The Response of Primary Metabolites to Abiotic Stresses: Comparison
Please note this is a comparison between Version 2 by Jessie Wu and Version 1 by Sajid Ali.

Metabolomics is a promising way to interpret biotic and abiotic stress tolerance in plants. The study of metabolic profiling revealed different types of metabolites, e.g., amino acids, carbohydrates, phenols, polyamines, terpenes, etc, which are accumulated in plants. Among all, primary metabolites, such as amino acids, carbohydrates, lipids polyamines, and glycine betaine, are considered the major contributing factors that work as osmolytes and osmoprotectants for plants from various environmental stress factors. In contrast, plant-derived secondary metabolites, e.g., phenolics, terpenoids, and nitrogen-containing compounds (alkaloids), have no direct role in the growth and development of plants. Nevertheless, such metabolites could play a significant role as a defense by protecting plants from biotic factors such as herbivores, insects, and pathogens. In addition, they can enhance the resistance against abiotic factors. 

 

  • metabolomics
  • tolerance
  • abiotic stress

1. Introduction

Plants established several adaptive mechanisms to endure abiotic factors, containing variations of metabolism in various directions, to confirm their existence in combative environmental situations [66][1] (Table 1). Several plant metabolites could assist and reduce the effect of the harsh stress of salt, drought, and water by acting as osmolytes and osmoprotectants [67][2]. Examples of such metabolites include dimethylsulfoniopropionate (DMSP) and glycine betaine; sugars, such as sucrose, trehalose, and fructan; amino acids, such as proline and ectoine, as well as some metabolites of polyols, sorbitol, and mannitol [68,69][3][4]. In plants, a wide range of waxy layers known as epicuticular wax keeps water balance during water shortage and acts as a mechanical stoppage to encounter disease-causing agents. Additionally, ascorbic acids, glutamine, alpha-tocopherol, anthocyanins, and carotene shield plant tissues by foraging the intermediates of bustling oxygen produced during oxidative stress [70][5]. Similarly, several other smaller compounds guard plants against oxidation damage related to various constrictions [65][6].
Besides, the plant’s defense system is related to generating phytoalexins, stimulating the common phenylpropanoid pathway and producing lignin biosynthesis [71][7]. Further, phytochemicals and hormones such as salicylic acid and methyl salicylate, methyl jasmonate and jasmonic acid, as well as other small molecules formed due to stress, play a significant role against environmental stresses [72,73,74][8][9][10]. All of these may also function as signaling compounds by stimulating the resistance system and reactions of acclimation [75][11]. Among the defense systems of plants, osmotic regulation is one of the broadly pronounced responses to the water shortage that needs the accretion of harmonious solutes, such as sugars, amino acids, polyols, and glycine betaine [76][12]. These chemical compounds do a significant job in sustaining cell turgor and stabilizing cell membranes and protein. Moreover, other studies designate the importance of these compounds in rehabilitating redox stability through the scavenging of ROS, which could adversely affect cellular structures and metabolism [68,77][3][13].

2. Amino Acids

Amino acids are considered a precursor for protein and other organic molecules, e.g., nucleic acids, which designate an active part in the responses of a plant under several stress factors. Amino acids could also play a significant role in signaling and controlling molecules [92][41]. Various studies showed that many amino acids stored in plants are apparent to different abiotic stresses [93,94][42][43]. Moreover, the exposure of plants to such stresses appearance an accumulation of proline and other amino acids. In plants, the role played by stored amino acids differs after acting as an osmolyte to adjust ions passage, reducing stomatal opening and reclamation of TM [95][17]. Moreover, amino acids can also disturb the synthesis and activity of several enzymes, gene expression, and redox state of homeostasis [96][14]. The accumulation of proline and ectoine is considered the most extensively dispersed osmolytes, as they act as osmoprotectants to protect plants from harmful effects and exciting environmental stresses, including low and high temperature, salinity, UVr, water, and osmotic stresses [68,97][3][15].
Primarily, proline is produced from a glutamate and proline metabolizing enzyme, pyrroline-5-carboxylate synthetase (P5CS), which reduces glutamate to pyrroline-5-carboxylate (P5C). At last, from the reduction of P5C, this stress-responsive amino acid forms by pyrroline-5-carboxylate reductases (P5CR) [98][44]. In transgenic plants, the significant role of proline was established during osmotic stress. For example, overexpression of the P5CS gene in soybean increased proline content and, thus, tolerance to salt stress in transgenic plants [99][45]. Besides osmolytes, proline is thought to accomplish many other important functions related to plant resistance, e.g., ROS scavenging, redox balancing, cytosolic pH buffer, molecular chaperon, and a stabilizer of protein structure [98][44]. Subsequently, in response to abiotic factors, the enlarged levels of proline were observed for several years to be the stress-responsive feature in plants. The relationship between the accumulation of proline as osmolytes and stress tolerance had a great share because of its applicability to different crops [100,101][46][47].
Remarkably, some of the metabolites were related to drought resistance and drought vulnerability of the considered hybrids [102][48]. Additionally, studies on drought responses at metabolomics levels indicated that Andean potatoes with a phenotype designating greater stress exposure have more proline related to the genetically assembled plant that was a higher dearth-tolerant [103][49]. It was established that the cultivar with a sensitive phenotype has high-level certain amino acids, containing proline and Gamma-aminobutyric acid (GABA) when barley exposed to salinity stress [104][50]. It may well advocate a greater liability of these plants to such stress. According to [96][14], this accretion could be associated with the deterioration of the leaf and slowing the development of a more subtle genotype. Furthermore, studies on Arabidopsis revealed that proline could be a lethal compound under heat stress [105][51], while Charlton et al. found that water deficiency was the cause of the decrease in isoleucine concentration in Pea and Arabidopsis plants [106][20].

3. Polyamines

Plants are tested by different stress factors and adversely affect their growth, yield, and geographical circulation [107][52]. To survive the combative environmental stress circumstances, plants have developed many adaptive strategies, amongst which the accumulation of metabolites plays an important defensive role [108][53]. Metabolites strongly involved in stress resistance are the low-molecular-weight (LMW) acyclic polyamines [109][54]. Polyamines are the LMW nitrogen-containing organic compounds with more than two amino groups with a positive charge at the cellular pH, allowing them to link with negatively charged molecules, such as nucleic acids, phospholipids, and proteins [110][55]. Usually, polyamines are polycations essential for plant growth and development and play an important role in abiotic stress resistance in higher plants. Triamine spermidine, tetraamine spermine, and their diamine predecessor, putrescine, are the general polyamines [111][56]. Because of their cationic nature, these compounds have often been correlated to environmental stresses, such as drought, chilling, heat, TM, and salinity [112][57].
The results of Khan et al. [95][17] and Capell et al. [113][58] showed that the accumulation of spermidine with the up-regulation of spermidine synthase of Cucurbita ficifolia augmented several stress responses in a recombinant Arabidopsis plant, such as waterlogging and salinity stresses. It was shown that spermidine acts as a signaling molecule and controls the assertion of intricate genes in drought resistance. Furthermore, it has been demonstrated that polyamines are attributed to being involved in maintaining membranes shielding from damage under stressful environments [114][59] and controlling the formation of nucleic acid as well as enzyme activity [115][60]. Additionally, different findings revealed that polyamines play a significant role in oxidative stress by mitigating the balance state of ROS through their direct contact or indirectly regulating the antioxidant system and suppressing ROS production. Moreover, some authors hypothesized that polyamines could act as a cellular signal in plants throughout the stress responses [116][61].

4. Carbohydrates

Carbohydrates produced during photosynthesis are the main building units that provide energy and support to the plant biomass [117][62]. Extensive studies revealed that non-living factors lead to the assemblage of non-structural saccharides, such as sucrose and lactose, simple sugars, or polyhydric compounds (alcohols and phenols), amongst various species of plants [118][63]. Particularly, there is a robust association between carbohydrate accretion and osmotic stress resistance, including oxidative stress (ROS) conditions, salt stress, and the scarcity of water [95][17]. As a source of carbon and energy in a cell, soluble carbohydrates may take a significant part in the metabolic processes of plants. Several stress factors may impact the level of these soluble carbohydrates because the accumulation of carbohydrates is associated with photosynthesis [119][64]. Rosa et al. [120][65] demonstrated that certain soluble sugars, such as sucrose and hexoses, improved stress tolerance by down-regulating the stress-related genes and up-regulating growth-related genes. Though, the contents of certain carbohydrates, such as raffinose, glucose, fructose, and maltose, are highly sensitive to environmental stresses and increase. However, the contents of myoinositol were reduced in barley roots during water-scarce conditions [121][66]. The findings of Sperdouli and Moustakas [122][67] revealed an increase and contact of augmented soluble carbohydrates, sustaining a great antioxidant defense in the leaves of Arabidopsis thaliana under dry environmental stress conditions. Studies showed renovation of carbon metabolism under salt-related stress (paraquat) in A. thaliana tissues and inferred by the researchers as a substitute approach to staying alive [122][67].
In water-deficit conditions, soluble sugars function as osmoprotectants, decreasing the harmful impact of osmotic stress and helps in sustaining the turgidity of cell and cell membrane stability by keeping plants from humiliation [123][68]. Under stress conditions, the increase in sugar quantity is generally the result of carbohydrate hydrolysis that needs enzymes with hydrolytic usage [124][18]. Moreover, carbohydrates that are soluble, such as disaccharides (sucrose and trehalose), oligosaccharides (raffinose and stachyose), and polymer of fructose molecules (fructans) next to their linked metabolic enzymes are essential compatible osmolytes associated with the scavenging of unstable molecules (ROS) during their assortment in plant tissues [125][69]. In low-temperature stress, sugar alcohols, such as polyols, function as osmoprotectants and shield cell membranes against ice adhesion [77][13]. Moreover, carbohydrates may act as signaling molecules [126][70]. The demonstrated data advocate a specific response of carbohydrates in plants. However, it should be noted that the accumulation of carbohydrates depends on the kind of stress to which it bared [127][71].

5. Glycine Betaine

Glycine betaine (GB) is a widely studied quat compound, which is active in retaining the water balance between the plant cell and the environment during drought conditions. Moreover, GB playing a significant role in stabilizing the macromolecules, shielding photosynthesis, detoxification of reactive oxygen radicals, and as an osmoprotectant [128,129][21][72]. Several studies indicated their importance in improving plant tolerance under various abiotic factors. It has been shown that plants are distinguished according to the formation of GB, such as barley, spinach, maize, and wheat, produce and accumulate a higher quantity of GB in their chloroplast. However, some plant species cannot obtain substantial amounts of GB during stress, such as A. thaliana, rice, and tobacco [130][73]. Furthermore, it has been shown that transgenic plants could mitigate the impact of abiotic stresses. Therefore, efforts have been made to improve tolerance through glycine betaine biosynthesis to achieve transgenic plants. In transgenic plants, such as Arabidopsis, the cyanobacteria genes, such as glycine sarcosine methyltransferase, and in transgenic maize, a greater amount of GB accumulates. As a result, in transgenic Arabidopsis, resistance to drought and salt is greater; nevertheless, a recombinant plant of maize retained well in cold-related to non-transgenic cultivars [131,132][74][75].
Moreover, through genetic engineering, other transgenic plants with a GB-producing capacity have been achieved, including Brassica juncea and tobacco with greater tolerance to salt and chilling, indicating a progressive ability to propagate and grow well related to wild-type in abiotic environmental conditions [133,134][76][77]. Besides, transgenic tomatoes with GB synthesis were more resistant to cold stress and produced fruit at a rate from 10 to 30% higher than the wild type. [135][78]. Though, the meditations of GB produced in every transgenic plant were scarce to control the osmotic stress to which plants were exposed. Similarly, previous studies showed that GB could enhance root growth and reduce oxidative stress. Additionally, the exogenous application of GB improves the stress tolerance of Cr in chickpea plants [136][22] and salinity stress in wheat [137][79]. Consequently, further protecting approaches of GB, such as defense against ROS and heavy metals stress, should be considered, which may enhance the tolerance level [138][23].

6. Lipids

Lipids are a fundamental component of biological membranes, particularly the plasma membrane, which serves as the contact between the cell and its surroundings [72][8]. Lipids can be grouped into eight major types based on the chemical structure in conjunction with distinctive hydrophobic and hydrophilic components, such as fatty acids, glycerides, phosphoglycerides, sphingolipids, steroids, isoprenoids, glycolipids, and polyketides [139][80]. Being sessile organisms, plants are subjected to a wide variety of biotic and abiotic factors, such as temperature, drought, heavy metals, salinity, and pathogen attack. However, lipid-mediated signaling occurs in response to all these stressors (Figure 1). The plasma membrane, which is typically the signaling source of lipids, is commonly used by plants to sense these stimuli and transform the signal into subsequent biochemical metabolism. Generally, these are acclimating enzymes that have all been proposed as signaling lipids, such as phospholipases, lipid kinases, and phosphatases [140][81]. Commonly, lysophospholipid, fatty acid, phosphatides, triacylglycerol, inositol phosphate, oxylipins, sphingolipids, and nacylethanolamine are considered the major contributing signaling lipids molecules [141][82]. The conformation and activity of cellular proteins and metabolites are influenced by signaling lipids because they have the ability to temporarily attract molecular markers to the membrane.
Figure 1. Environmental stresses of biotic and abiotic factors affecting the growth and homeostasis of plants.
  and abiotic factors affecting the growth and homeostasis of plants.
The enzyme phospholipase A (PLA) is very important in the formation of fatty acids and lysophospholipids. Usually, lysophospholipids are present in very limited amounts in plant tissues; however, in stressed conditions such as freezing their quantity increases [142][83]. Some reports revealed the physiological role of lysophospholipids against various environmental stresses. Similarly, the phospholipase A2 (PlA2) has been shown to increase the production of some elicitors in poppy plants [143][84], while lysophosphatidyl-choline and lysophosphatidyl-ethanolamine act as signals transducers in arbuscular symbiosis in potato [144][85].
Fatty acids have also been demonstrated as stress-responsive lipids in plants. Oleic acids modulate nitric oxide-related proteins, thereby regulating nitric oxide and mitigating tolerance in Arabidopsis [145][86]. Moreover, fatty acids also regulate drought, salt, and heavy metals tolerance, as well as the wound-induced responses of pathogens/herbivores in plants [146][87]. Likewise, the responsive role of phosphatidic acid (PA), inositol polyphosphates, oxylipins, sphingolipids, and some other lipids have been studied in various plant species [147,148,149][88][89][90]. Some of the environmental stress factors under which the plant lipid responses were reported to include chilling, freezing, and wounding [150][91], pathogens [151][24], low-temperature stress [152][92], salt stress [153][93], and water and drought [154][26] stress response.

 

References

  1. Ashapkin, V.V.; Kutueva, L.I.; Aleksandrushkina, N.I.; Vanyushin, B.F. Epigenetic mechanisms of plant adaptation to biotic and abiotic stresses. Int. J. Mol. Sci. 2020, 21, 7457.
  2. Teklić, T.; Parađiković, N.; Špoljarević, M.; Zeljković, S.; Lončarić, Z.; Lisjak, M. Linking abiotic stress, plant metabolites, biostimulants and functional food. Ann. Appl. Biol. 2021, 178, 169–191.
  3. Dikilitas, M.; Simsek, E.; Roychoudhury, A. Role of proline and glycine betaine in overcoming abiotic stresses. In Protective Chemical Agents in the Amelioration of Plant Abiotic Stress: Biochemical and Molecular Perspectives; Wiley: Hoboken, NJ, USA, 2020; pp. 1–23.
  4. Kebert, M.; Vuksanović, V. Species-Level Differences in Osmoprotectants and Antioxidants Contribute to Stress Tolerance of Quercus robur L., and Q. cerris L. Seedlings under Water Deficit and High Temperatures. Plants 2022, 11, 1744.
  5. Hüdig, M.; Schmitz, J.; Engqvist, M.; Maurino, V. Biochemical control systems for small molecule damage in plants. Plant Signal. Behav. 2018, 13, e1477906.
  6. Patel, J.; Khandwal, D.; Choudhary, B.; Ardeshana, D.; Jha, R.K.; Tanna, B.; Yadav, S.; Mishra, A.; Varshney, R.K.; Siddique, K.H. Differential Physio-Biochemical and Metabolic Responses of Peanut (Arachis hypogaea L.) under Multiple Abiotic Stress Conditions. Int. J. Mol. Sci. 2022, 23, 660.
  7. Kishi-Kaboshi, M.; Okada, K.; Kurimoto, L.; Murakami, S.; Umezawa, T.; Shibuya, N.; Yamane, H.; Miyao, A.; Takatsuji, H.; Takahashi, A. A rice fungal MAMP-responsive MAPK cascade regulates metabolic flow to antimicrobial metabolite synthesis. Plant J. 2010, 63, 599–612.
  8. Hou, Q.; Ufer, G.; Bartels, D. Lipid signalling in plant responses to abiotic stress. Plant Cell Environ. 2016, 39, 1029–1048.
  9. Li, X.; Zhang, L.-P.; Zhang, L.; Yan, P.; Ahammed, G.J.; Han, W.-Y. Methyl salicylate enhances flavonoid biosynthesis in tea leaves by stimulating the phenylpropanoid pathway. Molecules 2019, 24, 362.
  10. Creelman, R.A.; Tierney, M.L.; Mullet, J.E. Jasmonic acid/methyl jasmonate accumulate in wounded soybean hypocotyls and modulate wound gene expression. Proc. Natl. Acad. Sci. USA 1992, 89, 4938–4941.
  11. Sharma, A.; Shahzad, B.; Rehman, A.; Bhardwaj, R.; Landi, M.; Zheng, B. Response of phenylpropanoid pathway and the role of polyphenols in plants under abiotic stress. Molecules 2019, 24, 2452.
  12. Rady, M.O.; Semida, W.M.; Abd El-Mageed, T.A.; Hemida, K.A.; Rady, M.M. Up-regulation of antioxidative defense systems by glycine betaine foliar application in onion plants confer tolerance to salinity stress. Sci. Hortic. 2018, 240, 614–622.
  13. Janská, A.; Maršík, P.; Zelenková, S.; Ovesná, J. Cold stress and acclimation–what is important for metabolic adjustment? Plant Biol. 2010, 12, 395–405.
  14. Rodziewicz, P.; Swarcewicz, B.; Chmielewska, K.; Wojakowska, A.; Stobiecki, M. Influence of abiotic stresses on plant proteome and metabolome changes. Acta Physiol. Plant. 2014, 36, 1–19.
  15. Verbruggen, N.; Hermans, C. Proline accumulation in plants: A review. Amino Acids 2008, 35, 753–759.
  16. Grace, S.C. Phenolics as antioxidants. Antioxid. React. Oxyg. Species Plants 2005, 141, 168.
  17. Khan, N.; Ali, S.; Zandi, P.; Mehmood, A.; Ullah, S.; Ikram, M.; Ismail, M.A.S.; Babar, M. Role of sugars, amino acids and organic acids in improving plant abiotic stress tolerance. Pak. J. Bot 2020, 52, 355–363.
  18. Kaplan, F.; Guy, C.L. β-Amylase induction and the protective role of maltose during temperature shock. Plant Physiol. 2004, 135, 1674–1684.
  19. Pan, H.; Kou, P.; Yang, J.; Niu, L.; Wan, N.; Zhao, C.; Liu, Z.; Gu, C.; Fu, Y. A novel approach for efficient extraction and enrichment of phytochemicals with CO2-based switchable-solvent from pigeon pea leaves. J. Clean. Prod. 2021, 284, 124629.
  20. Charlton, A.J.; Donarski, J.A.; Harrison, M.; Jones, S.A.; Godward, J.; Oehlschlager, S.; Arques, J.L.; Ambrose, M.; Chinoy, C.; Mullineaux, P.M. Responses of the pea (Pisum sativum L.) leaf metabolome to drought stress assessed by nuclear magnetic resonance spectroscopy. Metabolomics 2008, 4, 312–327.
  21. Raza, S.H.; Athar, H.R.; Ashraf, M.; Hameed, A. Glycinebetaine-induced modulation of antioxidant enzymes activities and ion accumulation in two wheat cultivars differing in salt tolerance. Environ. Exp. Bot. 2007, 60, 368–376.
  22. Singh, D.; Singh, C.K.; Singh, D.; Sarkar, S.K.; Prasad, S.K.; Sharma, N.L.; Singh, I. Glycine betaine modulates chromium (VI)-induced morpho-physiological and biochemical responses to mitigate chromium toxicity in chickpea (Cicer arietinum L.) cultivars. Sci. Rep. 2022, 12, 8005.
  23. Emamverdian, A.; Ding, Y.; Mokhberdoran, F.; Xie, Y. Heavy metal stress and some mechanisms of plant defense response. Sci. World J. 2015, 2015, 756120 .
  24. Pinosa, F.; Buhot, N.; Kwaaitaal, M.; Fahlberg, P.; Thordal-Christensen, H.; Ellerström, M.; Andersson, M.X. Arabidopsis phospholipase Dδ is involved in basal defense and nonhost resistance to powdery mildew fungi. Plant Physiol. 2013, 163, 896–906.
  25. Ochoa-Villarreal, M.; Howat, S.; Hong, S.; Jang, M.O.; Jin, Y.-W.; Lee, E.-K.; Loake, G.J. Plant cell culture strategies for the production of natural products. BMB Rep. 2016, 49, 149.
  26. Uraji, M.; Katagiri, T.; Okuma, E.; Ye, W.; Hossain, M.A.; Masuda, C.; Miura, A.; Nakamura, Y.; Mori, I.C.; Shinozaki, K. Cooperative function of PLDδ and PLDα1 in abscisic acid-induced stomatal closure in Arabidopsis. Plant Physiol. 2012, 159, 450–460.
  27. Parida, A.K.; Dagaonkar, V.S.; Phalak, M.S.; Umalkar, G.; Aurangabadkar, L.P. Alterations in photosynthetic pigments, protein and osmotic components in cotton genotypes subjected to short-term drought stress followed by recovery. Plant Biotechnol. Rep. 2007, 1, 37–48.
  28. Pagare, S.; Bhatia, M.; Tripathi, N.; Pagare, S.; Bansal, Y. Secondary metabolites of plants and their role: Overview. Curr. Trends Biotechnol. Pharm. 2015, 9, 293–304.
  29. Martinez, V.; Mestre, T.C.; Rubio, F.; Girones-Vilaplana, A.; Moreno, D.A.; Mittler, R.; Rivero, R.M. Accumulation of flavonols over hydroxycinnamic acids favors oxidative damage protection under abiotic stress. Front. Plant Sci. 2016, 7, 838.
  30. Kabera, J.N.; Semana, E.; Mussa, A.R.; He, X. Plant secondary metabolites: Biosynthesis, classification, function and pharmacological properties. J. Pharm. Pharm. 2014, 2, 377–392.
  31. Saltveit, M.E. Synthesis and metabolism of phenolic compounds. Fruit Veget. Phytochem. Chem. Hum. Health 2017, 2, 115.
  32. Gu, C.-Z.; Xia, X.-M.; Lv, J.; Tan, J.-W.; Baerson, S.R.; Pan, Z.-q.; Song, Y.-Y.; Zeng, R.-S. Diterpenoids with herbicidal and antifungal activities from hulls of rice (Oryza sativa). Fitoterapia 2019, 136, 104183.
  33. Yang, C.-Q.; Wu, X.-M.; Ruan, J.-X.; Hu, W.-L.; Mao, Y.-B.; Chen, X.-Y.; Wang, L.-J. Isolation and characterization of terpene synthases in cotton (Gossypium hirsutum). Phytochemistry 2013, 96, 46–56.
  34. Vaughan, M.M.; Christensen, S.; Schmelz, E.A.; Huffaker, A.; Mcauslane, H.J.; Alborn, H.T.; Romero, M.; Allen, L.H.; Teal, P.E. Accumulation of terpenoid phytoalexins in maize roots is associated with drought tolerance. Plant Cell Environ. 2015, 38, 2195–2207.
  35. Massacci, A.; Nabiev, S.; Pietrosanti, L.; Nematov, S.; Chernikova, T.; Thor, K.; Leipner, J. Response of the photosynthetic apparatus of cotton (Gossypium hirsutum) to the onset of drought stress under field conditions studied by gas-exchange analysis and chlorophyll fluorescence imaging. Plant Physiol. Biochem. 2008, 46, 189–195.
  36. Wu, W.; Zhang, Q.; Zhu, Y.; Lam, H.-M.; Cai, Z.; Guo, D. Comparative metabolic profiling reveals secondary metabolites correlated with soybean salt tolerance. J. Agric. Food Chem. 2008, 56, 11132–11138.
  37. Divekar, P.A.; Narayana, S.; Divekar, B.A.; Kumar, R.; Gadratagi, B.G.; Ray, A.; Singh, A.K.; Rani, V.; Singh, V.; Singh, A.K. Plant secondary metabolites as defense tools against herbivores for sustainable crop protection. Int. J. Mol. Sci. 2022, 23, 2690.
  38. Davodnia, B.; Ahmahdi, J.; Fabriki Ourang, S. Evaluation of drought and salinity stresses on morphological and biochemical characteristics in four species of Papaver. Eco-Phytochem. J. Med. Plants 2017, 5, 24–36.
  39. Mewis, I.; Khan, M.A.; Glawischnig, E.; Schreiner, M.; Ulrichs, C. Water stress and aphid feeding differentially influence metabolite composition in Arabidopsis thaliana (L.). PLoS ONE 2012, 7, e48661.
  40. Taha, R.S.; Seleiman, M.F.; Alhammad, B.A.; Alkahtani, J.; Alwahibi, M.S.; Mahdi, A.H. Activated Yeast extract enhances growth, anatomical structure, and productivity of Lupinus termis L. plants under actual salinity conditions. Agronomy 2020, 11, 74.
  41. Heinemann, B.; Hildebrandt, T.M. The role of amino acid metabolism in signaling and metabolic adaptation to stress-induced energy deficiency in plants. J. Exp. Bot. 2021, 72, 4634–4645.
  42. Khan, A.; Ali, S.; Murad, W.; Hayat, K.; Siraj, S.; Jawad, M.; Khan, R.A.; Uddin, J.; Al-Harrasi, A.; Khan, A. Phytochemical and pharmacological uses of medicinal plants to treat cancer: A case study from Khyber Pakhtunkhwa, North Pakistan. J. Ethnopharmacol. 2021, 281, 114437.
  43. Hildebrandt, T.M. Synthesis versus degradation: Directions of amino acid metabolism during Arabidopsis abiotic stress response. Plant Mol. Biol. 2018, 98, 121–135.
  44. Meena, M.; Divyanshu, K.; Kumar, S.; Swapnil, P.; Zehra, A.; Shukla, V.; Yadav, M.; Upadhyay, R. Regulation of L-proline biosynthesis, signal transduction, transport, accumulation and its vital role in plants during variable environmental conditions. Heliyon 2019, 5, e02952.
  45. Yang, M.M.; Wang, J.; Dong, L.; Kong, D.J.; Teng, Y.; Liu, P.; Fan, J.J.; Yu, X.H. Lack of association of C3 gene with uveitis: Additional insights into the genetic profile of uveitis regarding complement pathway genes. Sci. Rep. 2017, 7, 1–8.
  46. Hanif, S.; Saleem, M.F.; Sarwar, M.; Irshad, M.; Shakoor, A.; Wahid, M.A.; Khan, H.Z. Biochemically triggered heat and drought stress tolerance in rice by proline application. J. Plant Growth Regul. 2021, 40, 305–312.
  47. Rashedy, A.A.; Abd-ElNafea, M.H.; Khedr, E.H. Co-application of proline or calcium and humic acid enhances productivity of salt stressed pomegranate by improving nutritional status and osmoregulation mechanisms. Sci. Rep. 2022, 12, 1–10.
  48. Abeed, A.H.; Eissa, M.A.; Abdel-Wahab, D.A. Effect of exogenously applied jasmonic acid and kinetin on drought tolerance of wheat cultivars based on morpho-physiological evaluation. J. Soil Sci. Plant Nutr. 2021, 21, 131–144.
  49. Vasquez-Robinet, C.; Mane, S.P.; Ulanov, A.V.; Watkinson, J.I.; Stromberg, V.K.; De Koeyer, D.; Schafleitner, R.; Willmot, D.B.; Bonierbale, M.; Bohnert, H.J. Physiological and molecular adaptations to drought in Andean potato genotypes. J. Exp. Bot. 2008, 59, 2109–2123.
  50. Li, Q.; Zhao, Y.; Ding, W.; Han, B.; Geng, S.; Ning, D.; Ma, T.; Yu, X. Gamma-aminobutyric acid facilitates the simultaneous production of biomass, astaxanthin and lipids in Haematococcus pluvialis under salinity and high-light stress conditions. Bioresour. Technol. 2021, 320, 124418.
  51. Lv, W.-T.; Lin, B.; Zhang, M.; Hua, X.-J. Proline accumulation is inhibitory to Arabidopsis seedlings during heat stress. Plant Physiol. 2011, 156, 1921–1933.
  52. Kalaji, H.M.; Jajoo, A.; Oukarroum, A.; Brestic, M.; Zivcak, M.; Samborska, I.A.; Cetner, M.D.; Łukasik, I.; Goltsev, V.; Ladle, R.J. Chlorophyll a fluorescence as a tool to monitor physiological status of plants under abiotic stress conditions. Acta Physiol. Plant. 2016, 38, 102.
  53. Mao, Y.-B.; Liu, Y.-Q.; Chen, D.-Y.; Chen, F.-Y.; Fang, X.; Hong, G.-J.; Wang, L.-J.; Wang, J.-W.; Chen, X.-Y. Jasmonate response decay and defense metabolite accumulation contributes to age-regulated dynamics of plant insect resistance. Nat. Commun. 2017, 8, 13925 .
  54. 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, 827.
  55. Fischer, W.; Calderón, M.; Haag, R. Hyperbranched polyamines for transfection. Nucleic Acid Transfection 2010, 95–129.
  56. Chen, D.; Shao, Q.; Yin, L.; Younis, A.; Zheng, B. Polyamine function in plants: Metabolism, regulation on development, and roles in abiotic stress responses. Front. Plant Sci. 2019, 9, 1945.
  57. Alcázar, R.; Bueno, M.; Tiburcio, A.F. Polyamines: Small amines with large effects on plant abiotic stress tolerance. Cells 2020, 9, 2373.
  58. Capell, T.; Bassie, L.; Christou, P. Modulation of the polyamine biosynthetic pathway in transgenic rice confers tolerance to drought stress. Proc. Natl. Acad. Sci. USA 2004, 101, 9909–9914.
  59. Zhao, J.; Wang, X.; Pan, X.; Jiang, Q.; Xi, Z. Exogenous putrescine alleviates drought stress by altering reactive oxygen species scavenging and biosynthesis of polyamines in the seedlings of Cabernet Sauvignon. Front. Plant Sci. 2021, 12.
  60. Vickers, N.J. Animal communication: When i’m calling you, will you answer too? Curr. Biol. 2017, 27, R713–R715.
  61. Gill, S.S.; Tuteja, N. Polyamines and abiotic stress tolerance in plants. Plant Signal. Behav. 2010, 5, 26–33.
  62. Saxena, R.; Adhikari, D.; Goyal, H. Biomass-based energy fuel through biochemical routes: A review. Renew. Sustain. Energy Rev. 2009, 13, 167–178.
  63. Ruppert, A.M.; Weinberg, K.; Palkovits, R. Hydrogenolysis goes bio: From carbohydrates and sugar alcohols to platform chemicals. Angew. Chem. Int. Ed. 2012, 51, 2564–2601.
  64. Cao, X.; Zhu, C.; Zhong, C.; Hussain, S.; Zhu, L.; Wu, L.; Jin, Q. Mixed-nitrogen nutrition-mediated enhancement of drought tolerance of rice seedlings associated with photosynthesis, hormone balance and carbohydrate partitioning. Plant Growth Regul. 2018, 84, 451–465.
  65. Rosa, M.; Prado, C.; Podazza, G.; Interdonato, R.; González, J.A.; Hilal, M.; Prado, F.E. Soluble sugars: Metabolism, sensing and abiotic stress: A complex network in the life of plants. Plant Signal. Behav. 2009, 4, 388–393.
  66. Sicher, R.C.; Timlin, D.; Bailey, B. Responses of growth and primary metabolism of water-stressed barley roots to rehydration. J. Plant Physiol. 2012, 169, 686–695.
  67. Sperdouli, I.; Moustakas, M. Interaction of proline, sugars, and anthocyanins during photosynthetic acclimation of Arabidopsis thaliana to drought stress. J. Plant Physiol. 2012, 169, 577–585.
  68. Basu, P.; Ali, M.; Chaturvedi, S. Osmotic adjustment increases water uptake, remobilization of assimilates and maintains photosynthesis in chickpea under drought. Experiment 2007, 45, 261–267.
  69. Keunen, E.; Peshev, D.; Vangronsveld, J.; Van Den Ende, W.; Cuypers, A. Plant sugars are crucial players in the oxidative challenge during abiotic stress: Extending the traditional concept. Plant Cell Environ. 2013, 36, 1242–1255.
  70. Saddhe, A.A.; Manuka, R.; Penna, S. Plant sugars: Homeostasis and transport under abiotic stress in plants. Physiol. Plant. 2021, 171, 739–755.
  71. Krasavina, M.S.; Burmistrova, N.A.; Raldugina, G.N. The role of carbohydrates in plant resistance to abiotic stresses. In Emerging Technologies and Management of Crop Stress Tolerance; Elsevier: Amsterdam, The Netherlands, 2014; pp. 229–270.
  72. Chen, T.H.; Murata, N. Glycinebetaine: An effective protectant against abiotic stress in plants. Trends Plant Sci. 2008, 13, 499–505.
  73. Giri, J. Glycinebetaine and abiotic stress tolerance in plants. Plant Signal. Behav. 2011, 6, 1746–1751.
  74. Sakamotto, A.; Muratta, N. The role of glycine betaine in protection of plants against stress: Clue from transgenic plants. Plant Cell Environ. 2002, 25, 163–171.
  75. Quan, R.; Shang, M.; Zhang, H.; Zhao, Y.; Zhang, J. Improved chilling tolerance by transformation with betA gene for the enhancement of glycinebetaine synthesis in maize. Plant Sci. 2004, 166, 141–149.
  76. Prasad, K.; Sharmila, P.; Kumar, P.; Saradhi, P.P. Transformation of Brassica juncea (L.) Czern with bacterial codA gene enhances its tolerance to salt stress. Mol. Breed. 2000, 6, 489–499.
  77. Holmström, K.O.; Somersalo, S.; Mandal, A.; Palva, T.E.; Welin, B. Improved tolerance to salinity and low temperature in transgenic tobacco producing glycine betaine. J. Exp. Bot. 2000, 51, 177–185.
  78. Niesche, R.; Haase, M. Emotions and ethics: A Foucauldian framework for becoming an ethical educator. Educ. Philos. Theory 2012, 44, 276–288.
  79. Khedr, R.A.; Sorour, S.G.R.; Aboukhadrah, S.H.; El Shafey, N.M.; Abd Elsalam, H.E.; El-Sharnouby, M.E.; El-Tahan, A.M. Alleviation of salinity stress effects on agro-physiological traits of wheat by auxin, glycine betaine, and soil additives. Saudi J. Biol. Sci. 2022, 29, 534–540.
  80. Fahy, E.; Subramaniam, S.; Brown, H.A.; Glass, C.K.; Merrill, A.H.; Murphy, R.C.; Raetz, C.R.; Russell, D.W.; Seyama, Y.; Shaw, W. A comprehensive classification system for lipids1. J. Lipid Res. 2005, 46, 839–861.
  81. Okazaki, Y.; Saito, K. Roles of lipids as signaling molecules and mitigators during stress response in plants. Plant J. 2014, 79, 584–596.
  82. Munnik, T.; Zarza, X. Analyzing plant signaling phospholipids through 32 P i-labeling and TLC. In Plant Lipid Signaling Protocols; Springer: Berlin/Heidelberg, Germany, 2013; pp. 3–15.
  83. Welti, R.; Li, W.; Li, M.; Sang, Y.; Biesiada, H.; Zhou, H.-E.; Rajashekar, C.; Williams, T.D.; Wang, X. Profiling membrane lipids in plant stress responses: Role of phospholipase Dα in freezing-induced lipid changes in Arabidopsis. J. Biol. Chem. 2002, 277, 31994–32002.
  84. Viehweger, K.; Dordschbal, B.; Roos, W. Elicitor-activated phospholipase A2 generates lysophosphatidylcholines that mobilize the vacuolar H+ pool for pH signaling via the activation of Na+-dependent proton fluxes. Plant Cell 2002, 14, 1509–1525.
  85. Drissner, D.; Kunze, G.; Callewaert, N.; Gehrig, P.; Tamasloukht, M.B.; Boller, T.; Felix, G.; Amrhein, N.; Bucher, M. Lyso-phosphatidylcholine is a signal in the arbuscular mycorrhizal symbiosis. Science 2007, 318, 265–268.
  86. Mandal, M.K.; Chandra-Shekara, A.; Jeong, R.-D.; Yu, K.; Zhu, S.; Chanda, B.; Navarre, D.; Kachroo, A.; Kachroo, P. Oleic acid–dependent modulation of NITRIC OXIDE ASSOCIATED1 protein levels regulates nitric oxide–mediated defense signaling in Arabidopsis. Plant Cell 2012, 24, 1654–1674.
  87. Upchurch, R.G. Fatty acid unsaturation, mobilization, and regulation in the response of plants to stress. Biotechnol. Lett. 2008, 30, 967–977.
  88. Markham, J.E.; Lynch, D.V.; Napier, J.A.; Dunn, T.M.; Cahoon, E.B. Plant sphingolipids: Function follows form. Curr. Opin. Plant Biol. 2013, 16, 350–357.
  89. Ballaré, C.L. Jasmonate-induced defenses: A tale of intelligence, collaborators and rascals. Trends Plant Sci. 2011, 16, 249–257.
  90. Kang, L.; Wang, Y.S.; Uppalapati, S.R.; Wang, K.; Tang, Y.; Vadapalli, V.; Venables, B.J.; Chapman, K.D.; Blancaflor, E.B.; Mysore, K.S. Overexpression of a fatty acid amide hydrolase compromises innate immunity in Arabidopsis. Plant J. 2008, 56, 336–349.
  91. Sun, Y.; Li, Y.; Sun, X.; Wu, Q.; Yang, C.; Wang, L. Overexpression of a Phosphatidylinositol-Specific Phospholipase C Gene from Populus simonii× P. nigra Improves Salt Tolerance in Transgenic Tobacco. J. Plant Biol. 2022, 65, 365–376.
  92. Arisz, S.A.; van Wijk, R.; Roels, W.; Zhu, J.-K.; Haring, M.A.; Munnik, T. Rapid phosphatidic acid accumulation in response to low temperature stress in Arabidopsis is generated through diacylglycerol kinase. Front. Plant Sci. 2013, 4, 1.
  93. McLoughlin, F.; Arisz, S.A.; Dekker, H.L.; Kramer, G.; De Koster, C.G.; Haring, M.A.; Munnik, T.; Testerink, C. Identification of novel candidate phosphatidic acid-binding proteins involved in the salt-stress response of Arabidopsis thaliana roots. Biochem. J. 2013, 450, 573–581.
More