Next Article in Journal
Influence of the Culture Substrate on the Agaricus blazei Murrill Mushrooms Vitamins Content
Next Article in Special Issue
Genetically Modified Heat Shock Protein90s and Polyamine Oxidases in Arabidopsis Reveal Their Interaction under Heat Stress Affecting Polyamine Acetylation, Oxidation and Homeostasis of Reactive Oxygen Species
Previous Article in Journal
Effects of NaHCO3 Acclimation on Rye (Secale Cereale) Growth Under Sodic-Alkaline Stress
Previous Article in Special Issue
Effect of Thermospermine on the Growth and Expression of Polyamine-Related Genes in Rice Seedlings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Interplay among Polyamines and Nitrogen in Plant Stress Responses

by
Konstantinos Paschalidis
1,†,
Georgios Tsaniklidis
2,†,
Bao-Quan Wang
3,†,
Costas Delis
4,
Emmanouil Trantas
1,
Konstantinos Loulakakis
1,
Muhammad Makky
5,
Panagiotis F. Sarris
6,7,8,
Filippos Ververidis
1 and
Ji-Hong Liu
9,*
1
Department of Agriculture, School of Agricultural Sciences, Hellenic Mediterranean University, Estavromenos, GR-71500 Heraklion, Greece
2
National Agricultural Research Foundation (NAGREF), GR-71103 Heraklion, Greece
3
School of Horticulture and Landscape Architecture, Henan Province Engineering Research Center of Horticultural Plant Resource Utilization and Germplasm Enhancement, Henan Institute of Science and Technology, Xinxiang 453003, China
4
Department of Agriculture, University of the Peloponnese, GR-24100 Kalamata, Greece
5
Department of Agricultural Engineering, Universitas Andalas, Padang 25163, Indonesia
6
Biosciences, University of Exeter, Geoffrey Pope Building, Exeter EX4 4QD, UK
7
Institute of Molecular Biology & Biotechnology, GR-70013 Heraklion, Greece
8
Department of Biology, University of Crete, GR-70013 Heraklion, Greece
9
Key Laboratory of Horticultural Plant Biology, College of Horticulture and Forestry Sciences, Huazhong Agricultural University, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2019, 8(9), 315; https://doi.org/10.3390/plants8090315
Submission received: 26 July 2019 / Revised: 27 August 2019 / Accepted: 28 August 2019 / Published: 30 August 2019
(This article belongs to the Special Issue Plant Polyamines)

Abstract

:
The interplay between polyamines (PAs) and nitrogen (N) is emerging as a key factor in plant response to abiotic and biotic stresses. The PA/N interplay in plants connects N metabolism, carbon (C) fixation, and secondary metabolism pathways. Glutamate, a pivotal N-containing molecule, is responsible for the biosynthesis of proline (Pro), arginine (Arg) and ornithine (Orn) and constitutes a main common pathway for PAs and C/N assimilation/incorporation implicated in various stresses. PAs and their derivatives are important signaling molecules, as they act largely by protecting and preserving the function/structure of cells in response to stresses. Use of different research approaches, such as generation of transgenic plants with modified intracellular N and PA homeostasis, has helped to elucidate a plethora of PA roles, underpinning their function as a major player in plant stress responses. In this context, a range of transgenic plants over-or under-expressing N/PA metabolic genes has been developed in an effort to decipher their implication in stress signaling. The current review describes how N and PAs regulate plant growth and facilitate crop acclimatization to adverse environments in an attempt to further elucidate the N-PAs interplay against abiotic and biotic stresses, as well as the mechanisms controlling N-PA genes/enzymes and metabolites.

1. Introduction

Nitrogen metabolites, polyamines (PAs), and several important plant phytohormones, such as ethylene, jasmonates, abscisic acid, salicylic acid, have shown to act as crucial growth regulators that can cross talk with each other in stress signaling processes [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22]. PA biosynthesis/degradation and their homeostasis undergo extensive alterations in response to various stress conditions, such as cell wall degradation [23] oxidative and developmental stress [2,24,25,26,27,28,29,30,31], phytopathogenic bacteria/fungi/viruses [30], water deficiency [31,32,33], ammonia toxicity and nutrient availability [34], salinity [2,35,36,37] and heat [38,39].
Plants absorb nitrogen (N) mostly as nitrate or ammonia ions. The nitrate molecules are enzymically converted to ammonia, which is assimilated in plants for amino acid synthesis. Ammonia assimilation is mainly catalyzed by the glutamine synthetase (GS)/glutamate synthase (GOGAT) cycle [40]. Ammonia detoxification, on the other hand, is catalyzed mainly by glutamate dehydrogenase (GDH); however, under stress conditions GDH partly contributes to ammonia assimilation [34,41]. Furthermore, accumulating evidence shows that stress induces PA export and subsequent oxidations in the apoplast that play a role in production of H2O2 [22,30,42]. The apoplastic polyamine oxidase (PAO) in cooperation with the NADPH-oxidase creates a feedforward reactive oxygen species (ROS) magnification loop, affecting the oxidative status and climaxes in programmed cell death (PCD) performance. This loop may be a crucial point in many reactions governing salinity stress resistance, with possible functions spreading outside stress resistance [36]. In tobacco transgenic plants overexpressing ZmPAO, we detected greater apoplastic/cytoplasmic contents of H2O2 and superoxide, accompanied by increase in antioxidant genes; however, these antioxidants cannot efficiently scavenge ROS [27,28]. On the other hand, repression of ZmPAO in young tobacco seedlings enables them to resist short-term salinity, which can be attributed either to higher PA content or to lower ROS contents, because of the perturbed PA apoplastic oxidation [27,28].
The links between PAs and growth-regulatory pathways at molecular, biochemical and physiological levels, suggest that altering the expression of specific PA-response factors could provide a new strategy for targeted PA-response engineering. This review elucidates the concerted roles of N and PAs against plant abiotic and biotic stressors, as well as their interplay mechanisms, as far as the related genes/enzymes and metabolites are concerned, in order to help plants adapt to unfavorable environmental conditions.

2. Major Genes Involved in Abiotic and Biotic Stress Responses

It has been documented that drought and salinity, two main abiotic stress factors, disturb at least 20% of the arable land and nearly 40% of the irrigated land in the world [43,44]. These factors severely limit crop yields and result in the loss of more than US$100 billion per annum to the agricultural sector [45]. Drought represents a reduced soil water capacity, causing a decrease in root water uptake, while salinity leads to enhanced salt ion levels in the soil. Because of both lower water potential in the soil and higher ion levels, higher osmosis in plant cells/tissues might lead to elevated levels of osmolytes in order to achieve osmotic balancing. Furthermore, the higher concentration of ions inside the plant may induce an ionic chain reaction, which increases with the duration of the stress, leading to endocellular permeability of toxic ions. In order to combat with these harsh conditions, the plant can respond by either excluding ions or compartmentalizing them in vacuoles. In this regard, the osmotic phenomenon occurs very quickly and is found in all stress conditions, whereas the ionic phenomenon is quite long, progresses with time, and only under salinity [46]. In comparison to abiotic stress, several biotic stress factors, such as viruses, bacteria, fungi, nematodes, insects, and weeds, cause a direct deprivation of plant nutritional agents, leading to decreased strength in host plants. In agricultural terms, both biotic and abiotic stresses cause dramatical pre- and postharvest damages [47].
Plants adopt specific morphological and cellular alterations by sensing stress signals in order to adapt to environmental conditions. However, very few presumed sensors have been recognized. This is mainly due to the fact that functional genes encoding sensor proteins may exist in a redundant way, so alteration of one gene does not cause substantial phenotype alterations under stress. Alternatively, a sensor protein may be essential to plants and loss-of-function mutants are lethal to plants, prohibiting further analysis [43].
Numerous sets of genes/products are linked to abiotic/biotic stress responses at transcriptional and translational level. Genes leading to effective plant adaptation/tolerance may be categorized into four major groups: (i) Genes coding for enzymes of osmolyte biosynthesis, such as proline, mannitol, glycine, betaine, and trehalose; (ii) genes coding for antioxidant enzymes, such as superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT); (iii) genes coding for stress-induced proteins, such as antifreeze proteins, chaperons, and heat shock proteins; and (iv) genes coding for regulatory proteins, such as protein kinases and transcription factors [44]. So far, a large number of functional and regulatory genes implicated in abiotic and biotic stress responses have been identified in a diversity of plants. In particular, we characterized some functional or regulatory genes controlling metabolism of N, PAs, ethylene, and abscisic acid that are involved in various biological processes [22,33,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72].

3. Stress-Related Nitrogen Flow and Polyamines

Nitrogen is one of the main essential nutrient elements in plants. The N molecules inside plant cells are derived from soil inorganic N uptake, usually in the form of nitrate and ammonium ions, and from ammonia assimilation, N transport throughout the plant, and N remobilization (Figure 1) [34,73,74,75,76,77,78].
Nitrogen levels influence plant productivity and quality, due to association with various growth substances that are involved in plant stress responses [31,34,55,79,80]. PAs often increase in plants as a result of N application. They usually result from N-induced higher concentrations of their precursor amino acids, such as Orn and Arg, which are converted to putrescine (Put) [20,31,81]. The interplay among PAs and N is evolving as a major participant in plant stress reactions. The first synthesized amino acid is commonly glutamate (Glu), which participates in N recycling/remobilization into other nitrogenous molecules, ensuring N homeostasis in plants (Figure 1). The Glu, as a central N molecule, leads to biosynthesis of proline (Pro), Orn, Arg, and PAs, which constitute a crucial cooperating pathway for carbon (C) and nitrogen (N) assimilation [78]. Pro, Arg, and Put concentrations in plants are further known as some of the important indicators for both biotic and abiotic stress response [31]. S-adenosylmethionine (SAM), which is formed by methionine and involved in ethylene biosynthesis, and Orn, an amino acid involved in the urea cycle, are two important precursor molecules in PA synthesis (Figure 1). Put, Spd, Spm, and thermospermine, in turn, are important products of the organic N, as they are found at relatively high endogenous levels. PAs and their C scaffold are involved in several biochemical pathways. PA catabolism has a crucial role in N/C assimilation/remobilization, as it recycles C and N and produces H2O2 by PAO [22,26,39,42,73,78]. However, PA accumulation inside plant cells is the consequence of biosynthesis/catabolism, inter-conversions, and conjugation.
Carbon and N are crucial for developmental and stress responses in plants, in terms of life cycle accomplishment and crop production. Therefore, an appropriate N/C balancing is extremely critical for a variety of physiological and biological processes, including stress response. However, the N/C signaling machineries remain fundamentally undiscovered. The N/C cooperative genome-wide function has revealed that the majority of genes in Arabidopsis are over- or under-controlled by C and/or N input [78]. Furthermore, PA remobilization is related with the nitrate transport in parenchymal shoot tissues [82]. PA catabolism, producing H2O2 and GABA in the cell wall, is also tightly involved in preserving the N/C homeostasis and balance inside plant tissues [83].
It is widely accepted that the PA/N interplay in plants is of major interest, because it connects N metabolism, C fixation, and secondary metabolism pathways. Stress conditions, such as salinity and drought, increase the activity of proteases causing augmentation of ammonium ions inside cells (Figure 1). Ammonia is converted into glutamine and Glu by GS/GOGAT, respectively. Glu gives Orn that requires higher PA biosynthesis in response to various stressful conditions [41].
Orn is a key amino acid participating in cooperating pathways with major amino acids (Figure 1). Transgenic mouse ornithine decarboxylase (mODC) plants with depleted Orn exhibited higher Glu/Orn conversion into Put, leading to N shortage in the cell and to decreased protein synthesis. The same plants also transformed more Glu into Orn, which was partially compensated for enhanced Glu synthesis from integrated N and C [78]. Therefore, overall N assimilation/partitioning in plants is largely dependent on C availability/reallocation and vice versa. Under threatening abiotic and biotic stress conditions, the plants respond by remobilizing N and C into signaling molecules, such as PAs, Pro, GABA, glycine betaine, and β-Ala [78], as they have stress-protective key roles and partly alleviate ammonia cell toxicity. A Glu-Pro-Arg-PA-GABA coordinated path is therefore of major importance to accomplish an equilibrium among assimilated/partitioned N/C inside plant cells [34,78]. Proteomics and transcriptomics studies on PA-stress interaction and classification of major proteins implicated in important plant developmental/stress responses may provide new insights into the molecular mechanisms underlying these processes [84,85]. Moreover, by RNA-RNA in situ hybridization (ISH) methodologies, we have further elucidated the functions of N/PA genes in crop plants. Use of ISH has helped to identify the localization of PA anabolic and catabolic gene transcripts in tissues, such as the locular parenchyma and the vascular bundles, supporting the viewpoint that Put biosynthetic and catabolic genes are mostly expressed in fast growing tissues and that PAs are strongly implicated in fruit ripening [20].

4. N/PA Biotechnological Approaches for Enhanced Tolerance to Abiotic and Biotic Stress

Plants usually circumvent stress conditions by stimulating appropriate responses that lead to altered metabolism and growth. Tolerance to abiotic stress conditions might be achieved via genetic engineering through modifying the endogenous concentrations of osmoprotectants, by increasing ROS scavenging capacity or by robustly excluding ions with efficient transporter/symporter systems [86]. Taking into account that dissecting the function of stress-related genes would assist in elucidating the potential biochemical and molecular machineries for stress adaptation, enormous efforts and approaches have been expended to unravel the genes/proteins/metabolites associated with a plethora of cellular processes that regulate the complicated character of abiotic and biotic stress resistance [44].
As PAs have pleiotropic roles, their homeostasis control is complex. Genetic transformation of N assimilation/detoxification genes and PA biosynthetic genes coding for GDH, arginine decarboxylase (ADC), ODC, SAM decarboxylase (SAMDC), or Spd synthase (SPDS), significantly enhances abiotic stress resistance in numerous plant species [4,5,34,36,37,38,39,41,63,72,87,88,89,90,91,92,93,94,95,96].
The GS/GOGAT pathway is the main ammonia assimilation cycle in plants. However, under stress conditions, photorespiratory ammonia may hyperaccumulate due to decreased activity of GOGAT or GS. In this regard, we have discovered alternative metabolic ways like GDH that are activated in order to decrease ammonia buildup in the cells (Figure 1) [34]. Transgenic tobacco plants overexpressing the plant GDH gene encoding for the a-subunit polypeptide of GDH (gdh-NAD;A1) also exhibit higher ammonium assimilation activity [41]. Furthermore, transgenic rice overexpressing the GDH gene from Eurotium cheralieri (a lower organism that has stronger ammonium affinity compared to higher plants) showed higher N assimilation efficacy and yield, especially under low N conditions [87].
Under abiotic stress environment, PAs are apoplastically delivered and oxidized by PAO (Figure 1), generating several intermediates. We revealed two different outputs based on the level of the PA oxidized products. On the one hand, low apoplastic PAO generates less amount of H2O2, which in turn initiates a ROS protective pathway that triggers tolerance reactions. On the other hand, high apoplastic PAO could produce a large level of H2O2, thus triggering plant cell death (PCD) [22,28,42].
The above scenario illuminates mainly the intercellular PAs’ role. In another study, transgenic tobacco plants with down-regulated SAMDC underwent abiotic stress-induced PCD, and displayed lower endocellular levels of soluble Spd and Spm. However, we found that PA contents and apoplastic oxidation in the transgenic plants were unpredictably comparable to those of the wild type [28]. The down-regulated SAMDC transgenics, thus, present a balanced PA interplay among developmental and stress responses [32].
However, during biotic stress an opposite scenario is observed. In the PAO-overexpressing tobacco plants, we detected a stress-induced up-regulation of the PAO gene upon exposure to infection by Pseudomonas syringae pv tabaci [30]. The increased expression of the ADC gene may promote PA stability. Spm, in turn, is apoplastically excreted and broken down by the enhanced PAO, generating excess H2O2 that helps plants to cope with the pathogen attack [30]. Therefore, transgenic plants with increased PAO exhibited pre-induced resistance towards infections, including biotrophic and hemibiotrophic diseases [30]. Our stress defense model may represent a pioneering way for creating transgenic plants resistant to both abiotic and biotic stresses.
Figure 1. Proposed model for the nitrogen–polyamine (N-PA) interplay in plant abiotic/biotic stress signaling and defense. Abiotic or biotic stress induces proteases activity and increases in ammonium ions inside the cell. The reactive oxygen species (ROS) stress signal triggers the induction of the gene encoding the α-subunit of glutamate dehydrogenase (gdh-NAD; A1), which contributes, together with the GS/GOGAT cycle, to ammonia assimilation [34,41]. Stress-induced glutamate (Glu) production by GDH is diverted to Pro biosynthesis. Stressful conditions cause a further increase in endocellular PAs that are excreted and apoplastically oxidized by polyamine oxidase (PAOs), thus, producing H2O2 and numerous N composites. Depending on the level of H2O2 produced under abiotic stresses, programmed cell death (PCD; high H2O2 levels above a certain threshold) or H2O2 scavenging (low levels below a certain threshold) is activated [22,27,28,33]. However, the biotic stress-induced H2O2 causes a reverse pattern, as high H2O2 levels form an apoplastic “barrier” protecting the plant from fungi and bacteria [30]. Ascorbate peroxidase (APX) and other antioxidant genes are also involved during the protection response. Moreover, PAs are peroxisomally back-converted to generate H2O2 and N compounds that could activate Ca2+ permeable channels [2,22,28,32,73,97]. PAO (1): Decrease of PAO activity results in increased Spd and Spm contents and low levels of H2O2, leading to expression of defense genes and plant tolerance to abiotic stress, but susceptibility to biotic stress (fungi and bacteria); PAO (2): Increase of PAO activity results in lower Spd and Spm contents and high levels of H2O2, leading to abiotic stress-induced PCD accompanied by plant abiotic stress susceptibility, but tolerance to biotic stress due to high levels of H2O2, which form a “barrier” to fungi and bacteria. The other abbreviations are found in the text.
Figure 1. Proposed model for the nitrogen–polyamine (N-PA) interplay in plant abiotic/biotic stress signaling and defense. Abiotic or biotic stress induces proteases activity and increases in ammonium ions inside the cell. The reactive oxygen species (ROS) stress signal triggers the induction of the gene encoding the α-subunit of glutamate dehydrogenase (gdh-NAD; A1), which contributes, together with the GS/GOGAT cycle, to ammonia assimilation [34,41]. Stress-induced glutamate (Glu) production by GDH is diverted to Pro biosynthesis. Stressful conditions cause a further increase in endocellular PAs that are excreted and apoplastically oxidized by polyamine oxidase (PAOs), thus, producing H2O2 and numerous N composites. Depending on the level of H2O2 produced under abiotic stresses, programmed cell death (PCD; high H2O2 levels above a certain threshold) or H2O2 scavenging (low levels below a certain threshold) is activated [22,27,28,33]. However, the biotic stress-induced H2O2 causes a reverse pattern, as high H2O2 levels form an apoplastic “barrier” protecting the plant from fungi and bacteria [30]. Ascorbate peroxidase (APX) and other antioxidant genes are also involved during the protection response. Moreover, PAs are peroxisomally back-converted to generate H2O2 and N compounds that could activate Ca2+ permeable channels [2,22,28,32,73,97]. PAO (1): Decrease of PAO activity results in increased Spd and Spm contents and low levels of H2O2, leading to expression of defense genes and plant tolerance to abiotic stress, but susceptibility to biotic stress (fungi and bacteria); PAO (2): Increase of PAO activity results in lower Spd and Spm contents and high levels of H2O2, leading to abiotic stress-induced PCD accompanied by plant abiotic stress susceptibility, but tolerance to biotic stress due to high levels of H2O2, which form a “barrier” to fungi and bacteria. The other abbreviations are found in the text.
Plants 08 00315 g001
PA catabolism in plants plays a key role in the antioxidant machinery under stress conditions. Overexpression of an apoplastic PAO in tobacco plants led to higher expression of antioxidant machinery, including SOD and CAT [27]. However, the induced machinery did not conclude in stress defense, as it represented an effort to neutralize the PAO-produced H2O2. Thus, we suggested that continuous PA oxidation may lead to a continuous stress condition. The same genetically modified Nicotiana tabacum plants with altered PA/H2O2 levels due to over/underexpression of the ZmPAO gene were examined under heat stress. When the ZmPAO gene was repressed in transgenic plants, they exhibited better thermotolerance, higher biomass growth, and higher enzymatic and non-enzymatic antioxidant levels. In contrast, the ZmPAO-overexpressing plants showed a compromised thermotolerance [38]. Moreover, the ZmPAO-underexpressing plants possessed higher Ca2+ levels with salinity, associated with lower chlorophyll levels, leaf area and biomass, and a taller phenotype, than the wild type. The ZmPAO-overexpressing plants, on the contrary, had a higher number of leaves with slightly greater size and higher antioxidant genes/enzyme levels than the underexpressing ones [37]. Therefore, different phenotypes are found in PAO-overexpressing and underexpressing plants under abiotic/biotic stress conditions, revealing a multifaceted character of the apoplastic PAO. In this regard, PAO exerts an important role in rendering plants to survive under both abiotic and biotic stress conditions. It is further proposed that the PAO/NADPH oxidase loop is a focal point in the control of several defense processes in plants, including stress tolerance (Figure 1) [22,24,25,35,36].

5. Conclusion and Perspectives

Metabolic engineering has great potential to enhance abiotic and biotic stress tolerance. Many compounds play dynamic roles in incorporating stress signals, regulating stress response through modifying gene expressions and controlling numerous transporters and biochemical pathways in plants. Modification of a single step in the N/PA cycle (e.g., elevated Glu synthesis via transgenic GDH and/or increase of a specific PA via overexpressing of the respective PA biosynthetic gene) might cause a substantial redistribution of the metabolome in the cell. In this regard, for model plants (e.g., Arabidopsis) [35,98,99] or plants of industrial use (e.g., grapevine, tobacco, tomato, citrus, etc.) [22,27,28,29,30,33,34,37,38,39,41], metabolic engineering to alter N/PA metabolism and the accompanying N/C accumulation might provide a suitable means for elucidating the physiological mechanisms underlying the increase in crop yield and quality under stress conditions. Extensive knowledge of the N-PA crosstalk by means of engineering technologies may further open new avenues or suggest alternative possibilities for improving the quality of agricultural food products with additional paybacks, such as nutraceuticals and functional components.

Author Contributions

K.P., G.T., B.Q.W., F.V. and J.H.L., collected literatures and wrote the manuscript; C.D., E.T., K.L., M.M. and P.F.S. organized and edited the manuscript; K.P. and J.H.L., proof-read, and finalized the manuscript.

Funding

This work was financially supported by National Key Research and Development Program of China (2018YFD1000300), National Natural Science Foundation of China (31501727, 31320103908), and Hubei Provincial Natural Science Foundation for Innovative Group (2017CFA018).

Acknowledgments

The authors thank anonymous reviewers and the handling editor for providing valuable advice to improve the quality of this review.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ioannidis, N.E.; Malliarakis, D.; Torne, J.M.; Santos, M.; Kotzabasis, K. The over-expression of the plastidial transglutaminase from maize in arabidopsis increases the activation threshold of photoprotection. Front. Plant Sci. 2016, 7, 635. [Google Scholar] [CrossRef] [PubMed]
  2. Moschou, P.N.; Paschalidis, K.A.; Roubelakis-Angelakis, K.A. Plant polyamine catabolism: The state of the art. Plant Signal. Behav. 2008c, 3, 1061–1066. [Google Scholar] [CrossRef] [PubMed]
  3. Tavladoraki, P.; Cona, A.; Angelini, R. Copper-containing amine oxidases and fad-dependent polyamine oxidases are key players in plant tissue differentiation and organ development. Front. Plant Sci. 2016, 7, 824. [Google Scholar] [CrossRef] [PubMed]
  4. Minocha, R.; Majumdar, R.; Minocha, S.C. Polyamines and abiotic stress in plants: A complex relationship. Front. Plant Sci. 2014, 5, 175. [Google Scholar] [CrossRef] [PubMed]
  5. Tiburcio, A.F.; Alcazar, R. Potential applications of polyamines in agriculture and plant biotechnology. Methods Mol. Biol. 2018, 1694, 489–508. [Google Scholar] [PubMed]
  6. Handa, A.K.; Fatima, T.; Mattoo, A.K. Polyamines: Bio-molecules with diverse functions in plant and human health and disease. Front. Chem. 2018, 6, 10. [Google Scholar] [CrossRef]
  7. Fortes, A.M.; Agudelo-Romero, P. Polyamine metabolism in climacteric and non-climacteric fruit ripening. Methods Mol. Biol. 2018, 1694, 433–447. [Google Scholar]
  8. Wang, Y.; Ye, X.; Yang, K.; Shi, Z.; Wang, N.; Yang, L.; Chen, J. Characterization, expression, and functional analysis of polyamine oxidases for their role in selenium-induced hydrogen peroxide production in brassica rapa. J. Sci. Food Agric. 2019, 99, 4082–4093. [Google Scholar] [CrossRef]
  9. Podlesakova, K.; Ugena, L.; Spichal, L.; Dolezal, K.; De Diego, N. Phytohormones and polyamines regulate plant stress responses by altering gaba pathway. N. Biotechnol. 2019, 48, 53–65. [Google Scholar] [CrossRef]
  10. Igarashi, K.; Kashiwagi, K. The functional role of polyamines in eukaryotic cells. Int. J. Biochem. Cell Biol. 2019, 107, 104–115. [Google Scholar] [CrossRef]
  11. Bordenave, C.D.; Granados Mendoza, C.; Jimenez Bremont, J.F.; Garriz, A.; Rodriguez, A.A. Defining novel plant polyamine oxidase subfamilies through molecular modeling and sequence analysis. BMC Evol. Biol. 2019, 19, 28. [Google Scholar] [CrossRef] [PubMed]
  12. Gimenez, E.; Salinas, M.; Manzano-Agugliaro, F. Worldwide research on plant defense against biotic stresses as improvement for sustainable agriculture. Sustainability 2018, 10, 391. [Google Scholar] [CrossRef]
  13. Miao, H.; Sun, P.; Liu, Q.; Liu, J.; Xu, B.; Jin, Z. The agpase family proteins in banana: Genome-wide identification, phylogeny, and expression analyses reveal their involvement in the development, ripening, and abiotic/biotic stress responses. Int. J. Mol. Sci. 2017, 18, 1581. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, D.D.; Shao, Q.S.; Yin, L.H.; Younis, A.; Zheng, B.S. Polyamine function in plants: Metabolism, regulation on development, and roles in abiotic stress responses. Front. Plant Sci. 2019, 9, 1945. [Google Scholar] [CrossRef] [PubMed]
  15. Xu, J.J.; Liu, T.; Yang, S.C.; Jin, X.Q.; Qu, F.; Huang, N.; Hu, X.H. Polyamines are involved in gaba-regulated salinity-alkalinity stress tolerance in muskmelon. Environ. Exp. Bot. 2019, 164, 181–189. [Google Scholar] [CrossRef]
  16. Takahashi, T. Thermospermine: An evolutionarily ancient but functionally new compound in plants. Methods Mol. Biol. 2018, 1694, 51–59. [Google Scholar] [PubMed]
  17. Takahashi, T.; Kakehi, J. Polyamines: Ubiquitous polycations with unique roles in growth and stress responses. Ann. Bot. 2010, 105, 1–6. [Google Scholar] [CrossRef] [PubMed]
  18. Takano, A.; Kakehi, J.; Takahashi, T. Thermospermine is not a minor polyamine in the plant kingdom. Plant Cell Physiol. 2012, 53, 606–616. [Google Scholar] [CrossRef]
  19. Tsaniklidis, G.; Delis, C.; Nikoloudakis, N.; Katinakis, P.; Aivalakis, G. Low temperature storage affects the ascorbic acid metabolism of cherry tomato fruits. Plant Physiol. Biochem. 2014, 84, 149–157. [Google Scholar] [CrossRef]
  20. Tsaniklidis, G.; Kotsiras, A.; Tsafouros, A.; Roussos, P.A.; Aivalakis, G.; Katinakis, P.; Delis, C. Spatial and temporal distribution of genes involved in polyamine metabolism during tomato fruit development. Plant Physiol. Biochem. 2016, 100, 27–36. [Google Scholar] [CrossRef]
  21. Paschalidis, K.A.; Aziz, A.; Geny, L.; Primikirios, N.I.; Roubelakis-Angelakis, K.A. Polyamines in grapevine. In Molecular Biology & Biotechnology of the Grapevine; Roubelakis-Angelakis, K.A., Ed.; Springer: Dordrecht, The Netherlands, 2001; pp. 109–151. [Google Scholar]
  22. Wang, W.; Paschalidis, K.; Feng, J.C.; Song, J.; Liu, J.H. Polyamine catabolism in plants: A universal process with diverse functions. Front. Plant Sci. 2019, 10, 561. [Google Scholar] [CrossRef] [PubMed]
  23. Papadakis, A.K.; Paschalidis, K.A.; Roubelakis-Angelakis, K.A. Biosynthesis profile and endogenous titers of polyamines differ in totipotent and recalcitrant plant protoplasts. Physiol. Plant 2005, 125, 10–20. [Google Scholar] [CrossRef]
  24. Papadakis, A.K.; Roubelakis-Angelakis, K.A. Polyamines inhibit nadph oxidase-mediated superoxide generation and putrescine prevents programmed cell death induced by polyamine oxidase-generated hydrogen peroxide. Planta 2005, 220, 826–837. [Google Scholar] [CrossRef] [PubMed]
  25. Paschalidis, K.A.; Roubelakis-Angelakis, K.A. Sites and regulation of polyamine catabolism in the tobacco plant. Correlations with cell division/expansion, cell cycle progression, and vascular development. Plant Physiol. 2005b, 138, 2174–2184. [Google Scholar] [CrossRef] [PubMed]
  26. Paschalidis, K.A.; Roubelakis-Angelakis, K.A. Spatial and temporal distribution of polyamine levels and polyamine anabolism in different organs/tissues of the tobacco plant. Correlations with age, cell division/expansion, and differentiation. Plant Physiol. 2005a, 138, 142–152. [Google Scholar] [CrossRef] [PubMed]
  27. Moschou, P.N.; Delis, I.D.; Paschalidis, K.A.; Roubelakis-Angelakis, K.A. Transgenic tobacco plants overexpressing polyamine oxidase are not able to cope with oxidative burst generated by abiotic factors. Physiol. Plant 2008a, 133, 140–156. [Google Scholar] [CrossRef] [PubMed]
  28. Moschou, P.N.; Paschalidis, K.A.; Delis, I.D.; Andriopoulou, A.H.; Lagiotis, G.D.; Yakoumakis, D.I.; Roubelakis-Angelakis, K.A. Spermidine exodus and oxidation in the apoplast induced by abiotic stress is responsible for H2O2 signatures that direct tolerance responses in tobacco. Plant Cell 2008b, 20, 1708–1724. [Google Scholar] [CrossRef] [PubMed]
  29. Moschou, P.N.; Sanmartin, M.; Andriopoulou, A.H.; Rojo, E.; Sanchez-Serrano, J.J.; Roubelakis-Angelakis, K.A. Bridging the gap between plant and mammalian polyamine catabolism: A novel peroxisomal polyamine oxidase responsible for a full back-conversion pathway in arabidopsis. Plant Physiol. 2009b, 147, 1845–1857. [Google Scholar] [CrossRef]
  30. Moschou, P.N.; Sarris, P.F.; Skandalis, N.; Andriopoulou, A.H.; Paschalidis, K.A.; Panopoulos, N.J.; Roubelakis-Angelakis, K.A. Engineered polyamine catabolism preinduces tolerance of tobacco to bacteria and oomycetes. Plant Physiol. 2009a, 149, 1970–1981. [Google Scholar] [CrossRef]
  31. Paschalidis, K.; Moschou, P.N.; Aziz, A.; Toumi, I.; Roubelakis-Angelakis, Κ.A. Polyamines in grapevine: An update. In Grapevine Molecular Physiology & Biotechnology; Roubelakis-Angelakis, K.A., Ed.; Springer: Dordrecht, The Netherlands, 2009a; pp. 207–228. [Google Scholar]
  32. Toumi, I.; Moschou, P.N.; Paschalidis, K.A.; Bouamama, B.; Ben Salem-Fnayou, A.; Ghorbel, A.W.; Mliki, A.; Roubelakis-Angelakis, K.A. Abscisic acid signals reorientation of polyamine metabolism to orchestrate stress responses via the polyamine exodus pathway in grapevine. J. Plant Physiol. 2010, 167, 519–525. [Google Scholar] [CrossRef]
  33. Paschalidis, K.A.; Toumi, I.; Moschou, P.N.; Roubelakis-Angelakis, K.A. Aba-dependent amine oxidases-derived h2o2 affects stomata conductance. Plant Signal. Behav. 2010, 5, 1153–1156. [Google Scholar]
  34. Skopelitis, D.S.; Paranychianakis, N.V.; Paschalidis, K.A.; Pliakonis, E.D.; Delis, I.D.; Yakoumakis, D.I.; Kouvarakis, A.; Stephanou, E.; Papadakis, A.K.; Roubelakis-Angelakis, K.A. Abiotic stress generates ros that signal expression of anionic glutamate dehydrogenases to form glutamate for proline synthesis in tobacco and grapevine. Plant Cell 2006, 18, 2767–2781. [Google Scholar] [CrossRef] [PubMed]
  35. Andronis, E.A.; Moschou, P.N.; Roubelakis-Angelakis, K.A. Peroxisomal polyamine oxidase and nadph-oxidase cross-talk for ros homeostasis which affects respiration rate in arabidopsis thaliana. Front. Plant Sci. 2014, 5, 132. [Google Scholar] [CrossRef] [PubMed]
  36. Gemes, K.; Kim, Y.J.; Park, K.Y.; Moschou, P.N.; Andronis, E.; Valassaki, C.; Roussis, A.; Roubelakis-Angelakis, K.A. An nadph-oxidase/polyamine oxidase feedback loop controls oxidative burst under salinity. Plant Physiol. 2016, 172, 1418–1431. [Google Scholar] [CrossRef] [PubMed]
  37. Gemes, K.; Mellidou, I.; Karamanoli, K.; Beris, D.; Park, K.Y.; Matsi, T.; Haralampidis, K.; Constantinidou, H.I.; Roubelakis-Angelakis, K.A. Deregulation of apoplastic polyamine oxidase affects development and salt response of tobacco plants. J. Plant Physiol. 2017, 211, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Mellidou, I.; Karamanoli, K.; Beris, D.; Haralampidis, K.; Constantinidou, H.A.; Roubelakis-Angelakis, K.A. Underexpression of apoplastic polyamine oxidase improves thermotolerance in Nicotiana tabacum. J. Plant Physiol. 2017, 218, 171–174. [Google Scholar] [CrossRef] [PubMed]
  39. Mellidou, I.; Moschou, P.N.; Pankou, C.; Valassakis, C.; Ioannidis, N.; Gémes, K.; Andronis, E.A.; Roussis, A.; Beris, D.; Haralampidis, K.; et al. Nicotiana tabacum plants with silenced s-adenosyl -l-methionine decarboxylase (samdc) reveal a pa-dependent trade-off between growth and tolerance responses. Front. Plant Sci. 2016, 7, 1–17. [Google Scholar] [CrossRef] [PubMed]
  40. Lea, P.J.; Miflin, B.J. Alternative route for nitrogen assimilation in higher plants. Nat. Rev. Genet. 1974, 251, 614–616. [Google Scholar] [CrossRef]
  41. Skopelitis, D.S.; Paranychianakis, N.V.; Kouvarakis, A.; Spyros, A.; Stephanou, E.G.; Roubelakis-Angelakis, K.A. The isoenzyme 7 of tobacco nad(h)-dependent glutamate dehydrogenase exhibits high deaminating and low aminating activities in vivo. Plant Physiol. 2007, 145, 1726–1734. [Google Scholar] [CrossRef]
  42. Moschou, P.N.; Roubelakis-Angelakis, K.A. Polyamines and programmed cell death. J. Exp. Bot. 2014, 65, 1285–1296. [Google Scholar] [CrossRef]
  43. Zhu, J.-K. Abiotic stress signaling and responses in plants. Cell 2016, 167, 313–324. [Google Scholar] [CrossRef] [PubMed]
  44. Singhal, P.; Jan, A.T.; Azam, M.; Haq, Q.M.R. Plant abiotic stress: A prospective strategy of exploiting promoters as alternative to overcome the escalating burden. Front. Life Sci. 2016, 9, 52–63. [Google Scholar] [CrossRef]
  45. Shabala, S.; Bose, J.; Fuglsang, A.T.; Pottosin, I. On a quest for stress tolerance genes: Membrane transporters in sensing and adapting to hostile soils. J. Exp. Bot. 2016, 67, 1015–1031. [Google Scholar] [CrossRef] [PubMed]
  46. Kosová, K.; Vítámvás, P.; Urban, M.O.; Prášil, I.T.; Renaut, J. Plant abiotic stress proteomics: The major factors determining alterations in cellular proteome. Front. Plant Sci. 2018, 9, 1–22. [Google Scholar] [CrossRef]
  47. Romero, F.M.; Maiale, S.J.; Rossi, F.R.; Marina, M.; Ruiz, O.A.; Garriz, A. Polyamine metabolism responses to biotic and abiotic stress. Methods Mol. Biol. 2018, 1694, 37–49. [Google Scholar] [PubMed]
  48. Gago, C.; Drosou, V.; Paschalidis, K.; Guerreiro, A.; Miguel, G.; Antunes, D.; Hilioti, Z. Targeted gene disruption coupled with metabolic screen approach to uncover the leafy cotyledon1-like4 (l1l4) function in tomato fruit metabolism. Plant Cell Rep. 2017, 36, 1065–1082. [Google Scholar] [CrossRef] [PubMed]
  49. Manganaris, G.A.; Drogoudi, P.; Goulas, V.; Tanou, G.; Georgiadou, E.C.; Pantelidis, G.E.; Paschalidis, K.A.; Fotopoulos, V.; Manganaris, A. Deciphering the interplay among genotype, maturity stage and low-temperature storage on phytochemical composition and transcript levels of enzymatic antioxidants in prunus persica fruit. Plant Physiol. Biochem. 2017, 119, 189–199. [Google Scholar] [CrossRef] [PubMed]
  50. Dhima, Κ.; Vasilakoglou, Ι.; Stefanou, S.; Gatsis, T.; Paschalidis, K.; Aggelopoulos, S.; Eleftherohorinos, I. Differential competitive and allelopathic ability of cyperus rotundus on solanum lycopersicum, solanum melongena and capsicum annuum. Arch. Agron. Soil Sci. 2016, 62, 1250–1263. [Google Scholar] [CrossRef]
  51. Dhima, K.; Vasilakoglou, I.; Paschalidis, K.A.; Gatsis, T.; Keco, R. Productivity and phytotoxicity of six sunflower hybrids and their residues effects on rotated lentil and ivy-leaved speedwell. Field Crop. Res. 2012, 136, 42–51. [Google Scholar] [CrossRef]
  52. Goumenaki, E.; Karidis, Z.; Paschalidis, K.A. Assessment of tropospheric ozone impact on crops in crete (greece) using snap bean as bioindicator. Acta Hortic. 2012, 938, 401–407. [Google Scholar] [CrossRef]
  53. Makky, M.; Paschalidis, K.A.; Dhima, K.; Manganaris, A. A new rapid gas chromatographic method for ethylene, respirational, and senescent gaseous production of climacteric fruits stored in prolonged low temperature. Proc. Int. Conf. Agric. Environ. Biol. Sci. (AEBS-2014) 2014a, 1, 24–25. [Google Scholar]
  54. Makky, M.; Paschalidis, K.A.; Dhima, K.; Mangganaris, A. Tomato fruits (solanaceae lycopersicon esculentum mill.) feedback mechanism in the presence of exogenous ethylene under prolonged chilling temperature storage. J. Nutr. Pharm. Res. 2015, 1, 4–12. [Google Scholar]
  55. Ninou, E.G.; Paschalidis, K.A.; Mylonas, I.G.; Vasilikiotis, C.; Mavromatis, A.G. The effect of genetic variation and nitrogen fertilization on productive characters of greek oregano. Acta Agric. Scand. Sect. B—Soil Plant Sci. 2017b, 67, 372–379. [Google Scholar] [CrossRef]
  56. Vasilakoglou, I.; Dhima, K.; Paschalidis, K.; Gatsis, T.; Zacharis, K.; Galanis, M. Field bindweed (convolvulus arvensis l.) and redroot pigweed (amaranthus retroflexus l.) control in potato by pre-or post-emergence applied flumioxazin and sulfosulfuron. Chil. J. Agric. Res. 2013, 73, 24–30. [Google Scholar] [CrossRef]
  57. Mantzorou, A.; Navakoudis, E.; Paschalidis, K.; Ververidis, F. Microalgae: A potential tool for remediating aquatic environments from toxic metals. Int. J. Environ. Sci. Technol. 2018, 15, 1815–1830. [Google Scholar] [CrossRef]
  58. Paschalidis, K.A.; Moschou, P.N.; Toumi, I.; Roubelakis-Angelakis, K.A. Polyamine anabolic/catabolic regulation along the woody grapevine plant axis. J. Plant Physiol. 2009b, 166, 1508–1519. [Google Scholar] [CrossRef]
  59. Mougiou, Ν.; Trikka, F.; Trantas, E.; Ververidis, F.; Makris, A.; Argiriou, A.; Vlachonasios, K.E. Expression of hydroxytyrosol and oleuropein biosynthetic genes are correlated with metabolite accumulation during fruit development in olive, olea europaea, cv. Koroneiki. Plant Physiol. Biochem. 2018, 128, 41–49. [Google Scholar] [CrossRef]
  60. Papadakis, I.E.; Tsiantas, P.I.; Tsaniklidis, G.; Landi, M.; Psychoyou, M.; Fasseas, C. Changes in sugar metabolism associated to stem bark thickening partially assist young tissues of eriobotrya japonica seedlings under boron stress. J Plant Physiol. 2018, 231, 337–345. [Google Scholar] [CrossRef]
  61. Trantas, A.E.; Sarris, P.F.; Mpalantinaki, E.; Papadimitriou, M.; Ververidis, F.; Goumas, D.E. First report of xanthomonas hortorum pv. Hederae causing bacterial leaf spot on ivy in greece. Plant Disease 2016, 100, 1–10. [Google Scholar] [CrossRef]
  62. Trantas, E.A.; Mpalantinaki, E.; Pagoulatou, M.; Sarris, P.F.; Ververidis, F.; Goumas, D.E. First report of bacterial apical necrosis of mango caused by pseudomonas syringae pv. Syringae in greece. Plant Disease 2017, 101, 1541. [Google Scholar] [CrossRef]
  63. Wang, W.; Wu, H.; Liu, J.H. Genome-wide identification and expression profiling of copper-containing amine oxidase genes in sweet orange (citrus sinensis). Tree Genet. Genomes 2017, 13, 31. [Google Scholar] [CrossRef]
  64. Wu, H.; Fu, B.; Sun, P.; Xiao, C.; Liu, J.H. A nac transcription factor represses putrescine biosynthesis and affects drought tolerance. Plant Physiol. 2016, 172, 1532–1547. [Google Scholar] [CrossRef] [PubMed]
  65. Gong, X.; Zhang, J.; Hu, J.; Wang, W.; Wu, H.; Zhang, Q.; Liu, J.H. Fcwrky70, a wrky protein of fortunella crassifolia, functions in drought tolerance and modulates putrescine synthesis by regulating arginine decarboxylase gene. Plant Cell Environ. 2015, 38, 2248–2262. [Google Scholar] [CrossRef] [PubMed]
  66. Huang, X.S.; Zhang, Q.; Zhu, D.; Fu, X.; Wang, M.; Zhang, Q.; Moriguchi, T.; Liu, J.H. Ice1 of poncirus trifoliata functions in cold tolerance by modulating polyamine levels through interacting with arginine decarboxylase. J. Exp. Bot. 2015, 66, 3259–3274. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, J.; Liu, J.H.; Yu, A.; Xiang, Y.; Kurosawa, T.; Nada, K.; Ban, Y. Cloning, biochemical identification, and expression analysis of a gene encoding s-adenosylmethionine decarboxylase in citrus sinensis osbeck. J. Hortic. Sci. Biotechnol. 2010, 85, 219–226. [Google Scholar] [CrossRef]
  68. Fu, X.Z.; Chen, C.W.; Wang, Y.; Liu, J.H.; Moriguchi, T. Ectopic expression of mdspds1 in sweet orange (citrus sinensis osbeck) reduces canker susceptibility: Involvement of H2O2 production and transcriptional alteration. BMC Plant Biol. 2011, 11, 55. [Google Scholar] [CrossRef] [PubMed]
  69. Shi, J.; Fu, X.Z.; Peng, T.; Huang, X.S.; Fan, Q.J.; Liu, J.H. Spermine pretreatment confers dehydration tolerance of citrus in vitro plants via modulation of antioxidative capacity and stomatal response. Tree Physiol. 2010, 30, 914–922. [Google Scholar] [CrossRef] [PubMed]
  70. Zhang, Q.; Wang, M.; Hu, J.; Wang, W.; Fu, X.; Liu, J.H. Ptrabf of poncirus trifoliata functions in dehydration tolerance by reducing stomatal density and maintaining reactive oxygen species homeostasis. J. Exp. Bot. 2015, 66, 5911–5927. [Google Scholar] [CrossRef]
  71. Zhang, X.; Wang, W.; Wang, M.; Zhang, H.Y.; Liu, J.H. The mir396b of poncirus trifoliata functions in cold tolerance by regulating acc oxidase gene expression and modulating ethylene-polyamine homeostasis. Plant Cell Physiol. 2016, 57, 1865–1878. [Google Scholar] [CrossRef]
  72. Liu, J.-H.; Wang, W.; Wu, H.; Gong, X.; Moriguchi, T. Polyamines function in stress tolerance: From synthesis to regulation. Front. Plant Sci. 2015, 6, 1–10. [Google Scholar] [CrossRef]
  73. Moschou, P.N.; Wu, J.; Cona, A.; Tavladoraki, P.; Angelini, R.; Roubelakis-Angelakis, K.A. The polyamines and their catabolic products are significant players in the nitrogenous turnover in plants. J. Exp. Bot. 2012, 63, 5003–5015. [Google Scholar] [CrossRef] [PubMed]
  74. Loulakakis, K.A.; Primikirios, N.I.; Nikolantonakis, M.A.; Roubelakis-Angelakis, K.A. Immunocharacterization of vitis vinifera l. Ferredoxin-dependent glutamate synthase, and its spatial and temporal changes during leaf development. Planta 2002, 215, 630–638. [Google Scholar] [CrossRef] [PubMed]
  75. Syntichaki, K.M.; Loulakakis, K.A.; Roubelakis-Angelakis, K.A. The amino-acid sequence similarity of plant glutamate dehydrogenase to the extremophilic archaeal enzyme conforms to its stress-related function. Gene 1996, 168, 87–92. [Google Scholar] [CrossRef]
  76. Loulakakis, K.A.; Roubelakis-Angelakis, K.A.; Kanellis, A.K. Regulation of glutamate dehydrogenase and glutamine synthetase in avocado fruit during development and ripening. Plant Physiol. 1994, 106, 217–222. [Google Scholar] [CrossRef] [PubMed]
  77. Loulakakis, K.A.; Roubelakis-Angelakis, K.A. Plant nad(h)-glutamate dehydrogenase consists of two subunit polypeptides and their participation in the seven isoenzymes occurs in an ordered ratio. Plant Physiol. 1991, 97, 104–111. [Google Scholar] [CrossRef]
  78. Majumdar, R.; Barchi, B.; Turlapati, S.A.; Gagne, M.; Minocha, R.; Long, S.; Minocha, S.C. Glutamate, ornithine, arginine, proline, and polyamine metabolic interactions: The pathway is regulated at the post-transcriptional level. Front. Plant Sci. 2016, 7, 78. [Google Scholar] [CrossRef]
  79. Ninou, E.; Paschalidis, K.; Mylonas, I. Essential oil responses to water stress in greek oregano populations. J. Essent. Oil Bear. Plants 2017a, 20, 12–23. [Google Scholar] [CrossRef]
  80. Makky, M.; Paschalidis, K.A.; Dhima, K.; Manganaris, A. Harnessing untapped bio-ethylene sources from tomatoes climacteric effluent. Proc. Int. Conf. Agric. Environ. Biol. Sci. (AEBS-2014) 2014b, 1, 24–25. [Google Scholar]
  81. Serapiglia, M.J.; Minocha, R.; Minocha, S.C. Changes in polyamines, inorganic ions and glutamine synthetase activity in response to nitrogen availability and form in red spruce (picea rubens). Tree Physiol. 2008, 28, 1793–1803. [Google Scholar] [CrossRef]
  82. Tong, W.; Imai, A.; Tabata, R.; Shigenobu, S.; Yamaguchi, K.; Yamada, M.; Hasebe, M.; Sawa, S.; Motose, H.; Takahashi, T. Polyamine resistance is increased by mutations in a nitrate transporter gene nrt1.3 (atnpf6.4) in arabidopsis thaliana. Front. Plant Sci. 2016, 7, 1–10. [Google Scholar] [CrossRef]
  83. Wu, Q.Y.; Ma, S.Z.; Zhang, W.W.; Yao, K.B.; Chen, L.; Zhao, F.; Zhuang, Y.Q. Accumulating pathways of gamma-aminobutyric acid during anaerobic and aerobic sequential incubations in fresh tea leaves. Food Chem. 2018, 240, 1081–1086. [Google Scholar] [CrossRef] [PubMed]
  84. Liu, Y.; Liang, H.; Lv, X.; Liu, D.; Wen, X.; Liao, Y. Effect of polyamines on the grain filling of wheat under drought stress. Plant Physiol. Biochem. 2016, 100, 113–129. [Google Scholar] [CrossRef] [PubMed]
  85. Tanou, G.; Ziogas, V.; Belghazi, M.; Christou, A.; Filippou, P.; Job, D.; Fotopoulos, V.; Molassiotis, A. Polyamines reprogram oxidative and nitrosative status and the proteome of citrus plants exposed to salinity stress. Plant Cell Environ. 2014, 37, 864–885. [Google Scholar] [CrossRef] [PubMed]
  86. Nguyen, H.C.; Lin, K.H.; Ho, S.L.; Chiang, C.M.; Yang, C.M. Enhancing the abiotic stress tolerance of plants: From chemical treatment to biotechnological approaches. Physiol. Plant 2018, 164, 452–466. [Google Scholar] [CrossRef] [PubMed]
  87. Tang, D.; Peng, Y.; Lin, J.; Du, C.; Yang, Y.; Wang, D.; Liu, C.; Yan, L.; Zhao, X.; Li, X.; et al. Ectopic expression of fungal ecgdh improves nitrogen assimilation and grain yield in rice. J. Integr. Plant Biol. 2018, 60, 85–88. [Google Scholar] [CrossRef] [PubMed]
  88. Liu, J.H.; Ban, Y.; Wen, X.P.; Nakajima, I.; Moriguchi, T. Molecular cloning and expression analysis of an arginine decarboxylase gene from peach (prunus persica). Gene 2009, 429, 10–17. [Google Scholar] [CrossRef] [PubMed]
  89. Liu, J.H.; Moriguchi, T. Changes in free polyamine titers and expression of polyamine biosynthetic genes during growth of peach in vitro callus. Plant Cell Rep. 2007, 26, 125–131. [Google Scholar] [CrossRef] [PubMed]
  90. Liu, J.H.; Nada, K.; Honda, C.; Kitashiba, H.; Wen, X.P.; Pang, X.M.; Moriguchi, T. Polyamine biosynthesis of apple callus under salt stress: Importance of the arginine decarboxylase pathway in stress response. J. Exp. Bot. 2006, 57, 2589–2599. [Google Scholar] [CrossRef]
  91. Liu, J.H.; Nakajima, I.; Moriguchi, T. Effects of salt and osmotic stresses on free polyamine content and expression of polyamine biosynthetic genes in vitis vinifera. Biol. Plant. 2011, 55, 340–344. [Google Scholar] [CrossRef]
  92. Sun, P.; Zhu, X.; Huang, X.; Liu, J.H. Overexpression of a stress-responsive myb transcription factor of poncirus trifoliata confers enhanced dehydration tolerance and increases polyamine biosynthesis. Plant Physiol. Biochem. 2014, 78, 71–79. [Google Scholar] [CrossRef]
  93. Wang, B.Q.; Zhang, Q.F.; Liu, J.H.; Li, G.H. Overexpression of ptadc confers enhanced dehydration and drought tolerance in transgenic tobacco and tomato: Effect on ros elimination. Biochem. Biophys. Res. Commun. 2011b, 413, 10–16. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, J.; Sun, P.P.; Chen, C.L.; Wang, Y.; Fu, X.Z.; Liu, J.H. An arginine decarboxylase gene ptadc from poncirus trifoliata confers abiotic stress tolerance and promotes primary root growth in arabidopsis. J. Exp. Bot. 2011a, 62, 2899–2914. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, W.; Liu, J.H. Genome-wide identification and expression analysis of the polyamine oxidase gene family in sweet orange (citrus sinensis). Gene 2015, 555, 421–429. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, W.; Liu, J.H. Cspao4 of citrus sinensis functions in polyamine terminal catabolism and inhibits plant growth under salt stress. Sci. Rep. 2016, 6, 31384. [Google Scholar] [CrossRef] [PubMed]
  97. Wu, J.; Qu, H.; Shang, Z.; Jiang, X.; Moschou, P.N.; Roubelakis-Angelakis, K.A.; Zhang, S. Spermidine oxidase-derived H2O2 activates downstream ca2+ channels which signal pollen tube growth in pyruspyrifolia. Plant J. 2010, 63, 1042–1053. [Google Scholar] [CrossRef] [PubMed]
  98. Zarza, X.; Atanasov, K.E.; Marco, F.; Arbona, V.; Carrasco, P.; Kopka, J.; Fotopoulos, V.; Munnik, T.; Gomez-Cadenas, A.; Tiburcio, A.F.; et al. Polyamine oxidase 5 loss-of-function mutations in arabidopsis thaliana trigger metabolic and transcriptional reprogramming and promote salt stress tolerance. Plant Cell Environ. 2017, 40, 527–542. [Google Scholar] [CrossRef]
  99. Fincato, P.; Moschou, P.N.; Ahou, A.; Angelini, R.; Roubelakis-Angelakis, K.A.; Federico, R.; Tavladoraki, P. The members of arabidopsis thaliana pao gene family exhibit distinct tissue- and organ-specific expression pattern during seedling growth and flower development. Amino Acids 2012, 42, 831–841. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Paschalidis, K.; Tsaniklidis, G.; Wang, B.-Q.; Delis, C.; Trantas, E.; Loulakakis, K.; Makky, M.; Sarris, P.F.; Ververidis, F.; Liu, J.-H. The Interplay among Polyamines and Nitrogen in Plant Stress Responses. Plants 2019, 8, 315. https://doi.org/10.3390/plants8090315

AMA Style

Paschalidis K, Tsaniklidis G, Wang B-Q, Delis C, Trantas E, Loulakakis K, Makky M, Sarris PF, Ververidis F, Liu J-H. The Interplay among Polyamines and Nitrogen in Plant Stress Responses. Plants. 2019; 8(9):315. https://doi.org/10.3390/plants8090315

Chicago/Turabian Style

Paschalidis, Konstantinos, Georgios Tsaniklidis, Bao-Quan Wang, Costas Delis, Emmanouil Trantas, Konstantinos Loulakakis, Muhammad Makky, Panagiotis F. Sarris, Filippos Ververidis, and Ji-Hong Liu. 2019. "The Interplay among Polyamines and Nitrogen in Plant Stress Responses" Plants 8, no. 9: 315. https://doi.org/10.3390/plants8090315

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