Open access peer-reviewed chapter

Protein Metabolism in Plants to Survive against Abiotic Stress

Written By

Bharti Thapa and Abhisek Shrestha

Reviewed: 01 February 2022 Published: 15 May 2022

DOI: 10.5772/intechopen.102995

From the Edited Volume

Plant Defense Mechanisms

Edited by Josphert Ngui Kimatu

Chapter metrics overview

400 Chapter Downloads

View Full Metrics

Abstract

Plants are frequently subjected to several abiotic environmental stresses under natural conditions causing profound impacts on agricultural yield and quality. Plants can themselves develop a wide variety of efficient mechanisms to respond environmental challenges. Tolerance and acclimation of plants are always related to significant changes in protein, cellular localization, posttranscription, and posttranslational modifications. Protein response pathways as well as pathways unique to a given stress condition shared by plants under different stressed environment are discussed in this chapter. The various signaling of protein such as fluctuation, overexpression, and silencing of the protein gene are observed to be modulated in drought-tolerant plants. Similarly, gene expression, RNA processing, and metabolic process take place to cope with drought conditions. For adaption in water-submerged conditions, plants undergo reactive oxygen species (ROS), cell wall modification, proteolysis, and post-recovery protein metabolism. Heat shock protein and protein and lipid contents vary and play pivotal role in resisting low and high temperatures. In a nutshell, this paper provides an overview of several modification, synthesis, degradation, and metabolism of protein in plants to cope with and revive again to normal growing conditions against abiotic stress, emphasizing drought, submerged, extreme cold, and heat temperatures.

Keywords

  • protein
  • abiotic stress
  • tolerance
  • acclimation
  • yield

1. Introduction

Plants have developed a wide variety of highly sophisticated and efficient mechanisms to sense, respond, and acclimatize to a wide range of environmental changes. They have responded by activating tolerance mechanisms at multiple levels of organization (molecular, tissue, anatomical, and morphological), through the adjustment of membrane systems and cell wall architecture. This includes altering the cell cycle and rate of cell division and also by metabolic tuning [1]. Many molecular genes are induced and repressed by abiotic stresses at molecular level involving a precise regulation of extensive stress-gene networks [2, 3], and their products may function in stress response and tolerance at cellular level. Proteins involved in multiple protein functions, such as biosynthesis of osmo-protectant compounds, detoxification enzyme systems, proteases, transporters and chaperones, act as a first line of direct protection from stress. Moreover, regulatory proteins, for instance, transcription factors, protein phosphatases and kinases, and signaling molecules activation are essential in regulation of signal transduction and stress-responsive gene expression [4, 5].

Generally, observed tolerance responses toward abiotic stress in plants are composed of stress-specific response mechanisms and adaptive responses that confer strategic advantages in adverse conditions. In energy maintenance, general response mechanisms related to central pathway are involved, including calcium signal cascades [6], reactive oxygen species (ROS) signaling elements [7, 8], and energy deprivation signaling (energy sensor protein kinase, SnRK1) [9]; and induction of these central pathways is observed during plant acclimation toward different stress. Protein kinase SnRK1, despite being central metabolic regulator of the expression of genes related to energy-depleting conditions, also get activates when plants face different sorts of abiotic stresses such as drought, salt, flooding, or nutrient depravation [10, 11]. SnRk1 kinases alter over 1000 stress-responsive genes expression allowing the re-establishment of homeostasis by repressing energy consuming processes, thus promoting stress tolerance [10, 12]. Optimization of cellular energy resources during stress for plant acclimation has been found to be imperative; and partially arrested energetically expensive process, such as reproductive activities, translation, and some biosynthetic pathways [13]. For instance, in maize, during salt stress and potassium deficiency stress, nitrogen and carbon assimilations are impaired; also, the synthesis of free amino acids, chlorophyll, and protein is also affected [14, 15]. After cessation of energy-expensive process, energy resources can be redirected to activate protective mechanisms [16].

1.1 Plant stress tolerance and resistance

Plants are sessile organisms, which are continuously being confronted with several detrimental factors rising from ever-changing environment, and to cope with these problems, they have developed sophisticated and delicate defense mechanisms. In fact, diverse defense signal including the production of reactive oxygen species (ROS), change in redox potential or cellular level of Ca2+ ion, disruption of ion homeostasis, and membrane fluidity adjustments are activated [17, 18]. Once external stress is sensed via specific receptors, foreign signal is induced into intracellular downstream signaling pathways including the activation of protein kinase or phosphatase, stimulation of downstream target proteins, and biosynthesis of phytohormones for the control of plant growth/development [19, 20].

1.2 Role of amino acid during stress

Gene expression can be adversely affected by salinity, drought, and temperature stress, and many genes coding for enzymes involved in cellular metabolism are differentially expressed upon stress, thus modeling some stress-related transcription factors to induce changes in stress-associated metabolite levels [4].

For the synthesis of secondary metabolites and signaling molecules, several amino acids can act as precursors, for instance, polyamines are derived from Arg [21], Met synthesized the plant hormone ethylene [22], and conversion of Lys to N-hydroxy pipecoline is necessary for immune signaling [17, 23]. Moreover, several aromatic amino acids, such as Phe, Tyr, and Trp, or intermediates of their synthesis pathways produce a broad spectrum of secondary metabolites possessing multiple biological functions and health-promoting properties [24]. Usually, plants exposed to different abiotic stresses tend to accumulate free amino acids [25, 26], as exemplified to this response, [27] reported extensive accumulation of amino acid in response to drought stress in maize, cotton, tomato, and the resurrection plant. Also, recent studies conducted by [26, 28, 29, 30] suggest increment in free amino acids as a result of autophagy and abscisic acid. Similarly, plants surviving in stressed environment can use amino acids as an alternative for mitochondrial respiration substrates during inadequate carbohydrate supply due to a decrease in photosynthesis rates [29, 31]. Although ambiguity still remains for the specific role of catabolic pathways, the degradation pathways for Lys and the branched-chain amino acids Val, Leu, and Ile have already been identified as crucial factors for dehydration tolerance for Arabidopsis [32]. And, soon after reviving of plants to favorable growth conditions, reprogramming their metabolism to switch back for survival and active growth is necessary.

Members of the AP2/EREBP (Apetala2/ethylene-responsive element binding protein) family of transcription factors, CBF/DREB1 proteins (C-repeat binding factor or dehydration responsive element binding proteins), such as CBF1/DREB1B, CBF2/DREB1C, and CBF3/DREB1A play an important role in the transcriptional response to osmotic stress [33, 34, 35, 36, 37] and stated improved tolerance of Arabidopsis to freezing, drought, and/or salt stress via overexpression of these transcription factors, and [38, 39] supported above researchers confronting plants overexpression of CBF/DREB1 accumulated higher levels of proline and soluble sugars (glucose, fructose, sucrose, and raffinose) when grown under normal growth conditions and during cold acclimation. Conclusion made from [39, 40] suggested overall metabolic profile of CBF3/DREB1A overexpressers grown at normal growth temperatures resembled that of cold-exposed plants.

Advertisement

2. Proteomic overview on abiotic responses in plants

The biological research of abiotic stress in plants can be studied in broad range of transcriptomic and proteomic-based, provides the comprehensive information, during and following stress condition, on alteration of gene expression and proteome profile, the study about 30 min to 1 day after induction, and time lapse between transcriptomic and proteomic suggest more than 50% of genes responsive to flood, heat, and other stress were found to encode transcription regulators [41].

2.1 Protein metabolism in plant roots and shoots due to drought stress

Prolonged water deficit in the soil causes drought, which vastly affects the metabolism and physiological function in growing plant especially roots and responsive for water supply from soils to leaves and photosynthesis, respectively. Most of the proteomic evidence has been noticed due to drought condition, six steps prominently occur in the responsive drought stress. Signaling and sensing receptor, yet not specially but drought-responsive photoreceptor, phytochrome C1, found in maize, phytochrome gene (i.e., PHYA, PHYB, and PHYE) in Arabidopsis, believed to regulated the transcription of light responsive genes by modulating the activity of several transcription factors and involved in suppressing drought tolerance. Other signaling cascades, G protein subunits (alpha and beta), small G protein (e.g., Ras-related protein Rab7 and Ras-related nuclear protein Ran), and Ran-binding protein 1 (play important role in cell cycle and DNA synthesis) regulate positive role in drought stress [42], involved in vesicle trafficking, intercellular signaling, polar growth, plant hormone signal cross talk, and stress response [43]. The PgRab7 gene was upregulated by dehydration in Pennisetum glaucum [44], while overexpression of the peanut AhRabG3f exhibited an enhanced tolerance to drought stress in transgenic peanut (Arachis hypogaea L.) [45] but negative role in Arabidopsis. Calcium-binding proteins (CaBs), such as calmodulin (CaM), calcium-sensing receptor (CaSR), calreticulin (CRT), and calcium-dependent protein kinase (CDPK), enhanced the survival of Triticum aestivum [46], and several protein kinases (e.g., serine/threonine-protein kinase, germinal center kinase (GCK)-like kinase MIK, receptor-like protein kinase HERK 1-like, phototropin family protein kinase, and salt-inducible protein kinase), imply their role in drought response signaling pathway, in addition to phosphorylation level of protein phosphate 2C (PP2C), acts as negative regulator for plant drought tolerance in the abscisic acid (ABA) signaling pathway, which can inhibit the activity of SnRK, leading to a decrease of the phosphorylation of its substrates in the signaling cascade [47]. Similarly, 14-3-3 protein availability fluctuation shows the drought condition and reported that drought stress can directly alter the abundance of 14-3-3 proteins [48]. In addition, overexpression or silencing of the 14-3-3 protein genes can modulate drought tolerance of transgenic plants (e.g., Gossypium hirsutum and Arabidopsis) [49, 50].

Phytohormones play important role in signal transduction pathways such as drought-increased ethylene-responsive transcription factor (ERF) in Gossypium herbaceum [2] (and some members of drought-responsive auxin-binding protein (ABP) family in Quercus robur [51] Zea mays (41) and polar clones [52]. Under drought stress, ERF gene was induced in G. herbaceum [53, 54], and its overexpression in various plants, such as sugarcane SodERF3 overexpression in tobacco, tomato TERF1 in rice, and Brassica rapa BrERF4 in Arabidopsis, can improve plant drought tolerance. BpERF11 was found to negatively regulate osmotic tolerance in Betula platyphylla [42] ABP members (i.e., ABP2, ABP20, and ABP19a) are in response to drought stress, additionally TRIP-1 was phosphorylated by the brassinosteroid (BR)-insensitive I (BRI-1) protein, drought-increased TGF-β-receptor interacting protein 1 (TRIP1) was found in Sporobolus stapfianus and triggered the BR signaling pathways in water deficit condition.

Gene expression plays important role in the transcriptional regulatory networks, requires chromatin structure modification, i.e., histone, major protein of chromatin, and regulates the expression and high mobility group protein (HMG), involved in cell cycle progression. Among several histones, histone H1 was decreased in a drought-sensitive Z. mays cultivar, but increased in a drought-tolerant one [55], and H2B histone H1 was decreased in a drought-sensitive Z. mays cultivar, but increased in a drought-tolerant one [55]. Similarly, the phosphorylation level of HMG was significantly decreased in a drought-tolerant wheat cultivar, but increased in a drought-sensitive one [42] that reduced its binding to DNA, inhibiting replication and transcription.

Several RNA processing-related proteins changed over the stress condition, represent the critical for plants to cope with. Five glycine-rich RNA-binding proteins (GR-RBPs) increased with drought and three GR-RBPs decreased with drought, which bind to RNA molecules for transcriptional gene regulation and suspected to function in the regulation of specific gene expression. For instances, transgenic rice consists of GR-RBPs gene showing higher yield and drought recovery rate as compared with wild rice [56], besides overexpressed in Camelina sativa, reduced the drought tolerant. Similarly, S-like ribonucleases (RNases) specialized function as stress regulation, defense against microorganisms, phosphate scavenging, and even nitrogen storage, increased in rice under drought [57]. Additionally, an intron splicing-related protein, and maturase K (MatK) and multiple organelle RNA editing factor 9, involved in RNA editing in mitochondria and plastids, was found fluctuated in Brassica napus with the extension of drought stress [58] indicating the transcriptional regulation.

Most fundamental metabolic process to cope with drought stress, a plant can attribute to protein synthesis and turnover. Several proteins are involved in protein biosynthesis, such as ribosomal protein (RP), elongation factor (EF), translation initiation factor (TIF), tRNA synthase (TRS), and ribosome recycling factor (RRF), beneficial to protein synthesis, besides protein folding and processing varies cultivars and species. Instances, peptidyl-prolyl cis-trans isomerases (PPIases) were significantly increased in Oryza sativa [6] and Q. robur [51], but decreased in a drought-sensitive cultivar of Phaseolus vulgaris [59]. Protein disulfide isomerases (PDIs) were increased in barley and B. napus, but decreased in Agrostis stolonifera, Q. robur, and poplar. Additionally, ER-luminal binding protein (BiP), trigger factor-like protein (TIG), most heat shock proteins (HSPs), and other molecular chaperones (i.e., calnexin, endoplasmin) were increased, but T-complex protein and HSP70-HSP90 organizing protein were decreased in drought-treated leaves. This protein helps in maintaining the normal protein folding, repairing, and renaturation of the stress-damaged protein, whereas HSP must popular function in protein folding in Arabidopsis and yeast to improve the drought tolerance.

Protein degradation, process of removing the abnormal, damaged protein, and maintenance of certain level of regulatory proteins during drought, includes the components such as ubiquitin/26S proteasomes, small ubiquitin-like modifier (E3 SUMO) ligase, and proteases/peptidases (ATP-dependent Clp proteas, cysteine proteinase, zinc metalloprotease, aspartic proteinase, serine carboxypeptidase, and aminopeptidases (APs)). These components show positive response in P. vulgaris [59] Hordeum vulgare [60], B. napus [58], and Medicago sativa [61] under drought condition, involved in ubiquitination, exhibited significantly increased values in drought tolerant and decreased in drought-sensitive leaves.

Due to drought condition, it interrupts the normal cellular mechanism, results to produce the ROS. Plants evolve diverse mechanism to keep ROS homeostasis in cells, including antioxidative enzymes, e.g., SOD (first defense mechanism by converting O2 into H2O2) and CAT (convert H2O2 into H2O and O2) and chemical antioxidant (e.g., glutathione and ascorbate). Diverse abundance of SODs in cystolic, peroxisomeas as well as in chloroplast helps in the drought tolerance and avoidance. For instance, increment of cystolic Cu-Zn SODs drought avoidance CT9993 and drought tolerance IR62266), while e chloroplast Cu-Zn SODs were increased in CT9993, but decreased in IR62266 [57], additionally in cultivar of Malus domestica, Cu-Zn SOD decreased and Fe SOD increased [54]. Similarly, in Ascorbate-Glutathione (AsA-GSH) pathways, the ascorbate peroxidase (APX) reduces H2O2 to H2O using ascorbate (AsA) as an electron donor, then the oxidized AsA is restored by monodehydroascorbate reductase (MDHAR), dehydroascorbate reductase (DHAR), and glutathione reductase (GR) [62]. GR catalyzes the reduction of glutathione disulfide (GSSG) to the sulfhydryl form GSH and work out for the drought tolerance. In addition, some proteins that involved in glutathione-mediated ROS scavenging: glyoxalse (GLO), phospholipid hydroperoxide glutathione peroxidase, glutamate-cysteine ligase (GCL), glutaredoxin (Grx), and monothiol. GLo catalyzes the detoxification of methylglyoxyl, whereas GCL, first enzymes in GSH biosynthesis pathways found in drought stressed B. napus [58]. Additionally, Prx/Trx also catalyzes the reduction of H2O2, whose abundance response to drought. Trx-linked enzyme, methionine sulfoxide reductase (MSR), involved in conversion of methionine sulfoxide to methionine, protects cells and tissues from H2O2-induced stress. Besides, Glutathione Peroxidase/Glutathione S-Transferase pathways, GPX catalyzes the reduction of H2O2 using Trx [63], and GST catalyzes conjugation reactions between GSH and a number of xenobiotics, playing a crucial role in the degradation of toxic substances. To cope with drought, GPXs found increased in Boea hygrometrica [64], E. elongatum [65], and B. napus [58].

There will be occurrence of pathogen when left plant for water deficit condition, but some pathogenesis-related protein, namely chitinase, disease resistance protein (DRP), polyphenol oxidase (PPO), oryzacystain, pathogen defense-related protein 10 (PR10), and disease resistance gene analog PIC15 increased in the response of drought condition. These proteins act as the pathogen by acting on insect exoskeleton and fungi cell walls, catalyzing the oxygen-dependent oxidation of phenols to quinines’ during plant defense, acting as cysteine proteinase inhibitor in the phytocystatin family of proteinase inhibitors. For example, overexpression of oryzacystain gene in Tobacco displayed an increase of drought tolerance by improving total SOD and guaiacol POD activities.

Osmotic regulation will be hindered due to exposed to water deficit, but important osmotic homeostasis-related protein, namely embryogenesis abundant (LEA) protein, dehydrin (DHN), and betaine aldehyde dehydrogenase (BADH), which function as cellular protectants to stabilize cellular components, protein structure through detergents and chaperone like properties, act as calcium buffer. LEA proteins were also increased in Z. mays [55] and B. napus [58] under certain drought conditions.

Cell division and cell wall formation decreased due to decrease of phosphorylation of several protein (cell division cycle protein, division protein ftsZ1, and cyclin A2) when exposed to drought, which implies the suppression of cell growth. Cytoskeleton and cell wall component require for cell division, morphogenesis, and signal transduction, while cytoskeleton protein, namely actin, kinesin motor protein, tubulin, profilin, actin depolymerizing factor, and fibrillin to check the cell growth during stress. Additionally, the translationally controlled tumor protein homolog (TCTP) is a Ca2+-binding protein, which protect against stress and apotosis, cell growth, and microtubule organization, which was significantly drought increased in H. vulgare [60], T. aestivum [66], and B. napus [58], which would facilitate plant adaptation to drought stress.

Cell wall extensibility was directly affected by water loss, while cell wall polysaccharide synthesis/hydrolysis, lignin biosynthesis, and cell wall loosening in leaves were drought-responsive enzymes. Two enzymes, glycosylated polypeptide and pectinacetylesterase, involved for polysaccharides synthesis, another two enzymes xylanase inhibitor and polygalacturonase inhibitor, involved in polysaccharide hydrolysis inhibition. Three lignin biosynthesis-related proteins, phenylalanine ammonia-lyase (PAL), caffeic acid 3-O-methyltransferase, and caffeoyl-CoA O-methyl-transferase, catalyze the transformation of phenylalanine to cinnamylate of lignin biosynthesis, while two drought-increased cell wall structural proteins (i.e., glycine-rich protein and fasciclin-like arabinogalactan protein) enhance cell wall synthesis in response to drought by providing the UDP-glucose directly to the cellulose synthases and/or callose synthases [67], hence, improve the mechanical strength for minimizing water loss and cell dehydration. Another important activity of cell wall loosening/expansion, important aspect in the adaption to drought, which was related enzymes, polygalacturonase/pectin depolymerase (PG) in O. sativa [6] and xyloglucan endotransglycosylase (XTH), where PG degrade pectin, while XTH can cleave and reform the bonds between xyloglucan chains to regulate cell wall rigidity.

Membrane trafficking localized in mitochondrion, plasma, and vacuole. Two mitochondrion protein carriers (dicarboxylate/tricarboxylate carrier (DTC) and 2-oxoglutarate/malate carrier protein (OMC)), catalyze the transport of various metabolites (e.g., dicarboxylates, tricarboxylates, amino acids, and keto acids), play important role in gluconeogenesis, nitrogen metabolism, as well as biotic stress [68]. Another, Remorin, aquaporin and PEG, plant-specific plasma membrane protein have importance in plant-microbe interaction and signal transduction [69]. In addition, vacuolar H+-pyrophosphatase (V-PPase), vacuolar-ATPase (V-ATPase), and ABC transporter ATPase, important for or translocating H+ into the vacuoles to generate a gradient of H+, which provide a driving force for the accumulation of ions and other solutes in the vacuole and function for abiotic stress.

Photosynthesis inhibition is the primary detrimental effect due to drought stress, and related protein decrease. To cope with this situation, drought increased protein involved in the photoreaction and Calvin cycle in leaves. Light-harvesting chlorophyll a/b-binding proteins (LHCB), involved in ABA signaling partially by modulating ROS homeostasis, besides, abundance of sedoheptulose-1,7-bisphosphatase (SBPase) and carbonic anhydrase (CA), catalyzes the reversible hydration of CO2, and influence in internal conductance and abundance of protein involved in photorespiration, significantly increases and decreases glycolate oxidase, glycine dehydrogenase, serine glyoxylate aminotransferase, and serine transhydroxymethyl transferase, aminomethyl transferase (AMT), and glycine dehydrogenase to adapt the drought stress. The mechanism in photorespiration can protect the photosynthesis from photoinhibition and prevent ROS accumulation in green tissues.

Involvement in carbohydrate and energy metabolism is important step to cope with drought condition. Phosphoglucomutases (PGluMs), fructose-bisphosphate aldolase (FBPA) in glycolysis and aconitate hydratases in TCA cycle increased in drought condition, which inhibit the accumulation of sugars as osmolyte or energy source for recovery, while the increase of glycolysis and TCA may act as a strategy for providing energy during the activation of stress defenses, especially when the photosynthesis was inhibited. The change in mitochondrial electron transport chain and ATP synthesis related protein implies ability to enhance energy production to maintain physiological activity and inhibit stress damage.

Due to the drought condition, nitrogen assimilation decreased in the reduction of NR, GS, and GOGAT, which was main reason for yield reduction. Similarly, the decline of aspartate aminotransferase (AST) and alanine aminotransferase (ALAT) indicates that drought stress inhibits the amino-acid metabolism and synthesis of other metabolites. At the same time, S-adenosyl-l-methionine (SAM) cycle was generally increased in leaves, including drought-increased. 5-methyltetrahydropteroyltriglutamate-homocysteine methyltransferase (MetE), S-adenosyl-l-homocysteine hydrolase (SAHase), S-adenosylmethionine synthase (SAMS), and methionine synthase (MS), which implies that it enhances the methionine and osmo-regulant metabolism for plants to cope with drought stress.

Acetyl-coenzyme A carboxylase carboxyl transferase, acyl carrier protein, enoyl-acyl carrier protein reductase, and lipoxygenase 6 involved in fatty acid biosynthesis, and enzymes thiolase I, thiolase II, and acyl-CoA dehydrogenase used for fatty acid degradation. Greater composition of unsaturated fatty acid in membrane lipids contribute to superior leaf dehydration tolerance and maintain membrane integrity and preserve cell compartmentation under water stress, in addition, two flavonoid biosynthesis related proteins (i.e., chalcone isomerase (CHI) and dihydroflavonol-4-reductase) involved in secondary metabolism were also changed in response to drought.

2.2 Protein metabolism under flooding and submerged stress condition

Deprivation of the soil oxygen due to consequence of flooding and forced the plant to shift from aerobic to anerobic respiration [70], which regenerate NAD+, through ethanol fermentation by selectively synthesizing flooding-inducible proteins involved in sucrose breakdown, glycolysis, and fermentation [13]. Several glycolysis-related proteins, including fructose-bisphosphate aldolase, phosphoglycerate kinase [64], glyceraldehyde-3-phosphate dehydrogenase [71], enolase [72], sugar isomerase, phosphofructo-kinase [73], and pyruvate kinase [72] are increased in soybean under flooding stress, indicate the glycolysis and fermentation pathways activation, initiating for plants protecting plant from flooding induced damage, whereas decrease of fructose-1,6-bisphosphate aldolase and sucrose-fructan 6-fructosyl transferase in wheat show response to flooding stress. Othersides, fermentation under anaerobic condition, influence the accumulation of fermented related proteins such as alcohol dehydrogenase (ADH) and pyruvate carboxylase, and indicates that activation of the alcohol fermentation pathways, to cope the hypoxic condition. The conversion of acetaldehyde to ethanol by ADH with concurrent reoxidation of NAD+ for the continuation of glycolysis. The fermentation related enzyme pyruvate decarboxylase, and aldehyde dehydrogenase increase to accelerate the energy production via nonoxidative pathways, even growth is suppressed.

In other sides, flooding stress induces impairment of the electron transport chain in plants. Protein related to complexes II, IV, and V of the electron transport chain decreased in abundance and while, succinate-semialdehyde dehydrogenase, 2-oxoglutarate dehydrogenase, and gamma-amino butyrate are significantly increased, which are required for energy production via non-oxidative pathways [72]. Oxaloacetate produced in TCA cycle stimulates phosphoenol pyruvate synthesis and provides the indirect simulation for the continuation of glycolysis. Reduction of energy metabolism-related proteins, including citrate synthase, glutamate dehydrogenase, and adenosine kinase, in wheat roots under waterlogging stress [74]. In addition, energy-related proteins such as beta-amylase, malate dehydrogenase, fructose-1,6-bisphosphatase, and phosphoenol pyruvate carboxykinase are decreased in response to flooding stress, indicating that gluconeogenesis is suppressed in wheat under these conditions [10]. RuBisCo sub unit binding protein alpha sub unit and RuBisCO activate degraded and senescence in high ROS condition and decreased the chlorophyll content, results to decrease in net energy production.

ROS recognized as toxic byproduct of aerobic metabolism and controlled by anti-oxidants and anti-oxidative enzymes. The plant development of well-organized scavenging mechanism to overcome ROS toxicity likely to led to the use of reactive molecules as signal transducers in plant cells. ROS production in cellular organelles, such as plastids, mitochondria, and peroxisomes, involved in signaling cascades controlled by production and scavenging of ROS intermediates [27]. ROS scavengers, such as peroxidase, APX, cytosolic APX, and superoxide dismutase (SOD), linked to bio photon emissions and decreased photosynthesis and beneficial for normal metabolism and cell signaling.

Cell wall modification related proteins, namely polygalactouronase inhibitor-like and expansion-like B1-like proteins and cell wall synthesis related protein such as cinnamyl-alcohol dehydrogenase and cellulose synthase-interactive protein-like protein abundance response under water logged condition. Flooding stress induces the assimilation of methionine and promotes cell wall hydrolysis, thereby restricting growth so, under the waterlogged stress, cell wall synthesis related proteins decrease, cell wall loosening related protein increase and cell wall lignification is suppressed.

Proteolysis, protein folding and storage plays important role in the removing the flooding damage induced non-active proteins [40]. Heat shock proteins act as molecular chaperones in preventing protein aggregation, translocation of nascent chains across membranes, assembly or disassembly of multimeric protein complexes, and targeting proteins for lysosomal or proteasomal degradation [40]. The ubiquitin/proteasome-mediated proteolysis of enzymes involved in glycolysis and fermentation pathways may be negatively controlled under the hypoxic condition caused by flooding stress [40].

Post recovery protein metabolism is less studied but studied by [75] Gro-EL-like chaperone ATPase, 26 S proteasome regulatory subunit 7, 26 S regulatory subunit S 10B, and cyclophilin were decreased in seedlings recovering from flooding stress, whereas globulin-like protein, Kunitz trypsin protease inhibitor, and peptidyl-prolyl cis-trans isomerase 1 were increased, and soybean root recovers from flooding by altering cell structure, strengthening cell wall lignification, and scavenging toxic ROS.

2.3 Cold stress

One of the major abiotic stresses is cold or low temperatures (LTs) that severely affect plant growth and survival. Chilling or freezing with temperatures <20°C and < 0°C respectively can induce ice formation in plant tissues which causes cellular dehydration [39]. To be able to withstand in this adverse condition, plants adopt several strategies, such as production of more energy via activation of primary metabolisms, leveling up of antioxidants content and chaperones, and maintenance of osmotic balance by altering membrane structure [76].

2.3.1 Protein metabolism in cold stress

Several articles and reviews deals with the metabolic responses of plants at low temperature, where some attempted correlating metabolic and biochemical responses with cold tolerance. Solaw [77] noted correlative studies of biochemical changes does not enable understanding of cold acclimation (CA) leading to increased freezing tolerance and till date no any new approaches in molecular biology and genetics have been extensively enlisted on study of cold-tolerance and injury mechanisms. However, few studies of CA started focusing on some of the more rapid plant physiological and molecular responses subjected to LTs, which revealed that within the hours of LT exposure, plant and algal cells can rapidly initiate to alter membrane lipid composition [78], RNA [79], and protein content. These findings of rapid biochemical alterations in response to L T convince the rapid induction of freezing tolerance at inductive temperatures and by desiccation and ABA at non inductive temperatures [80] and ABA [10]. This suggests a possible molecular basis, at minimum, for the adjustment of metabolism to low nonfreezing temperature, and perhaps for freezing tolerance. Also, upon exposure to LT, it consists of repeated observations that a number of enzymes show shifts in isozymic composition, whereas both quantitative and qualitative differences in the protein content is shown by numerous electrophoretic studies between non-acclimated and cold acclimated tissues.

2.3.2 Enzyme variation

Compared with plants maintained at warm temperature, [20] reported changes in activity, freeze stability, and isozymic variation in plants subjected to LTs. He mentioned increased peroxidase activity in hardened stems of four widely unrelated woody species when electrophoretic techniques to separate enzymes from non-hardened and hardened tissues. Here, peroxidase isozymes present in hardened tissues were not found in other three non-hardened tissues. Similarly, during deacclimation, no change in peroxidases, glucose-6-phosphate, 6-phosphogluconate, and malate dehydrogenases was observed in willow stem [81], however, differences were observed in lactate dehydrogenase where the activity increased during deacclimation. Similarity as above findings was illustrated by [82] in which invertase enzyme in wheat leaves undergoes a shift from a lower-molecular weight form to a higher-molecular-weight form at LT. Different kinetic properties is exhibited in larger form functionally replacing small form in cold-hardened plants [83]. Also, Krasnuk and colleagues [6, 84] observed increased activity of a number of dehydrogenases associated with respiratory pathways, including glucose-6-phosphate dehydrogenase, lactate, and isocitrate dehydrogenase [6] during a comprehensive studies with alfalfa. Thus [85] suggests higher amounts of enzyme may increase in activity and soluble protein content indicating increased soluble protein content and enzyme activity could be part of the adjustment of metabolism to the kinetic constraints imposed by LTs.

A more recent study of glutathione reductase from spinach carried out by [26] demonstrated not only additional isozymic forms in cold-acclimated tissue but also increased activity, freezing stability, and altered kinetic behavior and the activity of this particular enzyme was decreased by freezing/thawing both in vitro and in vivo. However, the enzyme found in cold-acclimated plants was less sensitive than its counterpart from non-acclimated plants to freezing from nonacclimated plants. In contrary, kinetic parameters and freeze/thaw inactivation was observed identical in ferredoxin NADP reductase from nonacclimated and cold-acclimated wheat [86], whereas activity was increased during CA. Therefore, [85] illustrated the potential for alterations in enzymes in response to low temperature exposure and the apparent selective basis where such changes can occur.

Ribulose bisphosphate or oxygenase from winter rye is the best-characterized enzyme relative to non-acclimated and cold-acclimated plants. Early in vitro studies studied by [87] noted purified enzyme from both non- and cold-acclimated plants demonstrated an increased stability of catalytic activity to denaturants and storage at −25°C of the enzyme from cold-acclimated plants. Moreover, [87] presented evidence of a stability in vivo conformational change during low-temperature adaptation that was not altered by purification of the enzyme. Also, osmotic concentration of the purified enzyme caused a greater degree of aggregation through intermolecular disulfide bond formation of the large subunit from non-acclimated plants [4] also claimed, similar to rye, the enzyme purified from freezing sensitive and -tolerant potato species demonstrated structural differences that paralleled variation in freezing tolerance much in the same way. However, the study still remains in ambiguity for the stable change in conformation, kinetic properties, and differential cryostabilities of this enzyme from cold-acclimated leaves or cold tolerant potato species. Given that a single chloroplastic gene encodes for large subunit and not possess even a minute chance for the synthesis of an alternative cryostable large subunit from another gene. Also, in many plants, always, a small gene family codes the small subunit, a change in the small subunit may possess subtle effect on the cryostability and other properties of the holoenzyme. Equally possibility occurs in LT -specific posttranslational processing, although there is no evidence to support this concept. In addition to isozymic and conformational differences of enzymes in response to L T exposure, Griffith, stated supramolecular interactions can also be affected.

2.3.3 Protein content

According to [88], accumulation of soluble protein in cold-acclimated cortical bark cells of black locust trees was first correlated with freezing tolerance about 40 years ago. These study may not be explained as simply as stated in past, [88] suggests there are many reasons why some plants might accumulate soluble proteins during CA; but with the exception of a protoplasmic augmentation hypothesis without clear mechanistic rationale for conferring greater freezing tolerance for this hardening response. In temperate deciduous perennials like black locust could provide the nitrogen source for the accumulation of proteins in the cortical cells of the living bark, in expense of nitrogenous materials during senescence. Parsell and Lindquist [89] supports a possible functional role of the increased soluble protein in cold tolerance was the fact that an evergreen, red pine, also accumulated soluble proteins during the winter, similarly, cortical bark cells need not to act as vegetative storage in evergreens nevertheless, it cannot be refused that one or more minor components of the total protein content could function in freezing-tolerance mechanisms.

Most of the studies have confirmed the presence of new protein species in cold-acclimated and freezing-tolerant plants. When these plants are compared to non-acclimated plants, subtle shift in protein content in cold-acclimated tissues involving mostly the appearance and disappearance of minor bands in gel can be observed [85]. Existing evidence at present includes several studies of purified plasma membranes from non-acclimated and cold-acclimated tissue. Tzin and Galili [90] emphasized on declination of more than 20 proteins in cold-acclimated leaf plasma membranes, whereas 11 had increased their concentration, while 26 proteins were new and unique to membranes from hardened tissue, yet increased levels of high-molecular-weight glycoproteins were other alterations included during CA.

2.4 Heat stress

Some defense mechanisms can be triggered in response to several stresses, such as expression of obvious genes which were not expressed under normal situations, resulting increased synthesis of protein groups [60]. These groups in cases of heat are called as Heat Shock Proteins (HSPs), “Stress-induced proteins” or “Stress proteins” [49].

According to [14] declination in normal protein synthesis occurs when exposed to high temperatures, thereby increasing selective translation of mRNAs for characteristic sets of HSPs. Heat responding phenomena in plants generally observed with concomitant reduction in protein synthesis of new or constitutive HSPs. Jin et al. [35] observed reduction in total protein synthesis at of 40°C and above in soybean. The HSPs, in plants, consists an abundant group of low mo1 wt polypeptides with higher molecular weight families [91], where some of them found to function as chaperones minimizing high-temperature stress damage partially denaturing proteins and preventing breakdown or aggregation. Response toward heat includes increase in binding of ubiquitin to conglomerated high molecular weight protein [11] which both increase and decrease ubiquitin transcripts expression [17].

2.4.1 Role of heat shock proteins

Levitt et al. [92] reported formation and folding patterns of any protein in three dimensional structure determines their function and [93] favored above statement and confronted with the findings illustrating 50% of principle amino acids sequence is necessary for formation of three dimensional structure which signifies the importance of HSPs in folding of other proteins. As plants were induced in heat stress, HSPs protects cells from injury and facilitates their recovery and survival to normal growth conditions [49]. Also, [94] specified the role of HSps as molecular chaperones during heat stressed condition, apart from ensuring maintenance of correct protein structure, which is basically different than in non-thermal stress where proteins unfolding is not the primary effect and protection could occur in any ways.

Seo et al. [95] also focused the general role of HSPs as molecular chaperones, regulating folding and accumulation of proteins as well as localization and degradation in all plant and animal species, thus preventing the irreversible aggregation of other proteins and participate in refolding proteins during heat stress conditions [96].

Different HSPs with their unique role are described below:

2.4.1.1 Small heat shock proteins (sHSPs)

These contains a common alpha-crystallin domain containing 80–100 amino acid residues contained in the C-terminal region [97]. The characteristics features of this proteins is degradation of protein having unsuitable folding [11]. Another attributes that make it indifferent from other class is independency of its activity from ATP [98]. A recent review from [99] concluded the presence of some indications that sHSPs play crucial role in membrane quality control, thereby potentially contributing the maintenance of membrane integrity under stress conditions.

2.4.1.2 HSP60

This class, called as chaperonins, is known to be important in assisting plastid proteins such as Rubisco [100]. Some studies like [101] pointed out that they might participate in folding and aggregation of proteins that were transported to chloroplasts and mitochondria. These proteins are different from other proteins, after transcription and before folding, to prevent their aggregation [42].

2.4.1.3 HSP70

These HSP70 functions as chaperones, in almost all organisms, for newly formed proteins to check their accumulations as aggregates and folds in a proper way during their transfer to final location [102]. Furthermore, cooperation in the activity of HSp70 and sHSPs primarily function as molecular chaperone and play an important role in protecting plant cell from detrimental effects of heat stress [83] stressed on crucial role played by HSp70 and sHSP17.6 in the development of cross adaptation to temperature stress in grape vines induced by heat acclimation (HA) and cold acclimation (CA).

2.4.1.4 HSP90

The protein from HSP90 class shares the role in many chaperone complexes and has important role in signaling protein function and trafficking [89]. Furthermore, other important attributes retained by these class includes regulation of cellular signals, such as the regulation of glucocorticoid receptor (GR) activity [103].

2.4.1.5 HSP100

What makes it unique from other class is the reactivation of aggregated proteins [42] by re-solubilizing nonfunctional protein aggregates and also by degrading irreversibly damaged polypeptides [104, 105]. Members of this class are not restricted only to acclimation to high temperatures but also housekeeping functions necessary in chloroplast development are also provided by specific member (Figures 1 and 2) [106].

Figure 1.

Plant stress tolerance and resistance [100].

Figure 2.

Transcription [100].

Advertisement

3. Conclusion and future prospective

Abiotic stresses are major limiting factors for plant growth and yields and with various acclimation responses at morphological, physiological, metabolic, and molecular level coordinated by complicated regulatory networks comprising genes, phytohormones, ROS, and other signaling components. The abundance of ion channels protein and trans-membrane water found indicated the change in ions/osmotic balances, but the phenomenon was not observed in flooding conditions. In addition, the preventive measure against the oxidative damage caused due to ROS levels under abiotic stress, higher abundance of ROS scavengers plays a great role in this matter, whereas the abundance of ROS scavengers was low in the flooding condition. On the other hand, protein folding due to molecular chaperone and disease, defense-related proteins such as proteolytic enzymes and proteosomal factors under stress, indicating the refolding of denatured proteins and proteolytic elimination of damaged proteins. This review paper showed the different protein metabolism occurs during the metabolic stages, and secondary metabolism-associated proteins escape and tolerate mechanism under different abiotic stress.

At the recovery stages, increased lignin biosynthesis results in enhanced mechanical strength by hardening of cell wall. Changes in abundance for cyto-skeletons associated proteins can be overlooked upon compensation against the reduced cell size as well as repairing injuries caused by drought and flood stress. Moreover, the levels of proteins related to de novo proteins synthesis, growth-related signaling and secondary metabolism are enhanced during flood replenishment of the stress induced effects. These stress-induced effects can be recovered by compensatory mechanism.

Only after proteomic studies could make us aware about the mechanism involved in abiotic stress condition. Analyzing of the plant response and abundance of protein and stress tolerant crops will lead to better understanding of the mechanism of plant to overcome the stress and recover. Moreover, some proteins showed the dynamic changes depending on plant species and stress intensity, which gives a clear interpretation of the mechanism in stress response. The integration of those finding from physiological, gene expression, and other large scale “omics” will help us to establish molecular networks of stress response and tolerance.

References

  1. 1. Atkinson NJ, Urwin PE. The interaction of plant biotic and abiotic stresses: From genes to the field. Journal of Experimental Botany. 2012;63(10):3523-3543
  2. 2. Chinnusamy V, Zhu J, Zhu JK. Cold stress regulation of gene expression in plants. Trends in Plant Science. 2007;12(10):444-451
  3. 3. Fraire-Velazquez S, Emmanuel V. Abiotic stress in plants and metabolic responses. In: Abiotic Stress - Plant Responses and Applications in Agriculture. London: InTech; 2013. DOI: 10.5772/54859
  4. 4. Huner NPA, Macdowall FDH. Effect of cold adaptation of puma rye on properties of RUDP carboxylase. Biochemical and Biophysical Research Communications. 1976;73(2):411-420
  5. 5. Nanjo Y, Skultety L, Uváčková LU, Klubicová K, Hajduch M, Komatsu S. Mass spectrometry-based analysis of proteomic changes in the root tips of flooded soybean seedlings. Journal of Proteome Research. 2012;11(1):372-385
  6. 6. Krasensky J, Jonak C. Drought, salt, and temperature stress-induced metabolic rearrangements and regulatory networks. Journal of Experimental Botany. 2012;63(4):1593-1608
  7. 7. Ahmad P, Jaleel CA, Salem MA, Nabi G, Sharma S. Roles of enzymatic and nonenzymatic antioxidants in plants during abiotic stress. Critical Reviews in Biotechnology. 2010;30(3):161-175
  8. 8. Júnior DF, Gaion LA, Júnior GS, Santos DMM, Carvalho RF. Drought-induced proline synthesis depends on root-to-shoot communication mediated by light perception. Acta Physiologiae Plantarum. 2018;40(1):15
  9. 9. Baena-González E, Sheen J. Convergent energy and stress signaling. Trends in Plant Science. 2008;13(9):474-482
  10. 10. Cattivelli L, Bartels D. Cold-induced mRNAs accumulate with different kinetics in barley coleoptiles. Planta. 1989;178(2):184-188
  11. 11. Feder ME. Integrative biology of stress: Molecular actors, the ecological theater, and the evolutionary play. In: International Symposium on Environmental Factors, Cellular Stress and Evolution, Varanasi, India. Vol. 2006. Storrs, CT, United States: Cell Stress Society International; 2006
  12. 12. Baena-González E. Energy signaling in the regulation of gene expression during stress. Molecular Plant. 2010;3(2):300-313
  13. 13. Bailey-Serres J, Voesenek LACJ. Flooding stress: Acclimations and genetic diversity. Annual Review of Plant Biology. 2008;59:313-339
  14. 14. Guimaraes E, Jueneman E. The Global Partnership Initiative for Plant Breeding Capacity Building (GIPB) (No. IAEA-CN--167). Rome, Italy: FAO; 2008
  15. 15. Levitt J. Responses of Plants to Environmental Stresses (No581.24/L666). New York: Academic Press; 1962
  16. 16. Cho YH, Hong JW, Kim EC, Yoo SD. Regulatory functions of SnRK1 in stress-responsive gene expression and in plant growth and development. Plant Physiology. 2012;158(4):1955-1964
  17. 17. Briggs DR, Siminovitch D. The chemistry of the living bark of the black locust tree relation to frost hardiness; seasonal variations in the electrophoresis patterns of the water-soluble proteins of the bark. Archives of Biochemistry. 1949;23(1):18
  18. 18. Foyer CH, Noctor G. Oxidant and antioxidant signalling in plants: A re-evaluation of the concept of oxidative stress in a physiological context. Plant, Cell & Environment. 2005;28(8):1056-1071
  19. 19. Akimoto-Tomiyama C, Tanabe S, Kajiwara H, Minami E, Ochiai H. Loss of chloroplast-localized protein phosphatase 2Cs in Arabidopsis thaliana leads to enhancement of plant immunity and resistance to Xanthomonas campestris pv. Campestris infection. Molecular Plant Pathology. 2018;19(5):1184-1195
  20. 20. Marmiroli N, Terzi V, Stanca MO, Lorenzoni C, Stanca AM. Protein synthesis during cold shock in barley tissues. Theoretical and Applied Genetics. 1986;73(2):190-196
  21. 21. Alcázar R, Marco F, Cuevas JC, Patron M, Ferrando A, Carrasco P, et al. Involvement of polyamines in plant response to abiotic stress. Biotechnology Letters. 2006;28(23):1867-1876
  22. 22. Amir R. Current understanding of the factors regulating methionine content in vegetative tissues of higher plants. Amino Acids. 2010;39(4):917-931
  23. 23. Gilmour SJ, Sebolt AM, Salazar MP, Everard JD, Thomashow MF. Overexpression of the Arabidopsis CBF3transcriptional activator mimics multiple biochemical changes associated with cold acclimation. Plant Physiology. 2000;124(4):1854-1865
  24. 24. Morimoto RI, Tissieres A, Georgopoulos C. Heat Shock Proteins: Structure, Function and Regulation. New York: Cold Spring Harbor Lab Press, Cold Spring Harbor; 1994
  25. 25. Aleksza D, Horváth GV, Sándor G, Szabados L. Proline accumulation is regulated by transcription factors associated with phosphate starvation. Plant Physiology. 2017;175(1):555-567
  26. 26. Gurley WB, Key JL. Transcriptional regulation of the heat-shock response: A plant perspective. Biochemistry. 1991;30(1):1-12
  27. 27. Kilian J, Whitehead D, Horak J, Wanke D, Weinl S, Batistic O, et al. The AtGenExpress global stress expression data set: Protocols, evaluation and model data analysis of UV-B light, drought and cold stress responses. The Plant Journal. 2007;50(2):347-363
  28. 28. Barros JA, Cavalcanti JHF, Medeiros DB, Nunes-Nesi A, Avin-Wittenberg T, Fernie AR, et al. Autophagy deficiency compromises alternative pathways of respiration following energy deprivation in Arabidopsis thaliana. Plant Physiology. 2017;175(1):62-76
  29. 29. Good AG, Zaplachinski ST. The effects of drought stress on free amino acid accumulation and protein synthesis in Brassica napus. Physiologia Plantarum. 1994;90(1):9-14
  30. 30. Grativol C, Hemerly AS, Ferreira PCG. Genetic and epigenetic regulation of stress responses in natural plant populations. Biochimica et Biophysica Acta (BBA)-Gene Regulatory Mechanisms. 2012;1819(2):176-185
  31. 31. Araújo WL, Tohge T, Ishizaki K, Leaver CJ, Fernie AR. Protein degradation–an alternative respiratory substrate for stressed plants. Trends in Plant Science. 2011;16(9):489-498
  32. 32. Krasnuk M, Jung GA, Witham FH. Electrophoretic studies of several dehydrogenases in relation to the cold tolerance of alfalfa. Cryobiology. 1976;13(3):375-393
  33. 33. Ferguson DL, Guikema JA, Paulsen GM. Ubiquitin pool modulation and protein degradation in wheat roots during high temperature stress. Plant Physiology. 1990;92(3):740-746
  34. 34. Ford KL, Cassin A, Bacic AF. Quantitative proteomic analysis of wheat cultivars with differing drought stress tolerance. Frontiers in Plant Science. 2011;2:44
  35. 35. Jin L, Huang B, Li H, Liu J. Expression profiles and transactivation analysis of a novel ethylene-responsive transcription factor gene GhERF5 from cotton. Progress in Natural Science. 2009;19(5):563-572
  36. 36. Martinelli T, Whittaker A, Bochicchio A, Vazzana C, Suzuki A, Masclaux-Daubresse C. Amino acid pattern and glutamate metabolism during dehydration stress in the ‘resurrection’plant Sporobolus stapfianus: A comparison between desiccation-sensitive and desiccation-tolerant leaves. Journal of Experimental Botany. 2007;58(11):3037-3046
  37. 37. Morimoto RI, Santoro MG. Stress–inducible responses and heat shock proteins: New pharmacologic targets for cytoprotection. Nature Biotechnology. 1998;16(9):833-838
  38. 38. Achard P, Gong F, Cheminant S, Alioua M, Hedden P, Genschik P. The cold-inducible CBF1 factor–dependent signaling pathway modulates the accumulation of the growth-repressing DELLA proteins via its effect on gibberellin metabolism. The Plant Cell. 2008;20(8):2117-2129
  39. 39. Chen YC, Holmes EC, Rajniak J, Kim JG, Tang S, Fischer CR, et al. N-hydroxy-pipecolic acid is a mobile metabolite that induces systemic disease resistance in Arabidopsis. Proceedings of the National Academy of Sciences. 2018;115(21):E4920-E4929
  40. 40. Kim HJ, Hwang NR, Lee KJ. Heat shock responses for understanding diseases of protein denaturation. Molecules & Cells (Springer Science & Business Media BV). 2007;23(2)
  41. 41. Hildebrandt TM. Synthesis versus degradation: Directions of amino acid metabolism during Arabidopsis abiotic stress response. Plant Molecular Biology. 2018;98(1-2):121-135
  42. 42. Pan Z, Zhao Y, Zheng Y, Liu J, Jiang X, Guo Y. A high-throughput method for screening Arabidopsis mutants with disordered abiotic stress-induced calcium signal. Journal of Genetics and Genomics. 2012;39(5):225-235
  43. 43. Khan MN, Sakata K, Hiraga S, Komatsu S. Quantitative proteomics reveals that peroxidases play key roles in post-flooding recovery in soybean roots. Journal of Proteome Research. 2014;13:5812-5828
  44. 44. Chen Y, Chen X, Wang H, Bao Y, Zhang W. Examination of the leaf proteome during flooding stress and the induction of programmed cell death in maize. Proteome Science. 2014;12:1
  45. 45. McCown BH, McLeester RC, Beck GE, Hall TC. Environment-induced changes in peroxidase zymograms in the stems of deciduous and evergreen plants. Cryobiology. 1969;5(6):410-412
  46. 46. Huner NP, Carter JV. Differential subunit aggregation of a purified protein from cold-hardened and unhardened puma rye. Zeitschrift für Pflanzenphysiologie. 1982;106(2):179-184
  47. 47. Jiang G, Wang Z, Shang H, Yang W, Hu Z, Phillips J, et al. Proteome analysis of leaves from the resurrection plant Boea hygrometrica in response to dehydration and rehydration. Planta. 2007;225(6):1405
  48. 48. Jin LG, Liu JY. Molecular cloning, expression profile and promoter analysis of a novel ethylene responsive transcription factor gene GhERF4 from cotton (Gossypium hirstum). Plant Physiology and Biochemistry. 2008;46(1):46-53
  49. 49. Miller GAD, Suzuki N, Ciftci-Yilmaz SULTAN, Mittler RON. Reactive oxygen species homeostasis and signalling during drought and salinity stresses. Plant, Cell & Environment. 2010;33(4):453-467
  50. 50. Mohammadi PP, Moieni A, Hiraga S, Komatsu S. Organ-specific proteomic analysis of drought-stressed soybean seedlings. Journal of Proteomics. 2012;75(6):1906-1923
  51. 51. Ma QH. Small GTP-binding proteins and their functions in plants. Journal of Plant Growth Regulation. 2007;26(4):369-388
  52. 52. Bogeat-Triboulot MB, Brosché M, Renaut J, Jouve L, Le Thiec D, Fayyaz P, et al. Gradual soil water depletion results in reversible changes of gene expression, protein profiles, ecophysiology, and growth performance in Populus euphratica, a poplar growing in arid regions. Plant Physiology. 2007;143(2):876-892
  53. 53. Hartmann M, Zeier T, Bernsdorff F, Reichel-Deland V, Kim D, Hohmann M, et al. Flavin monooxygenase-generated N-hydroxypipecolic acid is a critical element of plant systemic immunity. Cell. 2018;173(2):456-469
  54. 54. Hashiguchi A, Sakata K, Komatsu S. Proteome analysis of early-stage soybean seedlings under flooding stress. Journal of Proteome Research. 2009;8(4):2058-2069
  55. 55. Batista-Silva W, Heinemann B, Rugen N, Nunes-Nesi A, Araújo WL, Braun HP, et al. The role of amino acid metabolism during abiotic stress release. Plant Cell and Environment. 2019;42(5):1630-1644. DOI: 10.1111/pce.13518
  56. 56. Nsimba-Lubaki M, Peumans WJ. Seasonal fluctuations of lectins in barks of elderberry (Sambucus nigra) and black locust (Robinia pseudoacacia). Plant Physiology. 1986;80(3):747-751
  57. 57. Liu Q, Kasuga M, Sakuma Y, Abe H, Miura S, Yamaguchi-Shinozaki K, et al. Two transcription factors, DREB1 and DREB2, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought-and low-temperature-responsive gene expression, respectively, in Arabidopsis. The Plant Cell. 1998;10(8):1391-1406
  58. 58. Holcik M, Sonenberg N. Translational control in stress and apoptosis. Nature Reviews Molecular Cell Biology. 2005;6(4):318-327
  59. 59. Oh M, Komatsu S. Characterization of proteins in soybean roots under flooding and drought stresses. Journal of Proteomics. 2015;114:161-181
  60. 60. Farooq M, Barsa SMA. Wahid. A., Lee, DJ, Cheema, SA and Aziz. 2006;86:336-345
  61. 61. Aranjuelo I, Molero G, Erice G, Avice JC, Nogués S. Plant physiology and proteomics reveals the leaf response to drought in alfalfa (Medicago sativa L.). Journal of Experimental Botany. 2011;62(1):111-123
  62. 62. Komatsu S, Kuji R, Nanjo Y, Hiraga S, Furukawa K. Comprehensive analysis of endoplasmic reticulum-enriched fraction in root tips of soybean under flooding stress using proteomics techniques. Journal of Proteomics. 2012;77:531-560
  63. 63. Délano-Frier JP, Avilés-Arnaut H, Casarrubias-Castillo K, Casique-Arroyo G, Castrillón-Arbeláez PA, Herrera-Estrella L, et al. Transcriptomic analysis of grain amaranth (Amaranthus hypochondriacus) using 454 pyrosequencing: Comparison with a. tuberculatus, expression profiling in stems and in response to biotic and abiotic stress. BMC Genomics. 2011;12(1):363
  64. 64. Haque E, Kawaguchi K, Komatsu S. Analysis of proteins in aerenchymatous seminal roots of wheat grown in hypoxic soils under waterlogged conditions (supplementary material). Protein and Peptide Letters. 2011;18(9):912-924
  65. 65. Dobson CM, Šali A, Karplus M. Protein folding: A perspective from theory and experiment. Angewandte Chemie International Edition. 1998;37(7):868-893
  66. 66. Panaretou B, Zhai C. The heat shock proteins: Their roles as multi-component machines for protein folding. Fungal Biology Reviews. 2008;22(3-4):110-119
  67. 67. Amor Y, Haigler CH, Johnson S, Wainscott M, Delmer DP. A membrane-associated form of sucrose synthase and its potential role in synthesis of cellulose and callose in plants. Proceedings of the National Academy of Sciences. 1995;92(20):9353-9357
  68. 68. Meehl GA, Washington WM, Collins WD, Arblaster JM, Hu A, Buja LE, et al. How much more global warming and sea level rise? Science. 2005;307:1769-1772. DOI: 10.1126/science.1106663
  69. 69. Imamura H, Matsuyama Y, Miyagawa Y, Ishida K, Shimada R, Miyagawa S, et al. Prognostic significance of anatomical resection and des-γ-carboxy prothrombin in patients with hepatocellular carcinoma. British Journal of Surgery. 1999;86(8):1032-1038
  70. 70. Hirota T, Izumi M, Wada S, Makino A, Ishida H. Vacuolar protein degradation via autophagy provides substrates to amino acid catabolic pathways as an adaptive response to sugar starvation in Arabidopsis thaliana. Plant and Cell Physiology. 2018;59(7):1363-1376
  71. 71. Komatsu S, Kamal AHM, Makino T, Hossain Z. Ultraweak photon emission and proteomics analyses in soybean under abiotic stress. Biochimica et Biophysica Acta. 2014;1844:1208-1218
  72. 72. Nakamoto H, Vigh L. The small heat shock proteins and their clients. Cellular and Molecular Life Sciences. 2007;64(3):294-306
  73. 73. Nanjo Y, Skultety L, Ashraf Y, Komatsu S. Comparative proteomic analysis of early-stage soybean seedlings responses to flooding by using gel and gel-free techniques. Journal of Proteome Research. 2010;9(8):3989-4002
  74. 74. Gygi SP, Rochon Y, Franza BR, Aebersold R. Correlation between protein and mRNA abundance in yeast. Molecular and Cellular Biology. 1999;19(3):1720-1730
  75. 75. Roberts DWA. Changes in the proportions of two forms of invertase associated with the cold acclimation of wheat. Canadian Journal of Botany. 1979;57(4):413-419
  76. 76. Timperio AM, Egidi MG, Zolla L. Proteomics applied on plant abiotic stresses: Role of heat shock proteins (HSP). Journal of Proteomics. 2008;71(4):391-411
  77. 77. Solaw FAO. The State of the world’s Land and Water Resources for Food and Agriculture. Rome, Italy: FAO; 2011
  78. 78. Dhaubhadel S, Browning KS, Gallie DR, Krishna P. Brassinosteroid functions to protect the translational machinery and heat-shock protein synthesis following thermal stress. The Plant Journal. 2002;29(6):681-691
  79. 79. Burke TJ, Callis J, Vierstra RD. Characterization of a polyubiquitin gene from Arabidopsis thaliana. Molecular and General Genetics MGG. 1988;213(2-3):435-443
  80. 80. Sheikh AH, Eschen-Lippold L, Pecher P, Hoehenwarter W, Sinha AK, Scheel D, et al. Regulation of WRKY46 transcription factor function by mitogen-activated protein kinases in Arabidopsis thaliana. Frontiers in Plant Science. 2016;7:61
  81. 81. Guy CL, Carter JV. Characterization of partially purified glutathione reductase from cold-hardened and nonhardened spinach leaf tissue. Cryobiology. 1984;21(4):454-464
  82. 82. Reddy AS, Ali GS, Celesnik H, Day IS. Coping with stresses: Roles of calcium-and calcium/calmodulin-regulated gene expression. The Plant Cell. 2011;23(6):2010-2032
  83. 83. Riov J, Brown GN. Comparative studies of activity and properties of ferredoxin–NADP+ reductase during cold hardening of wheat. Canadian Journal of Botany. 1976;54(16):1896-1902
  84. 84. Kosová K, Vítámvás P, Prášil IT, Renaut J. Plant proteome changes under abiotic stress - contribution of proteomics studies to understanding plant stress response. Journal of Proteomics. 2011;74(8):1301-1322. DOI: 10.1016/j.jprot.2011.02.006
  85. 85. Gupta SC, Sharma A, Mishra M, Mishra RK, Chowdhuri DK. Heat shock proteins in toxicology: How close and how far? Life Sciences. 2010;86(11-12):377-384
  86. 86. Su PH, Li HM. Arabidopsis stromal 70-kD heat shock proteins are essential for plant development and important for thermotolerance of germinating seeds. Plant Physiology. 2008;146(3):1231-1241
  87. 87. Huang T, Jander G. Abscisic acid-regulated protein degradation causes osmotic stress-induced accumulation of branched-chain amino acids in Arabidopsis thaliana. Planta. 2017;246(4):737-747
  88. 88. Showler AT. Effects of water deficit stress, shade, weed competition, and kaolin particle film on selected foliar free amino acid accumulations in cotton, Gossypium hirsutum (L.). Journal of Chemical Ecology. 2002;28(3):631-651
  89. 89. Parsell DA, Lindquist S. The function of heat-shock proteins in stress tolerance: Degradation and reactivation of damaged proteins. Annual Review of Genetics. 1993;27(1):437-496
  90. 90. Tzin V, Galili G. New insights into the shikimate and aromatic amino acids biosynthesis pathways in plants. Molecular Plant. 2010;3(6):956-972
  91. 91. Vickers NJ. Animal communication: When I’m calling you, will you answer too? Current Biology. 2017;27(14):R713-R715
  92. 92. Lee U, Rioflorido I, Hong SW, Larkindale J, Waters ER, Vierling E. The Arabidopsis ClpB/Hsp100 family of proteins: Chaperones for stress and chloroplast development. The Plant Journal. 2006;49(1):115-127
  93. 93. Dickens BF, Thompson GA Jr. Rapid membrane response during low-temperature acclimation correlation of early changes in the physical properties and lipid composition of tetrahymena microsomal membranes. Biochimica et Biophysica Acta (BBA)-Biomembranes. 1981;644(2):211-218
  94. 94. Thomashow MF. Molecular basis of plant cold acclimation: Insights gained from studying the CBF cold response pathway. Plant Physiology. 2010;154(2):571-577
  95. 95. Salekdeh GH, Siopongco J, Wade LJ, Ghareyazie B, Bennett J. Proteomic analysis of rice leaves during drought stress and recovery. PROTEOMICS: International Edition. 2002;2(9):1131-1145
  96. 96. Thomashow MF. Plant cold acclimation: Freezing tolerance genes and regulatory mechanisms. Annual Review of Plant Biology. 1999;50(1):571-599
  97. 97. Schmidhuber J, Tubiello FN. Global food security under climate change. Proceedings of the National Academy of Sciences. 2007;104(50):19703-19708
  98. 98. Maruyama K, Takeda M, Kidokoro S, Yamada K, Sakuma Y, Urano K, et al. Metabolic pathways involved in cold acclimation identified by integrated analysis of metabolites and transcripts regulated by DREB1A and DREB2A. Plant Physiology. 2009;150(4):1972-1980
  99. 99. Morimoto RI. Cells in stress: Transcriptional activation of heat shock genes. Science-New York Then Washington. 1993;259:1409-1409
  100. 100. Vierling E. The roles of heat shock proteins in plants. Annual Review of Plant Biology. 1991;42(1):579-620
  101. 101. Liu J, Feng L, Li J, He Z. Genetic and epigenetic control of plant heat responses. Frontiers in Plant Science. 2015;6:267
  102. 102. Song L, Li R, Xiang X, Wang J, Qiao L, Song X, et al. Overexpression of stress-inducible small GTP-binding protein AhRab7 (AhRabG3f) in peanut (Arachis hypogaea L.) enhances abiotic stress tolerance. Journal of Food, Agriculture and Environment. 2012;10:888-894
  103. 103. Pires MV, Pereira Júnior AA, Medeiros DB, Daloso DM, Pham PA, Barros KA, et al. The influence of alternative pathways of respiration that utilize branched-chain amino acids following water shortage in Arabidopsis. Plant, Cell & Environment. 2016;39(6):1304-1319
  104. 104. Bohler S, Sergeant K, Jolivet Y, Hoffmann L, Hausman JF, Dizengremel P, et al. A physiological and proteomic study of poplar leaves during ozone exposure combined with mild drought. Proteomics. 2013;13(10-11):1737-1754
  105. 105. Key JL, Lin CY, Chen YM. Heat shock proteins of higher plants. Proceedings of the National Academy of Sciences. 1981;78(6):3526-3530
  106. 106. Krasnuk M, Jung GA, Witham FH. Electrophoretic studies of the relationship of peroxidases, polyphenol oxidase, and indoleacetic acid oxidase to cold tolerance of alfalfa. Cryobiology. 1975;12(1):62-80

Written By

Bharti Thapa and Abhisek Shrestha

Reviewed: 01 February 2022 Published: 15 May 2022