruiz, J.; Giraldo-Acosta, M.; Mihyaoui, A.E.; Cano, A.; Arnao, M.B. Melatonin in Plant Biocontrol. Encyclopedia. Available online: https://encyclopedia.pub/entry/44749 (accessed on 27 July 2024).
Hernández-Ruiz J, Giraldo-Acosta M, Mihyaoui AE, Cano A, Arnao MB. Melatonin in Plant Biocontrol. Encyclopedia. Available at: https://encyclopedia.pub/entry/44749. Accessed July 27, 2024.
Hernández-Ruiz, Josefa, Manuela Giraldo-Acosta, Amina El Mihyaoui, Antonio Cano, Marino B. Arnao. "Melatonin in Plant Biocontrol" Encyclopedia, https://encyclopedia.pub/entry/44749 (accessed July 27, 2024).
Hernández-Ruiz, J., Giraldo-Acosta, M., Mihyaoui, A.E., Cano, A., & Arnao, M.B. (2023, May 24). Melatonin in Plant Biocontrol. In Encyclopedia. https://encyclopedia.pub/entry/44749
Hernández-Ruiz, Josefa, et al. "Melatonin in Plant Biocontrol." Encyclopedia. Web. 24 May, 2023.
Melatonin in Plant Biocontrol
Edit

Melatonin is a multifunctional and ubiquitous molecule. In animals, melatonin is a hormone that is involved in a wide range of physiological activities and is also an excellent antioxidant. In plants, it has been considered a master regulator of multiple physiological processes as well as of hormonal homeostasis.

melatonin resistance disease virus plant biocontrol plant pathogen

1. Introduction

Melatonin (N-acetyl-5-methoxytryptamine) is an indoleamine derived from tryptophan and produced by several organisms such as animals [1][2][3][4][5][6][7], bacteria [8][9][10][11], fungi [7][12] and plants [7][10][13]. In mammals, it was discovered in 1958 in the pineal gland where it is synthesized [14]. Since then, its chemical structure has been described [15]. Later, in 1995, melatonin was identified in several plants and so-called phytomelatonin [16][17][18].
In animals, melatonin is involved in the regulation of circadian rhythms, mood, locomotor activity, retina physiology and the seasonal behavior, intervening in processes such as reproductive behavior, sleep, food intake, etc. [19][20][21][22][23]. Melatonin regulates circadian rhythms by acting as a light/darkness signal and sending information to the central nervous system to synchronize physiological processes [24][25]. Moreover, at the cellular level, melatonin acts in many cell metabolic processes, from the scavenging of reactive species of oxygen and nitrogen (ROS and RNS) and activating antioxidative enzymes to being a great antioxidant molecule [26][27][28]. Additionally, melatonin seems to intervene in processes related to the immune system since it can be influenced by light signals through the neuroendocrine. In this way, melatonin acts as an anti-inflammatory, antioxidant and neuroprotective agent in many diseases such as neurodegenerative (dementia, Alzheimer, Parkinson), cardiovascular, obesity, cancer and other dysfunctions [29][30][31][32][33][34][35][36][37][38]. In addition, melatonin has been used to combat animal and human viral diseases, such as Venezuelan equine encephalomyelitis (VEE) virus [39] and Ebola virus diseases [40]. On the other hand and as a consequence of the global pandemic caused by SARS-CoV-2, melatonin has been used in various studies as an active or adjuvant drug for use at different stages of therapy for COVID-19 [41]. Melatonin is not virucidal but it has indirect antiviral actions due to its anti-inflammatory, antioxidant and immune enhancing properties [42].

2. Biosynthetic Pathway of Melatonin

Tryptophan, an amino acid that is synthesized from chorismic acid in plants, is the origin of the melatonin biosynthesis pathway [4][43][44] in animal and plant cells, but differs in some steps. In animal cells, tryptophan is converted to 5-hydroxytryptophan by tryptophan hydroxylase (TPH), an enzyme that apparently has not been identified in plants [45][46].
In plants, tryptophan is converted to tryptamine by the enzyme tryptophan decarboxylase (TDC) (Figure 1). Tryptamine is then converted into 5-hydroxytryptamine (serotonin) by tryptamine 5-hydroxylase (T5H), an enzyme that has been extensively studied in rice, and which could act on many substrates. Serotonin was N-acetylated by serotonin N-acetyltransferase (SNAT). N-acetylserotonin is then methylated by acetylserotonin methyl transferase (ASMT), a hydroxyindole-O-methyltransferase that generates melatonin. In plants, methylation of N-acetylserotonin can also be performed by a caffeic acid O-methyltransferase (COMT), a class of enzyme that can act on a variety of substrates, including caffeic acid and quercetin [47]. In plants, serotonin may also be transformed into 5-methoxytryptamine by ASMT/COMT to generate melatonin after the action of SNAT [48]. This route would occur in senescence and/or stress situations [46][49]. Figure 1 shows the steps of the biosynthesis of melatonin in animals (mainly mammals) and plants.
Figure 1. Pathways of melatonin biosynthesis in plants (green arrows) and animals (blue arrows). Enzyme abbreviations: TDC = tryptophan decarboxylase; TPH = tryptophan hydroxylase; T5H = tryptamine 5-hydroxylase; SNAT = serotonin N-acetyltransferase; ASMT = acetylserotonin methyl transferase; COMT = caffeic acid O-methyltransferase.

3. Physiological Roles of Phytomelatonin

The functions of phytomelatonin in plants have been widely studied, observing that phytomelatonin is a protective biomolecule that activates tolerance and resistance responses in plants, playing an important role against biotic (bacteria, viruses, fungi, insects, parasitic nematodes and weeds) and abiotic stressors (salinity, drought, waterlogging, UV-radiation, heavy metal, heat, cold, mineral deficit/excess, and pesticides) [50][51][52].
Melatonin/phytomelatonin plays an important role in plants as a large antioxidant molecule, similar to how it occurs in animal cells, generally acting as an excellent free radical scavenger, specifically on ROS and RNS [26][28][53]. The direct antioxidant activity of melatonin that neutralizes several ROS/RNS and other free radical species that can damage cells was demonstrated in several in vitro and in vivo models [54][55][56][57]. It also activates the antioxidant redox response, upregulating various transcription factors that trigger the expression of antioxidant enzymes such as superoxide dismutases, catalases, peroxidases, and those involved in the ascorbate-glutathione cycle, among others [28][44][58].
Phytomelatonin is considered a plant master regulator because it regulates the levels and actions of plant hormones. The endogenous levels of auxin, gibberellins (GA), cytokinins (CK), abscisic acid (ABA), ethylene (ET) and other phytohormones such as brassinosteroids, jasmonates (JA) and salicylates (SA) are affected by the action of phytomelatonin through up- or downregulation of transcripts of some biosynthesis/catabolism enzymes, and also hormone-related regulatory factors [51][59][60].
On the other hand, it should be noted that photosynthesis, photorespiration, and stomatal regulation, which are key pieces of the water and carbon economy in plants, are strongly regulated by phytomelatonin. Moreover, the metabolic pathways of carbohydrates, lipids and nitrogen and sulfur compounds are modulated by phytomelatonin, including the osmoregulatory response in stressful situations. In secondary metabolism, phytomelatonin induces the biosynthesis of flavonoids, anthocyanins and carotenoids, among other molecules [61][62][63].
Phytomelatonin also promotes the rooting process for primary and adventitious roots [64], and it regulates leaf senescence, delaying it [65]. In fruit post-harvest, phytomelatonin regulates ethylene and lycopene content, as well as cell wall-related enzymes and primary and secondary metabolism. It also helps preserve cut flowers, delaying senescence, and induces parthenocarpy in fruiting [66].
Due to the ability of phytomelatonin to protect plants against stresses, this molecule could be used as a safener in crops. This agronomical application consists of using phytomelatonin in combination with pesticides, not only to increase the plant tolerance to the possible stress caused by the pesticide (abiotic stress) but also to increase the pesticide efficacy against the biotic stressor [67][68][69][70]. Another application of the effect of phytomelatonin under abiotic stresses is its use postharvest. Phytomelatonin reduces the effect of cold storage of vegetables and fruits by minimizing the damage caused by ROS, improving the quality and commercial shelf-life of fruits and vegetables [71][72][73][74].
Lastly, its role in bacterial, fungal, and viral pathogenic infections should be emphasized. Phytomelatonin slows damage and stimulates systemic acquired resistance (SAR), which contributes to increasing both crop health and postharvest health quality [66][73][75][76][77].

4. Melatonin in Plant Disease Biocontrol

There are different approaches that can be used to prevent, mitigate, or control plant diseases. One of them is the application of chemical pesticides, which are used both for preventive as well as for curative disease management [78][79]. However, there is a concern about the negative effects of chemical pesticides due their possible harmful effects on human health, the environment, as well as their effect on the promotion of new resistant pathogens [80][81][82].
There is increasingly stricter legislation in relation to the accessibility and use of efficient pesticides and, therefore, their use is currently declining [83]. Consequently, one study has focused its efforts on developing alternatives to synthetic chemicals for pests and diseases control, some of these being alternatives to so-called biological control.
The term biocontrol refers to the use of naturally occurring (micro-)organisms to control plant diseases or pests [84]. The organism that suppresses the pest or pathogen is found widely in nature, including bacteria, fungi, viruses, yeasts, and protozoans. It can control plant diseases directly, which can be achieved through parasitism, antibiosis, or competition for nutrients or infection sites, or indirectly, where the biocontrol organism induces plant-mediated responses allowing the plant to react faster and more efficiently in case of subsequent pathogen attack [85].
The induced resistance strategy in plants can be indirectly carried out, not only through an organism but also using elicitors; that is, a natural molecule that mimics a pathogen attack or a state of danger, also by living organisms. Plant-induced resistance may represent an interesting strategy for crops [86].
Considering this, melatonin can be an excellent candidate to be used in biocontrol treatments as elicitor molecule. There are currently several studies conducted on the effect of melatonin treatment in the control or reduction of infectious diseases in plants, such as those caused by fungi, bacteria, and viruses.
Plant viruses produce local lesions and also cause systemic damage, resulting in malformations, stunting, and chlorosis in plant tissues, even if their hosts are often biotrophic pathogens [77][87]. Viral diseases in plants cause severe economic losses due to agricultural production and have hindered sustainable agricultural development globally for a long time. Unlike diseases induced by fungi and bacteria, viral infections are more difficult to control once the plants are infected [88][89].
In recent decades, various strategies have been developed to control viral diseases, including the breeding of virus-resistant/tolerant cultivars by conventional breeding techniques [90] and the use of eco-friendly chemical compounds that induce systemic resistance [91].
Chemical priming may be considered a timely and successful management technique to induce resistance/tolerance to viruses of plants. Several eco-friendly compounds that are considered non-toxic, biodegradable, and also biocompatible oligomers, such as proteins, polysaccharides and small molecules (alkaloids, flavonoids, phenolics, essential oils) from plants, proteins and polysaccharides from microorganisms, polysaccharides from algae and oligochitosan from animals, can be used to induce plant resistance to viruses [92][93].
Concerning virus infection in plants, there are a few reports (since the first studies were conducted in 2019) based on the study of the interaction of plant viruses with melatonin and its possible role as an inducer of viral resistance in plants. One of these studies was conducted by Chen and cols. (2019) on the potential role of melatonin in the eradication of Apple Stem Grooving Virus (ASGV) infection in in vitro Gala apple cultivars [94]. Apple is generally propagated vegetatively by grafting, resulting in transmission of the virus from generation to generation; therefore, it is important to obtain virus-free apple planting materials. Apple shoot segments excised from ASVG virus-infected shoots were cultured in various media supplemented with 0, 10, 15, or 20 µM melatonin and were maintained at 22 °C under a 16 h photoperiod. Ten samples were included in each treatment of three replicates. Treatments of 15 µM melatonin were the most efficient in promoting the number and length of the shoots, as well as the high level of endogenous hormone indole-3-acetic acid. On the other hand, in vitro culture of the virus-infected shoots tips in the medium with 15 µM melatonin resulted in 95% of these shoots being virus-free, while no virus-free shoots were obtained in shoot tips of the virus infected shoots cultured without melatonin. In addition, those plants that continued to be infected, even with 15 µM melatonin in the medium, showed a lower viral load than infected plants grown without melatonin. The virus localization showed that exogenous application of melatonin enlarged the virus-free area of the virus-infected shoot tips. The authors concluded that the exogenous application of melatonin can efficiently eradicate ASGV, being the frequency of the virus eradication related to the melatonin concentration used and the culture time duration on melatonin-containing shoot proliferation medium. Inclusion of 15 µM melatonin in the medium to proliferate shoots for 4 weeks followed by shoot tip culture was found to efficiently eradicate ASGV. This procedure produced 100% of survival and 85% of shoot regrowth levels, and also 95% of virus-free plants in shoot tip culture. The application of melatonin treatments may provide an alternative means for the eradication of plant viruses and could even be used to produce virus-free plants as an interesting biotechnological approach.
Resistance to Rice Stripe Virus (RSV) in rice plants treated with melatonin has been studied [95]. The optimum concentration of melatonin and SNP (sodium nitroprusside used as a nitric oxide (NO)-releasing reagent) to reduce disease incidence in the RSV-suscetible Nipponbare rice cultivar has been screened. The soil of the 14-day-old rice seedling pots was kept as dry as possible, followed by the application of the different treatments, 0.1, 1, 10, and 100 μM of melatonin or 10, 50, 100, 500 and 1000 µM of SNP. Rice seedlings are placed in the dark for 12 h. After that, they were inoculated with RSV for 3 days. The plants were then transferred to the soil in the greenhouse. Thirty plants were used for each treatment. The results showed that both (melatonin and SNP) can reduce disease incidence in a concentration dependent manner, with the largest effect being observed with 10 μM melatonin and 100 μM SNP. Therefore, the application of exogenous melatonin and NO can promote rice resistance to RSV. Additionally, both compounds positively modulated the expression of two genes (OsPR1b and OsWRKY45 involved in the induction of PR genes (pathogenesis related protein)), indicating that melatonin and NO are able to enhance the plant disease-resistance genes in RSV disease. So, Lu et al. (2019) also quantified the endogenous melatonin levels in two rice cultivars, Nipponbare and Zhendao-88 (susceptible and resistant cultivars to RSV, respectively) after RSV infection. The data showed that resistance to RSV was improved by increased endogenous melatonin and NO production, and established that melatonin was responsible for rice resistance to RSV infection by inducing NO. The authors conclude that rice resistance to RSV can be improved by increasing melatonin through a NO-dependent pathway. The authors postulate that increased melatonin in the resistant cultivar Zhendao-88 could lead to more NO, which might lead to more SA, which may be the explanation for the increased resistance of this cultivar to RSV [95].

References

  1. Wright, M.L. Melatonin, Diel Rhythms, and Metamorphosis in Anuran Amphibians. Gen. Comp. Endocrinol. 2002, 126, 251–254.
  2. Zeman, M.; Herichová, I. Circadian Melatonin Production Develops Faster in Birds than in Mammals. Gen. Comp. Endocrinol. 2011, 172, 23–30.
  3. Claustrat, B.; Leston, J. Melatonin: Physiological Effects in Humans. Neurochirurgie 2015, 61, 77–84.
  4. Zhao, D.; Yu, Y.; Shen, Y.; Liu, Q.; Zhao, Z.; Sharma, R.; Reiter, R.J. Melatonin Synthesis and Function: Evolutionary History in Animals and Plants. Front. Endocrinol. 2019, 10, 249.
  5. Saha, S.; Singh, K.M.; Gupta, B.B.P. Melatonin Synthesis and Clock Gene Regulation in the Pineal Organ of Teleost Fish Compared to Mammals: Similarities and Differences. Gen. Comp. Endocrinol. 2019, 279, 27–34.
  6. Singh, A.; Singh, R.; Tripathi, M.K. Photoperiodic Manipulation Modulates the Innate and Cell Mediated Immune Functions in the Fresh Water Snake, Natrix Piscator. Sci. Rep. 2020, 10, 14722.
  7. Cheng, G.; Ma, T.; Deng, Z.; Gutiérrez-Gamboa, G.; Ge, Q.; Xu, P.; Zhang, Q.; Zhang, J.; Meng, J.; Reiter, R.J.; et al. Plant-Derived Melatonin from Food: A Gift of Nature. Food Funct. 2021, 12, 2829–2849.
  8. Tilden, A.R.; Becker, M.A.; Amma, L.L.; Arciniega, J.; McGaw, A.K. Melatonin Production in an Aerobic Photosynthetic Bacterium: An Evolutionarily Early Association with Darkness. J. Pineal Res. 1997, 22, 102–106.
  9. Rodriguez-Naranjo, M.I.; Torija, M.J.; Mas, A.; Cantos-Villar, E.; Garcia-Parrilla, M.d.C. Production of Melatonin by Saccharomyces Strains under Growth and Fermentation Conditions. J. Pineal Res. 2012, 53, 219–224.
  10. Hardeland, R. Melatonin in the Evolution of Plants and Other Phototrophs. Melatonin Res. 2019, 2, 10–36.
  11. Que, Z.; Ma, T.; Shang, Y.; Ge, Q.; Zhang, Q.; Xu, P.; Zhang, J.; Francoise, U.; Liu, X.; Sun, X. Microorganisms: Producers of Melatonin in Fermented Foods and Beverages. J. Agric. Food Chem. 2020, 68, 4799–4811.
  12. Rehman, R.S.; Hussain, M.; Ali, M.; Zafar, S.A.; Pasha, A.N.; Bashir, H.; Ashraf, N.A.; Javed, A.; Shah, W.A. A Comprehensive Review on Melatonin Compound and Its Functions in Different Fungi and Plants. Int. J. Pathog. Res. 2022, 10, 9–21.
  13. Arnao, M.B. Phytomelatonin: Discovery, Content, and Role in Plants. Adv. Bot. 2014, 2014, e815769.
  14. Lerner, A.B.; Case, J.D.; Takahashi, Y.; Lee, T.H.; Mori, W. Isolation of Melatonin, the Pineal Gland Factor That Lightens Melanocytes 1. J. Am. Chem. Soc. 1958, 80, 2587.
  15. Lerner, A.B.; Case, J.D.; Heinzelman, R.V. Structure of Melatonin. J. Am. Chem. Soc. 1959, 81, 6084–6085.
  16. Kolar, J.; Machackova, I.; Illnerova, H.; Prinsen, E.; van Dongen, W.; van Onckelen, H. Melatonin in Higher Plant Determined by Radioimmunoassay and Liquid Chromatography-Mass Spectrometry. Biol. Rhythm Res. 1995, 26, 406–409.
  17. Dubbels, R.; Reiter, R.J.; Klenke, E.; Goebel, A.; Schnakenberg, E.; Ehlers, C.; Schiwara, H.W.; Schloot, W. Melatonin in Edible Plants Identified by Radioimmunoassay and by High Performance Liquid Chromatography-Mass Spectrometry. J. Pineal Res. 1995, 18, 28–31.
  18. Hattori, A.; Migitaka, H.; Iigo, M.; Itoh, M.; Yamamoto, K.; Ohtani-Kaneko, R.; Hara, M.; Suzuki, T.; Reiter, R.J. Identification of Melatonin in Plants and Its Effects on Plasma Melatonin Levels and Binding to Melatonin Receptors in Vertebrates. Biochem. Mol. Biol. Int. 1995, 35, 627–634.
  19. Xie, Z.; Chen, F.; Li, W.A.; Geng, X.; Li, C.; Meng, X.; Feng, Y.; Liu, W.; Yu, F. A Review of Sleep Disorders and Melatonin. Neurol. Res. 2017, 39, 559–565.
  20. Vadnie, C.A.; McClung, C.A. Circadian Rhythm Disturbances in Mood Disorders: Insights into the Role of the Suprachiasmatic Nucleus. Neural Plast. 2017, 2017, e1504507.
  21. Majidinia, M.; Reiter, R.J.; Shakouri, S.K.; Yousefi, B. The Role of Melatonin, a Multitasking Molecule, in Retarding the Processes of Ageing. Ageing Res. Rev. 2018, 47, 198–213.
  22. Blume, C.; Angerer, M.; Raml, M.; del Giudice, R.; Santhi, N.; Pichler, G.; Kunz, A.B.; Scarpatetti, M.; Trinka, E.; Schabus, M. Healthier Rhythm, Healthier Brain? Integrity of Circadian Melatonin and Temperature Rhythms Relates to the Clinical State of Brain-Injured Patients. Eur. J. Neurol. 2019, 26, 1051–1059.
  23. Socaciu, A.I.; Ionut, R.; Socaciu, M.A.; Ungur, A.P.; Bârsan, M.; Chiorean, A.; Socaciu, C.; Râjnoveanu, A.G. Melatonin, an Ubiquitous Metabolic Regulator: Functions, Mechanisms and Effects on Circadian Disruption and Degenerative Diseases. Rev. Endocr. Metab. Disord. 2020, 21, 465–478.
  24. Jan, J.E.; Reiter, R.J.; Wasdell, M.B.; Bax, M. The Role of the Thalamus in Sleep, Pineal Melatonin Production, and Circadian Rhythm Sleep Disorders. J. Pineal Res. 2009, 46, 1–7.
  25. Luo, C.; Yang, Q.; Liu, Y.; Zhou, S.; Jiang, J.; Reiter, R.J.; Bhattacharya, P.; Cui, Y.; Yang, H.; Ma, H.; et al. The Multiple Protective Roles and Molecular Mechanisms of Melatonin and Its Precursor N-Acetylserotonin in Targeting Brain Injury and Liver Damage and in Maintaining Bone Health. Free Radic. Biol. Med. 2019, 130, 215–233.
  26. Reiter, R.J.; Poeggeler, B.; Tan, D.X.; Chen, L.; Manchester, L.; Guerrero, J. Antioxidant Capacity of Melatonin. A Novel Action Not Requiring a Receptor. Neuroendocrinol. Lett. 1993, 15, 103–116.
  27. Tan, D.X.; Poeggeler, B.; Reiter, R.J.; Chen, L.D.; Chen, S.; Lucien, M.C.; Barlow-Walden, L.R. The Pineal Hormone Melatonin Inhibits DNA-Adduct Formation Induced by the Chemical Carcinogen Safrole in Vivo. Cancer Lett. 1993, 70, 65–71.
  28. Arnao, M.B.; Hernández-Ruiz, J. Melatonin and Reactive Oxygen and Nitrogen Species: A Model for the Plant Redox Network. Melatonin Res. 2019, 2, 152–168.
  29. Srinivasan, V.; Pandi-Perumal, S.R.; Maestroni, G.J.M.; Esquifino, A.I.; Hardeland, R.; Cardinali, D.P. Role of Melatonin in Neurodegenerative Diseases. Neurotoxicol. Res. 2005, 7, 293–318.
  30. Miller, S.C.; Pandi-Perumal, S.R.; Esquifino, A.I.; Cardinali, D.P.; Maestroni, G.J.M. The Role of Melatonin in Inmmuno-Enhancement: Potential Application in Cancer. Int. J. Exp. Pathol. 2006, 87, 81–87.
  31. Cardinali, D.P.; Hardeland, R. Inflammaging, Metabolic Syndrome and Melatonin: A Call for Treatment Studies. Neuroendocrinology 2017, 104, 382–397.
  32. Su, S.C.; Hsieh, M.J.; Yang, W.E.; Chung, W.H.; Reiter, R.J.; Yang, S.F. Cancer Metastasis: Mechanisms of Inhibition by Melatonin. J. Pineal Res. 2017, 62, e12370.
  33. Cardinali, D.P. Melatonin and Healthy Aging. In Vitamins and Hormones Hormones and Aging; Litwack, G., Ed.; Academic Press: Cambridge, MA, USA, 2021; Volume 115, pp. 67–88. ISBN 0083-6729.
  34. Malabadi, R.B.; Kolkar, K.P.; Meti, N.T.; Chalannavar, R.K. Melatonin: One Molecule One- Medicine for Many Diseases, Coronavirus (SARS-CoV-2) Disease (Covid19); Function in Plants. IJRSI 2021, 8, 155–181.
  35. Reiter, R.J.; Sharma, R.; Rodriguez, C.; Martin, V.; Rosales-Corral, S.; Zuccari, D.A.P.d.C.; Chuffa, L.G.d.A. Part-Time Cancers and Role of Melatonin in Determining Their Metabolic Phenotype. Life Sci. 2021, 278, 119597.
  36. Reiter, R.; Sharma, R.; Rosales-Corral, S.; Manucha, W.; Chuffa, L.G.; Zuccari, D.A.P.C. Melatonin and Pathological Cell Interactions: Mitochondrial Glucose Processing in Cancer Cells. Int. J. Mol. Sci. 2021, 22, 12494.
  37. Lauritzen, E.S.; Kampmann, U.; Pedersen, M.G.B.; Christensen, L.-L.; Jessen, N.; Møller, N.; Støy, J. Three Months of Melatonin Treatment Reduces Insulin Sensitivity in Patients with Type 2 Diabetes—A Randomized Placebo-Controlled Crossover Trial. J. Pineal Res. 2022, 73, e12809.
  38. Zhang, Y.; Yang, N.; Huang, X.; Zhu, Y.; Gao, S.; Liu, Z.; Cao, F.; Wang, Y. Melatonin Engineered Adipose-Derived Biomimetic Nanovesicles Regulate Mitochondrial Functions and Promote Myocardial Repair in Myocardial Infarction. Front. Cardiovasc. Med. 2022, 9, 789203.
  39. Bonilla, E.; Valero, N.; Chacin-Bonilla, L.; Medina-Leendertz, S. Melatonin and Viral Infections. J. Pineal Res. 2004, 36, 73–79.
  40. Tan, D.-X.; Korkmaz, A.; Reiter, R.J.; Manchester, L.C. Ebola Virus Disease: Potential Use of Melatonin as a Treatment. J. Pineal Res. 2014, 57, 381–384.
  41. Cardinali, D.P.; Brown, G.M.; Pandi-Perumal, S.R. Can Melatonin Be a Potential “Silver Bullet” in Treating COVID-19 Patients? Diseases 2020, 8, 44.
  42. Zhang, R.; Wang, X.; Ni, L.; Di, X.; Ma, B.; Niu, S.; Liu, C.; Reiter, R.J. COVID-19: Melatonin as a Potential Adjuvant Treatment. Life Sci. 2020, 250, 117583.
  43. Arnao, M.B.; Hernández-Ruiz, J. Melatonin: Synthesis from Tryptophan and Its Role in Higher Plant. In Amino Acids in Higher Plants; D’Mello, J.P.F., Ed.; CABI: Boston, MA, USA, 2015; pp. 390–435.
  44. Tan, D.-X.; Reiter, R.J. An Evolutionary View of Melatonin Synthesis and Metabolism Related to Its Biological Functions in Plants. J. Exp. Bot. 2020, 71, 4677–4689.
  45. Tan, D.X.; Manchester, C.L.; Esteban-Zubero, E.; Zhou, Z.; Reiter, J.R. Melatonin as a Potent and Inducible Endogenous Antioxidant: Synthesis and Metabolism. Molecules 2015, 20, 18886–18906.
  46. Back, K.; Tan, D.-X.; Reiter, R.J. Melatonin Biosynthesis in Plants: Multiple Pathways Catalyze Tryptophan to Melatonin in the Cytoplasm or Chloroplasts. J. Pineal Res. 2016, 61, 426–437.
  47. Byeon, Y.; Lee, H.Y.; Lee, K.; Back, K. Caffeic Acid O-Methyltransferase Is Involved in the Synthesis of Melatonin by Methylating N-Acetylserotonin in Arabidopsis. J. Pineal Res. 2014, 57, 219–227.
  48. Back, K. Melatonin Metabolism, Signaling and Possible Roles in Plants. Plant J. 2021, 105, 376–391.
  49. Tan, D.X.; Hardeland, R.; Back, K.; Manchester, L.C.; Latorre-Jimenez, M.A.; Reiter, R.J. On the Significance of an Alternate Pathway of Melatonin Synthesis via 5-Methoxytryptamine: Comparisons across Species. J. Pineal Res. 2016, 61, 27–40.
  50. Arnao, M.B.; Hernández-Ruiz, J. Functions of Melatonin in Plants: A Review. J. Pineal Res. 2015, 59, 133–150.
  51. Arnao, M.B.; Hernández-Ruiz, J. Is Phytomelatonin a New Plant Hormone? Agronomy 2020, 10, 95.
  52. Arnao, M.B.; Hernández-Ruiz, J. Melatonin Against Environmental Plant Stressors: A Review. Curr. Protein Pept. Sci. 2021, 22, 413–429.
  53. Ahmad, P.; Tripathi, D.K.; Deshmukh, R.; Pratap Singh, V.; Corpas, F.J. Revisiting the Role of ROS and RNS in Plants under Changing Environment. Environ. Exp. Bot. 2019, 161, 1–398.
  54. Melchiorri, D.; Reiter, R.J.; Attia, A.M.; Hara, M.; Burgos, A.; Nistico, G. Potent Protective Effect of Melatonin on in Vivo Paraquat-Induced Oxidative Damage in Rats. Life Sci. 1995, 56, 83–89.
  55. Reiter, R.J.; Tan, D.X.; Qi, W.; Manchester, L.C.; Karbownik, M.; Calvo, J.R. Pharmacology and Physiology of Melatonin in the Reduction of Oxidative Stress in Vivo. Biol. Signals Recept. 2000, 9, 160–171.
  56. Tan, D.X.; Manchester, L.C.; Terron, M.P.; Flores, L.J.; Reiter, R.J. One Molecule, Many Derivates: A Never-Ending Interaction of Melatonin with Reactive Oxygen and Nitrogen Species? J. Pineal Res. 2007, 42, 28–42.
  57. Prasad, A.; Sedlářová, M.; Balukova, A.; Rác, M.; Pospíšil, P. Reactive Oxygen Species as a Response to Wounding: In Vivo Imaging in Arabidopsis Thaliana. Front. Plant Sci. 2020, 10, 1660.
  58. Li, D.; Wei, J.; Peng, Z.; Ma, W.; Yang, Q.; Song, Z.; Sun, W.; Yang, W.; Yuan, L.; Xu, X.; et al. Daily Rhythms of Phytomelatonin Signaling Modulate Diurnal Stomatal Closure via Regulating Reactive Oxygen Species Dynamics in Arabidopsis. J. Pineal Res. 2020, 68, e12640.
  59. Arnao, M.B.; Hernández-Ruiz, J. Melatonin as a Regulatory Hub of Plant Hormone Levels and Action in Stress Situations. Plant Biol. 2021, 23, 7–19.
  60. Arnao, M.B.; Hernandez-Ruiz, J. Melatonin in Its Relationship to Plant Hormones. Ann. Bot. 2018, 121, 195–207.
  61. Arnao, M.B.; Cano, A.; Hernández-Ruiz, J. Phytomelatonin: An Unexpected Molecule with Amazing Performances in Plants. J. Exp. Bot. 2022, 73, 5779–5800.
  62. Arnao, M.B.; Hernández-Ruiz, J.; Cano, A.; Reiter, R.J. Melatonin and Carbohydrate Metabolism in Plant Cells. Plants 2021, 10, 1917.
  63. Arnao, M.B.; Hernández-Ruiz, J.; Cano, A. Role of Melatonin and Nitrogen Metabolism in Plants: Implications under Nitrogen-Excess or Nitrogen-Low. Int. J. Mol. Sci. 2022, 23, 15217.
  64. Arnao, M.B.; Hernández-Ruiz, J. Growth Activity, Rooting Capacity, and Tropism: Three Auxinic Precepts Fulfilled by Melatonin. Acta Physiol. Plant. 2017, 39, 127.
  65. Arnao, M.B.; Hernández-Ruiz, J. Protective Effect of Melatonin against Chlorophyll Degradation during the Senescence of Barley Leaves. J. Pineal Res. 2009, 46, 58–63.
  66. Arnao, M.B.; Hernández-Ruiz, J. Melatonin in Flowering, Fruit Set and Fruit Ripening. Plant Reprod. 2020, 33, 77–87.
  67. Giraldo-Acosta, M.; Cano, A.; Hernández-Ruiz, J.; Arnao, M.B. Melatonin as a Possible Natural Safener in Crops. Plants 2022, 11, 890.
  68. Caputo, G.A.; Wadl, P.A.; McCarty, L.; Adelberg, J.; Jennings, K.M.; Cutulle, M. In Vitro Safening of Bentazon by Melatonin in Sweetpotato (Ipomoea Batatas). Hortscience 2020, 55, 1406–1410.
  69. Abu-Qare, A.W.; Duncan, H.J. Herbicide Safeners: Uses, Limitations, Metabolism, and Mechanisms of Action. Chemosphere 2002, 48, 965–974.
  70. Giraldo-Acosta, M.; Martínez-Andújar, C.; Martínez-Melgarejo, P.A.; Cano, A.; Hernández-Ruiz, J.; Arnao, M.B. Protective Effect (Safener) of Melatonin on Vigna Radiata L. Seedlings in the Presence of the Fungicide Copper Oxychloride. J. Plant Growth Regul. 2022.
  71. Hernández-Ruiz, J.; Ruiz-Cano, D.; Giraldo-Acosta, M.; Cano, A.; Arnao, M.B. Melatonin in Brassicaceae: Role in Postharvest and Interesting Phytochemicals. Molecules 2022, 27, 1523.
  72. Wei, D.; Yang, J.; Xiang, Y.; Meng, L.; Pan, Y.; Zhang, Z. Attenuation of Postharvest Browning in Rambutan Fruit by Melatonin Is Associated With Inhibition of Phenolics Oxidation and Reinforcement of Antioxidative Process. Front. Nutr. 2022, 9, 905006.
  73. Zhu, L.; Hu, H.; Luo, S.; Wu, Z.; Li, P. Melatonin Delaying Senescence of Postharvest Broccoli by Regulating Respiratory Metabolism and Antioxidant Activity. Trans. Chin. Soc. Agric. Eng. 2018, 34, 300–308.
  74. Cano, A.; Giraldo-Acosta, M.; García-Sánchez, S.; Hernández-Ruiz, J.; Arnao, M.B. Effect of Melatonin in Broccoli Postharvest and Possible Melatonin Ingestion Level. Plants 2022, 11, 2000.
  75. Aghdam, M.S.; Mukherjee, S.; Flores, F.B.; Arnao, M.B.; Luo, Z.; Corpas, F.J. Functions of Melatonin During Postharvest of Horticultural Crops. Plant Cell Physiol. 2021, 63, 1764–1786.
  76. Zhao, C.; Nawaz, G.; Cao, Q.; Xu, T. Melatonin Is a Potential Target for Improving Horticultural Crop Resistance to Abiotic Stress. Sci. Hortic. 2022, 291, 110560.
  77. Moustafa-Farag, M.; Almoneafy, A.; Mahmoud, A.; Elkelish, A.; Arnao, M.B.; Li, L.; Ai, S. Melatonin and Its Protective Role against Biotic Stress Impacts on Plants. Biomolecules 2020, 10, 54.
  78. Collinge, D.B.; Jørgensen, H.J.L.; Latz, M.A.C.; Manzotti, A.; Ntana, F.; Rojas, E.C.; Jensen, B. Searching for Novel Fungal Biological Control Agents for Plant Disease Control Among Endophytes. In Endophytes for a Growing World; Murphy, B.R., Doohan, F.M., Saunders, M.J., Hodkinson, T.R., Eds.; Cambridge University Press: Cambridge, UK, 2019; pp. 25–51. ISBN 978-1-108-47176-3.
  79. Lamichhane, J.R.; Dachbrodt-Saaydeh, S.; Kudsk, P.; Messéan, A. Toward a Reduced Reliance on Conventional Pesticides in European Agriculture. Plant Dis. 2016, 100, 10–24.
  80. Geiger, F.; Bengtsson, J.; Berendse, F.; Weisser, W.W.; Emmerson, M.; Morales, M.B.; Ceryngier, P.; Liira, J.; Tscharntke, T.; Winqvist, C.; et al. Persistent Negative Effects of Pesticides on Biodiversity and Biological Control Potential on European Farmland. Basic Appl. Ecol. 2010, 11, 97–105.
  81. Kim, K.-H.; Kabir, E.; Jahan, S.A. Exposure to Pesticides and the Associated Human Health Effects. Sci. Total Environ. 2017, 575, 525–535.
  82. Raymaekers, K.; Ponet, L.; Holtappels, D.; Berckmans, B.; Cammue, B.P.A. Screening for Novel Biocontrol Agents Applicable in Plant Disease Management—A Review. Biol. Control 2020, 144, 104240.
  83. Bruce, T.J.A.; Smart, L.E.; Birch, A.N.E.; Blok, V.C.; MacKenzie, K.; Guerrieri, E.; Cascone, P.; Luna, E.; Ton, J. Prospects for Plant Defence Activators and Biocontrol in IPM—Concepts and Lessons Learnt so Far. Crop Prot. 2017, 97, 128–134.
  84. Fravel, D.R. Commercialization and Implementation of Biocontrol. Annu. Rev. Phytopathol. 2005, 43, 337–359.
  85. Vos, C.M.F.; De Cremer, K.; Cammue, B.P.A.; De Coninck, B. The Toolbox of Trichoderma Spp. in the Biocontrol of Botrytis Cinerea Disease. Mol. Plant Pathol. 2015, 16, 400–412.
  86. Burketova, L.; Trda, L.; Ott, P.G.; Valentova, O. Bio-Based Resistance Inducers for Sustainable Plant Protection against Pathogens. Biotechnol. Adv. 2015, 33, 994–1004.
  87. Tiwari, R.K.; Lal, M.K.; Kumar, R.; Mangal, V.; Altaf, M.A.; Sharma, S.; Singh, B.; Kumar, M. Insight into Melatonin-Mediated Response and Signaling in the Regulation of Plant Defense under Biotic Stress. Plant Mol. Biol. 2021, 109, 385–399.
  88. Balconi, C.; Stevanato, P.; Motto, M.; Biancardi, E. Breeding for Biotic Stress Resistance/Tolerance in Plants. In Crop Production for Agricultural Improvement; Ashraf, M., Öztürk, M., Ahmad, M.S.A., Aksoy, A., Eds.; Springer: Dordrecht, The Netherlands, 2012; pp. 57–114. ISBN 978-94-007-4116-4.
  89. Sofy, A.R.; Sofy, M.R.; Hmed, A.A.; El-Dougdoug, N.K. Potential Effect of Plant Growth-Promoting Rhizobacteria (PGPR) on Enhancing Protection Against Viral Diseases. In Field Crops: Sustainable Management by PGPR; Sustainable Development and Biodiversity; Maheshwari, D.K., Dheeman, S., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 411–445. ISBN 978-3-030-30926-8.
  90. Wang, S.M.; Hou, X.L.; Ying, L.; Cao, X.W.; Zhang, S.; Wang, F. Effects of Turnip mosaic virus (TuMV) on endogenous hormones and transcriptional level of related genes in infected non-heading Chinese cabbage. J. Nanjing Agric. Univ. 2011, 5, 13–19.
  91. Sofy, A.R.; Sofy, M.R.; Hmed, A.A.; Dawoud, R.A.; Refaey, E.E.; Mohamed, H.I.; El-Dougdoug, N.K. Molecular Characterization of the Alfalfa Mosaic Virus Infecting Solanum melongena in Egypt and the Control of Its Deleterious Effects with Melatonin and Salicylic Acid. Plants 2021, 10, 459.
  92. Iriti, M.; Faoro, F. Abscisic Acid Is Involved in Chitosan-Induced Resistance to Tobacco Necrosis Virus (TNV). Plant Physiol. Biochem. 2008, 46, 1106–1111.
  93. Zhao, L.; Feng, C.; Wu, K.; Chen, W.; Chen, Y.; Hao, X.; Wu, Y. Advances and Prospects in Biogenic Substances against Plant Virus: A Review. Pestic. Biochem. Physiol. 2017, 135, 15–26.
  94. Chen, L.; Wang, M.-R.; Li, J.-W.; Feng, C.-H.; Cui, Z.-H.; Zhao, L.; Wang, Q.-C. Exogenous Application of Melatonin Improves Eradication of Apple Stem Grooving Virus from the Infected in Vitro Shoots by Shoot Tip Culture. Plant Pathol. 2019, 68, 997–1006.
  95. Lu, R.; Liu, Z.; Shao, Y.; Sun, F.; Zhang, Y.; Cui, J.; Zhou, Y.; Shen, W.; Zhou, T. Melatonin Is Responsible for Rice Resistance to Rice Stripe Virus Infection through a Nitric Oxide-Dependent Pathway. Virol. J. 2019, 16, 141.
More
Information
Subjects: Agronomy
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , ,
View Times: 382
Revisions: 2 times (View History)
Update Date: 24 May 2023
1000/1000
Video Production Service