Open access peer-reviewed chapter

Molecular Defense Mechanisms in Plants to Tolerate Toxic Action of Heavy Metal Environmental Pollution

Written By

Istvan Jablonkai

Submitted: 19 July 2021 Reviewed: 21 December 2021 Published: 21 March 2022

DOI: 10.5772/intechopen.102330

From the Edited Volume

Plant Defense Mechanisms

Edited by Josphert Ngui Kimatu

Chapter metrics overview

206 Chapter Downloads

View Full Metrics

Abstract

Toxic action of heavy metals on plants growing in contaminated soils intensified the research on detoxification and sequestering mechanisms existing in plants to understand and manipulate defense mechanisms that confer tolerance against metal ions. Increased biosynthesis of plant biomolecules to confer tolerance during toxic action of heavy metals is an intrinsic ability of plants. Induced formation of low-molecular weight amino acids, peptides or proteines as chelators such as proline (Pro), glutathione (GSH), phytochelatins (PCs) or metallothioneins (MTs) under heavy metal stress enhances metal binding and detoxification capability of plants. In addition, proline and GSH related enzymes such as GSH reductase, GSH peroxidases and glutathione S-transferases are also key components of the antioxidant defense system in the cells to scavenge reactive oxygen species (ROS). Protective action of oxidized fatty acids oxylipins at toxic levels of heavy metals is considered to activate detoxification processes as signaling molecules.

Keywords

  • heavy metals
  • stress
  • detoxification
  • glutathione
  • chelators

1. Introduction

Abiotic stress factors such as extreme temperatures, drought, salinity, heavy metals, xenobiotics have been considered as potential threats for agricultural productivity. To cope with abiotic stress, plants can initiate a number of molecular, cellular, and physiological changes to respond and adapt to such stresses. Stress-tolerance of plants involves the activation of cascades of molecular networks leading to expression of stress-related enzymes.

The accumulation of heavy metals in the plant is accompanied by damage to the structural components and an imbalance of various metabolic processes in the cells, which leads to disruption of plant growth and development. Plants have evolved a number of mechanisms to adapt to increasing concentrations of metal ions. These include the immobilization, exclusion, chelation, and compartmentalisation of metal ions.

The toxic action of heavy metals can produce excessive reactive oxygen species (ROS). Glutathione and GSH related enzymes such as GSH reductase, GSH peroxidases and glutathione S-transferases are fundamental parts of the antioxidant defense system in the cells to scavenge ROS and electrophilic organic molecules as well. The significance of plant thiols and glutathione S-transferases involved in plant response to almost all stress factors will be discussed with special attention to their overexpression, redox status and regulation that confer stress tolerance.

A number of metal-binding ligands have now been recognized in plants [1]. Polypeptide ligands include the metallothioneins (MTs), small, gene-encoded, cysteine-rich polypeptides, and the phytochelatins (PCs), which, in contrast, are enzymatically synthesized are effective metal binding ligands have been identified. Recent advances in understanding the role, biosynthesis and protective action of phytochelatins and metallothioneins as metal-binding ligands in heavy metal detoxification are reviewed.

Proteinogenic amino acid proline protects cell from environmental stress factors by several protective mechanisms such as osmoprotectant, acting as chemical chaperone by stabilization of proteins and antioxidant enzymes, chelation of metals, scavenging ROS, balancing the intracellular redox homeostasis (NADP+/NADPH ratio, GSH pool) and participating in cellular metabolic signaling. Proline accumulation has been observed in response to environmental stress in a variety of living organism including plants.

Reactive electrophile species oxylipins exhibit protective action under toxic concentration of pollutants by activation of detoxification processes. The less studied oxylipin signaling in plant stress response will be detailed as an important factor in plant adaptation to stress by heavy metal pollutants.

Advertisement

2. Role of glutathione (GSH) and glutathione related enzymes in protection of plants against toxic action of heavy metals

The accumulation of heavy metals in plant organs results in the damage of the structural components and disruption of cell metabolic processes leading to plant growth retardation. The toxic action of heavy metals can produce excessive reactive oxygen species (ROS). Glutathione and GSH related enzymes such as GSH reductase, GSH peroxidases and glutathione S-transferases are fundamental parts of the antioxidant defense system in the cells to scavenge ROS and electrophilic organic molecules as well (Figure 1).

Figure 1.

Biosynthesis of GSH and GSH/GSSG redox system with related enzymes such as glutathione peroxidase (GPX) and glutathione reductase (GR) as well as glutathione S-transferase.

2.1 Glutathione (GSH) and GSH/GSSG redox system

The most abundant biological thiol the tripeptide glutathione (GSH, γ-Glu-Cys-Gly) is a low molecular weight, water soluble compound that is ubiquitous in most plant tissues and has a key function in stress management. Besides the ROS detoxification, GSH also participates in detoxication of methylglyoxal and electrophilic xenobiotics [2]. As a component of the glutathione-ascorbate (AsA-GSH) cycle GSH has a key role in H2O2 detoxication. GSH is used as co-factor by a) various peroxidases in detoxication of peroxides formed in the reaction of oxygen radical with biological molecules and b) by glutathione S-transferases (GST) to conjugate GSH with endogenous substances and xenobiotics. Interaction of GSH with thioredoxin systems fine-tune photosynthetic and respiratory metabolism by modifying the sensitive protein Cys residues [3]. GSH is a substrate of phytochelatin synthase and oligomeric GSH products phytochelatins (PCs) can effectively sequester heavy metals by complexation alleviating metal stress of plants [4]. Interaction of GSH with known signaling molecules such as salicilic acid, jasmonic acid and ethylene can be important in treatment of plant biotic stress [5].

GSH biosynthesis is two-step pathway mediated by γ -glutamylcysteine synthase (γ-ECS) and glutathione synthase (GSHS) enzymes using 2 molecules of ATP (Figure 1). The first step occurs mainly in the chloroplast while the second step predominantly takes place in the cytosol [3, 6]. GSH produced in the cytosol can be transported directly to other cellular organelles by glutathione transporters [7].

The GSH redox state in plants, particularly in leaves is normally very stable but is extremely sensitive to oxidative stress. In the absence of stress, in leaf tissues measurable GSH: GSSG ratios typically around 20:1 [8]. Conversion of the reduced GSH into oxidized glutathione or glutathione disulfide (GSSG) can occur under stress conditions. Non-enzymatic reactions of GSH with ROS species such as 1O2, O2, OH, H2O2 and O22− convert the reduced GSH to the oxidized form (GSSG, glutathione disulfide) [9, 10]. Glutathione reductase (GR) and glutathione peroxidase (GPX) enzymes in conjuction with AsA-GSH cycle are responsible for the balanced state of GSH/GSSG and GSH homeostasis [11].

Plethora of information on alteration of plant GSH pool and GSH/GSSG redox system as a results of heavy metal stress are available. Elevated levels of GSH have been observed in various plant species with increasing Cd concentration. But depletion of GSH in roots of variety of species under cadmium and lead stress has also been reported [12]. In two genotypes of Brassica juncea treated with Cd increased levels of GSH and GSSG were detected. Higher increase in GSH content was found in Cd-tolerant genotype while in Cd-sensitive genotype the enhancement of level of the oxidized form, GSSH was more pronounced [13]. Remarkable decreases in O2, H2O2, and MDA (malondialdehyde) accumulation were detected with exogenously applied GSH in rice seedlings treated with Cd as a result of modulation of the antioxidant system [14]. The oxidative stress induced by Cd in coontail, a free-floating freshwater macrophyte (Ceratophyllum demersum) was alleviated by exogenous Zn application. The application of Zn restored thiols and also inhibited oxidation of AsA and GSH, and sustained the redox state balance. Application of zinc enhanced the activities of AsA-GSH enzymes (APX, MDHAR, DHAR, and GR), GST, and GPX, conferring tolerance to Cd stress [15].

Toxic action of heavy metals seemed to induce the expression of genes encoding γ -glutamylcysteine (γ-ECS) and glutathione synthase (GSHS) enzymes by enhancing GSH levels. Moreover, GSH can efficiently influence coordination of metals to the active sites of affected enzymes preserving their activity [16].

2.2 Glutathione reductase

Glutathione reductase (GR) is a flavo-protein oxidoreductase mediates the reduction of GSSG to GSH using NADPH as an electron donor. GR is predominantly located in chloroplasts but some isoforms were also detected in mitochondria and cytosol [17]. GR activity secure the redox potential of cells at highly reduced GSH/GSSG and AsA/DHA ratios under regular and oxidative stress conditions. Biotic and abiotic stress factors including toxic metals were found to influence GR activities in plants [17, 18]. Under Cd stress GR activities were increased in sugarcane callus cultures in time- and concentration-dependent manner [19]. Elevated GR activities were detected in response to CD-treatment in a variety of plant species such as Solanum tuberosum, Raphanus sativus, Glycine max, Saccharum officinarum, Capsicum annuum, Arabidopsis thaliana, B. juncea, Brassica campestris, Vigna mung, and Pisum sativum [12]. On the contrary significant reduction of GR activity was shown in Ceratophylum demersum treated with Cd [20]. A seedling age specific changes in GR activities were observed in Oryza sativa genotypes under Pb stress. After initial deacreases in GR activities in both roots and shoots up to 10 days a remarkable increase of enzyme activities was found after 15 days of treatment [21].

2.3 Glutathione peroxidases

Glutathione peroxidases (GPXs) are a large family of broad substrate spectrum multiple isozymes. Contrary to most of their counterparts in animal cells, plant GPXs contain cysteine instead of selenocysteine in their active site. Antioxidant GPX protecting cells from oxidative damage by reducing H2O2, organic and lipid hydroperoxides in association with the GSH pool [22]. Presence of GPXs were detected in cytosol, chloroplasts, mitochondria, peroxisome and apoplast.

Stress responses of GPXs are contradictory. Cd stress increased GPX activity in cultivars of C. annuum plants [23] but reduced activities were found in roots of Cd-exposed P. sativum plants [24] while no change was observed in GPX activities in maize roots exposed to Ni [25]. The activity of GPX activity was significantly enhanced in Lolium perenne shoots exposed to Se but chronic metal exposure decreased enzyme activity [26]. Externally supplied GSH to Cd treated barley genotypes was shown to counteract Cd-induced elevation of root GPX activity by reducing GPX activity to the control level [14].

2.4 Detoxification action of glutathione S-transferases

Glutathione S-transferases (glutathione sulfo-transferases, GSTs) are major phase II, GSH-dependent detoxication enzyme superfamily. GSTs catalyze the conjugation of glutathione (GSH) to a wide variety of endogenous and exogenous electrophilic compounds to form water soluble, non-toxic GSH conjugates [27, 28]. GSTs are divided into two distinct super-family members: the membrane-bound microsomal and cytosolic family members. Microsomal GSTs are structurally distinct from the cytosolic in that they homo- and heterotrimerize rather than dimerize to form a single active site [29]. In cytoplasm GSTs account for roughly 1% of soluble proteins in maize leaves [30]. Various plant GST izozymes were shown to possess with GSTs/GPX activity mediating lipid hydroperoxide metabolism to non-toxic alcohols [31].

Elevated GST activities were found in leaves and roots of Cd exposed pea plants [24] and in roots of Phragmites australis plants by Iannelli et al. [32]. Cadmium- and mercury-induced root growth inhibition is strongly correlated with increased GST and GPX activity in barley [33] while in maize seedlings, Cd treatment strongly induced (20-50-fold increase) GST Bronze2 and GST III [34]. A detailed summary of literature studies on GST induction by heavy metals such as Cd, Pb, Cu, As(III) and As(V) has been published [12].

Advertisement

3. Phytochelatins (PCs) as heavy metal chelators

One of the detoxication mechanisms in plants to overcome heavy metal stress is the production of thiol-containing oligomer peptides from a precursor glutathione (GSH) by phytochelatin synthase (PCS, γ-glutamylcysteine dipeptidyltranspeptidase) (Figure 2).

Figure 2.

General structure of (a) and biosynthesis (b) phytochelatins.

PC synthase-deficient mutants of Arabidopsis and S. pombe exhibited high sensitivity towards Cd and arsenite providing strong evidence for the role of PCs in heavy metal detoxification [35]. Phytochelatin level of plant can serve as biomarkers for the initial detection of heavy metal stress. Since the immobilized metals are less toxic than the free ions, PCs are considered as part of the detoxication mechanism of higher plants, [36, 37]. Phytochelatins, with (γ-Glu-Cys)n-Gly general formula (where n = 2-11), can act as chelators to bind heavy metals in the cytosol and the metal-phytochelatin complexes are compartmentalized to vacuoles. Varieties in structures of PCs include the replacement of glycine residue (Gly) with β-alanine (β-Ala), Ala, glutamine (Gln), serine (Ser) or glutamate (Glu) [38].

As a result of Cys residues PCs have high thiol (SH) contents and ability to strongly bind heavy metals exhibiting increased metal-binding capacity with increasing size [39]. PCs chain lengths varies with plant species and metal forms. The relative affinity of metals such as Cd, Pb, Hg, Cu to GSH and PCn oligomers increases with chain length (GSH < PC2 < PC3 < PC4) [40]. After the formation, the high molecular weight metal-phytochelatin complexes are sequestered in the vacuoles by the involvement of ABC-type (ATP-binding casette) transporters [41]. PC synthase was shown to be activated by heavy metal ions such as Cd2+, Cu2+, Ag+, Hg2+, Pb2+, Ni2+. Cd tolerance of Arabidopsis and tobacco was found to be mainly related to PC content. Contents of PCs and GSH in Arabidopsis were 3.5 and 3 times higher than in tobacco plants and the concentration of various PCs oligomers in the two species was different: PC3 and PC4 oligomers were prevalent in wild-type tobacco as compared to high concentration of PC2 and PC3 in Arabidopsis [42].

Cd2+ ions were found as the most effective stimulator of PCs biosynthesis and 4-6-fold higher induction of PCs were detected than with Cu2+ and Zn2+ compounds in cell cultures of Indian snakeroot (Rauvolfia serpentina) [43] and red spruce (Picea rubens Sarg) [44], respectively.

Biosynthesis of PCs takes place in the cytosol of root cells. PCs are produced from glutathione, homoglutathione, hydroxymethyl-gluthathione or glutamylcysteinyl-glutamate by a transpeptidase, the constitutive PC synthase enzyme [45]. In Solanum nigrum L. copper treatment enhanced the biosynthesis of PCs was shown followed by the immobilization of toxic Cu in the roots by inhibiting the translocation into the shoots [46]. Reduced transport of As from roots to shoots was found in rice cultivars subjected to the elevated levels of arsenic due to the stimulation of formation of phytochelatin-arsenic complexes [47]. Arsenite and arsenate anions are readily absorbed by plants and both anions were found to induce effectively the biosynthesis of PCs in vivo and in vitro [48]. A higher degree of production, accumulation, and transportation of PCs has a definite role in tolerance of plants to heavy metal stress. However, biosynthesis of PCs does not neccessarily take place in roots. In hyperaccumulator Sedum alfredii plants PC synthesis and Cd accumulation were most abundant in the leaves followed by the stems but hardly detected in the roots [49]. In a Cd hyperaccumulator wheat plant, Cd accumulation was not only the sequestration of Cd-phytochelatin complexes in the roots but their translocation to shoots also takes place. The translocation of Cd from root to shoot is through the xylem appears to be the main process for shoot accumulation. At a relatively high Cd treatment (20 μM) phytochelatin biosynthesis was enhanced more evidently in shoots [50]. The phytochelatin independent mechanism of tolerance of higher plants to Cd toxicity also exists and can be attributed to the highly developed apoplastic transport systems [51].

PCs enzyme activities were not only also induced Cd stress. The principal mechanism of intracellular metal detoxification by complexing and transporting metals into the vacuole was also established for stress by various heavy metals such as mercury, copper, arsenic, silver, nickel, gold, and zinc [52, 53, 54].

Stochiometry of complexes of Pb2+ formed with various PCn (n = 2–4) were examined in details. Mass spectrometry analysis of Pb– PC2 revealed four different complexes corresponding to [Pb–PC2]+, [Pb2–PC2]+, [Pb–(PC2)2]+, and [Pb2– (PC2)2]+. The coordination of Pb2+ with PC2 was postulated via the thiol groups of cysteine residue of PC2 and possibly by carboxylic groups. In case of PC3 and PC4, two complexes were detected for each metal such as Pb–PC3, Pb2–PC3, Pb–PC4, and Pb2–PC4 [55]. The metal-PCn complexes formed with Cd2+ differed from those of Pb2+. Higher metal/peptide molar ratios were estimated in Cd-PCn complexes than in Pb-PCn complexes (n = 3, 4), suggesting that phytochelatins of marine algae Phaeodactylum tricornutum are capable to sequester and detoxify more Cd2+ than Pb2+ ions forming complexes with a different structure and stoichiometry. As a confirmation 80% of Cd2+ was detected in a complex form, and only 40% of the absorbed Pb2+ was bound to phytochelatins in the cellular extract [56].

Advertisement

4. Metallothioneins (MTs), the metal-chelating proteins

Involvement of MTs has been reported in a number of physiological processes such as regulation of cell growth and proliferation, toxic metal protection and homeostasis, free radical scavenging or protection from oxidative stress, DNA damage repair [57]. For a long MTs was considered to be expressed only in mammals while in plants enzymatically formed PCs are the protective biomolecules against heavy metal toxicity. MTs is a cytosolic superfamily of cysteine-rich proteins capable to bind both physiological and xenobiotic heavy metals [58, 59]. While phytochelatins are formed enzymatically MTs are the products of mRNA translation [60, 61]. In the structure of MTs cysteine (C) residues representing about 30% of the constitutional amino acids. Primary structures of this low molecular weight protein family (Mw ranging from 5 to10 kDa) are characteristically rich in highly conserved CC, CXC (where X is a general amino acid) and CXXC motifs that render a unique ability to bind mono- or divalent metal ions, such as Cu2+, Zn2+and Cd2+ [62]. While in animal MTs no aromatic acids occur histidine (His) residues can be found in a number of plant MT sequences. The replacement a part of the Cys residues by His would be an increased selectivity for Zn2+ over Cd2+, and thus the function of the respective MT is more selective in maintaining Zn2+ homeostasis than heavy metal detoxification [63]. In addition, the thiol(ate)s in MTs can act as powerful antioxidants, and therefore MTs have definite roles in protection against oxidative stress [64].

Based on the arrangement of Cys residues four types of plant MTs exist [59]. Type 1 MTs are mainly expressed in roots, while the expression of type 2 MTs mostly occurs in shoots, type 3 MTs are induced in leaves and during fruit ripening, and type 4 MTs are abundant in the developing seeds [65]. Regarding high level of sequence diversity of plant metallothioneins, type 1-4 MTs is further subdivided to various isoforms. All four types of plant MTs and their isoforms are able to chelate heavy metals. In general, the primary structures of plant MTs in type 1, 2 and 3 have a similar cysteine topology. The two Cys-rich domains (α and β) are attached by a 30-40 amino acid long cysteine-poor linker depending on plant species. Cysteine topology of type 4 MTs different from that observed in MTs of type 1-3. In angiosperm species three Cys-rich regions linked by two Cys-poor linkers containing 15 and 40 amino acids [58, 60]. Experiments with Arabidopsis thaliana MTs expressed in copper and zinc sensitive yeast mutants provided evidences that MT1a, 2a, 2b and MT3 function as copper binding MTs. The seed-specific type-4 MTs were more effective than other Arabidopsis MTs in providing protection against Zn toxicity and enhancing Zn accumulation [66].

Studies on copper tolerance and expression MT1 and MT2 genes in several A. thaliana species revealed that MT1 was uniformly expressed in all ecotypes and MT2 was copper inducible. In cross-induction experiments, Ag+, Cd2+, Zn2+, Ni2+ significantly enhanced the levels of MT2 genes [67]. A. thaliana plants knocked down for MT1a and b isoform expression exhibited increased cadmium sensitivity. These lines accumulated less As, Cd and Zn in the leaves than wild-types indicating that MT1 have a definitive role in Cd tolerance and possibly involved in Zn homeostasis [68]. Lack of MTs increased Cd and Cu sensitivity in PC-deficient Arabidopsis plants suggesting that PCs and MTs contribute to Cu and Cd tolerance and may overlap in their functions [66]. Experiments with A. thaliana MTs expressed in copper and zinc sensitive yeast mutants provided evidences that MT1a, 2a, 2b and MT3 function as copper binding MTs. The seed-specific type 4 MTs were more effective than other Arabidopsis MTs in confering protection against Zn toxicity and enhancing Zn accumulation [66].

MTs typically bind metal ions in characteristic metal– thiolate clusters that provide high thermodynamic stability coupled with kinetic lability [69]. The large diversity in the metal binding regions of plant MTs confers the ability to bind a greater range of metals than in animals possessing a greater range of function [59]. A model of cadmium binding to mammalian MTs showed that all cysteine residues participate in the coordination of 7 mol of Cd per mol of MT. Two polynuclear metal clusters formed during binding with 3 and 4 metal atoms on β- and α-domain, respectively.

In one metal cluster requires 9 cysteine SH group to bind 3 Cd in a six-membered ring while the four-metal-cluster forms a bicyclo[3.1.3] structure with the participation of 11 Cys-SH groups (Figure 3) [70].

Figure 3.

Proposed structures of the four-metal and three-metal clusters of rat liver metallothioneins based on 113Cd NMR data [71]. Adapted from Klaassen et al. [70].

Experimentally and predicted stochiometries of metal-plant MTs are in agreement. Type 1 plant MTs with 12 Cys residues can bind 4-5 metal ions, type 2 with 14 Cys can coordinate 5 metal ions while type 3 MTs with only 10 Cys residues exhibit the lowest capacity for metal binding (4 metal ions) [58].

The key characteristics of metal-MT complexes are the high thermodynamic but low kinetic stability. Thermodynamically, MTs are the most stable zinc sites in eukaryotes but have appropriate kinetic lability for the protein to intermolecularly exchange zinc with proteins [72]. Metal binding affinities of MTs characterized by complex stability constants are hardly available in the literature. However, pH of half-displacement values (pH (1/2)) are available for several plant MTs indicating that the more stable protein-metal complex have lower pH(1/2) value [63]. pH(1/2) values for all MTs follow the order Cu(I) < Cd(II) < Zn(II). Accordingly, MTs will bind Cu(I) in vitro more strongly than Cd(II) or Zn(II), and Cd(II) will be bound more powerfully than Zn(II). Numerically The pH values at which MT1 of pea loses half of the initially bound metals are 5.6 for Zn(II), 4.0 for Cd(II), and 1.5 for Cu(I) [73].

Further studies on large and complex MT gene families in higher plants may exhibit beneficial metal binding and induction properties to enhance the phytoremediation capacity of plants used for heavy metal removal in soils. To understand the function and the mechanism of action of plant MTs requires further manipulations on the expression of this protein family.

Advertisement

5. Protective action of proline (pro) against heavy metals toxicity

Proline (Pro) is an essential proteinogenic amino acid that fulfill several developmental functions in plants and has a fundamental role in responses to biotic and abiotic stress. In plants proline can protect cells from environmental stress factors by several protective mechanisms such as acting as osmoprotectant, functioning as chemical chaperone stabilizing proteins and antioxidant enzymes, chelating metals, scavenging reactive oxygen species (ROS), balancing the intracellular redox homeostasis (NADP+/NADPH ratio, GSH pool) and participating in cellular metabolic signaling [74, 75, 76].

Protective mechanism of proline as ROS scavanger include direct reaction with ROS. Free and polypeptide-bound proline was demonstrated to react with hydrogen peroxide (H2O2) and hydroxy radicals (OH.) producing stable free radical adducts and hydroxyproline (Figure 4). However, the most important ROS-scavenging mechanism in the stress protection is the reaction with singlet oxygen (1O2). A direct reaction between H2O2 and proline plays a minor role in scavenging of cellular H2O2 and the formation of nitroxyl radical accumulate during is very sluggish as compared to that of proline and OH.. Nevertheless, proline effectively quenches 1O2 via a charge transfer mechanism to form the ground triplet oxygen (3O2) as a ground state molecular oxygen. Due to its action as a 1O2 quencher, proline may help stabilize proteins, DNA, and membranes [77].

Figure 4.

Routes for scavenging reactive oxygen species (ROS) by proline. Adapted from Liang et al., 2013 [74].

Toxic action of lead, copper, and zinc on proline, malondialdehyde (MDA), and superoxide dismutase (SOD) has been studied in the cyanobacterium Spirulina platensis-S5. Parallel to growth reduction elevated MDA, SOD, and proline contents were found with increasing concentrations of metals. Elevated levels of MDA indicated the formation of free radicals in generated heavy metals stress while enhanced amounts of SOD and proline demonstrated the undergoing scavenging mechanism [78].

The molecular mechanism of proline protection of cells during stress may involve the effects on redox systems such as the glutathione (GSH) pool. Pro reduces heavy metal stress by detoxification of free radicals produced as a result of Cd poisoning. Pro may physically quench oxygen singlets or react directly with hydroxyl radicals. These reactions result in reduced free radical damage (lower malondialdehyde levels) and a more reducing cellular environment (higher GSH levels). The high GSH levels in turn facilitate phytochelatin synthesis and sequestration of heavy metal phytochelatin conjugates in the vacuole. This enhanced sequestration of Cd-phytochelatin complexes in the vacuole accounts for the transiently increased Cd content of P5CS (Δ1-pyrroline-5-carboxylate synthetase)-expressing cells in transgenic algae (Figure 5) [76].

Figure 5.

Participation of proline in reducing cadmium stress by detoxification of free radicals generated by Cd toxicity. In wild type algae (Figure 5A), cd2+ induces the production of reactive oxygen species that rapidly oxidize GSH to GSSG. In transgenic algae (Figure 5B) the resulting decreased GSH are reduced with free pro leading to increased GSH levels and therefore enhanced phytochelatin synthesis to coordinate and sequester Cd. Adapted from Siripornadulsil et al., 2002 [76].

The chelation of metals is also considered as possible mechanism responsible for the protective effect of proline in cells against stress. Reversal on Cd- and Zn-induced inhibition of glucose-6-phosphate dehydrogenase and nitrate reductase enzymes by proline supported the function of proline as a metal chelator by forming proline-metal complexes [79]. Nevertheless, the stability of metal-proline complexes was found to be relatively low [80]. to effectively influence the inhibitory concentration of the metal ions in the assay mixtures. A copper–proline complex was found in the roots of copper-tolerant sea thrift (Armeria maritima) [75].

Accumulation of free proline plants in response to abiotic stresses has been demonstrated by Delauney and Verma [81]. In Arabidopsis, accumulation of Pro occurs after NaCl stress and can reach 20% of the total free amino acid pool [82]. The metal-tolerant plant species such as Armeria maritima, Deschampsia cespitosa, and Silene vulgaris were reported to contain substantially higher levels of constitutive proline than non-tolerant plants [46]. In hyperaccumulator artichoke (Cynara scolymus L.), Pro elevation in cells correlated with increasing lead concentration [83]. Enhancement of Pro concentration in root of rape seed (Brassica napus L.) was more significant than in shoot tissues as a result of increasing concentrations of Pb2+ [84]. Root specific accumulation of the proline was detected in Indian mustard (Brassica juncea L.) under lead and cadmium stress [85]. Similarly elevated prolin levels were detected in roots of a variety of plants such as black nightshade (Solanum nigrum L.) exposed to copper [4], in wheat stressed by cadmium [86], and mercury and cadmium stressed lemongrass (Cymbopogon flexuosus Stapf) [87]. In hybrid poplar (Populus trichocarpa × deltoides) Pro accumulation in roots was almost double than in leaf tissues due to highly toxic Cd2+ exposure [88]. In the roots of pepper plants (Capsicum annuum L.) elevated proline level was detected with increasing Cr concentration, while Pro decreased in leaves [89]. On the contrary, shoot specific accumulation of Pro was detected in sunflower (Helianthus annuus L.) seedlings as a result of copper stress [90]. In metal non-tolerant bladder campion (S. vulgaris) Cu was most effective inducer of the proline accumulation followed by Cd and Zn, respectively [46]. Treatment of cauliflower seedlings (Brassica oleracea var. botrytis) with Cd, Hg, and Zn resulted in the concentration dependent accumulation of Pro up to double. Among these heavy metals Hg was the most effective inducer of Pro biosynthesis [91]. Treatment of sal seedlings (Shorea robusta) with heavy metals increased the proline accumulation in plants in the order of Cd2+ > Pb2+ > As3+ [92]. In wheat seedlings stimulation of Pro levels was more pronounced by copper than zinc [93]. Additionally, induction of Pro contents in shoots of black gram (Vigna mungo L.) seedlings treated with heavy metals changed in the order of highest to lowest as Hg > Pb > Co > Zn and followed the toxicities of heavy metals. The results suggest that heavy metal stress induced proline accumulation is strongly dependent on the concentration and type of heavy metal and plant species [94].

It seems that two mechanisms are responsible by which proline provides protection against heavy metal stressors. One of these, up-regulation of proline biosynthesis to accumulate proline to serve as an osmolyte, a chemical chaperone, and a direct scavenger of hydroxyl radical and singlet oxygen. Secondly, linkage of proline metabolic flux to metabolic pathways to maintain the intracellular redox homeostasis (NADP+/NADPH ratio, GSH pool).

Advertisement

6. Oxylipin signaling under heavy metal stress

Oxidized fatty acids, oxylipins, are an important class of signaling molecule in plants in responses to abiotic stresses [95, 96].

Oxylipins regulate growth, development, and responses to environmental stimuli of organisms [97]. Lipoxygenases, allene oxide synthase and a series of cytochromes P450 related oxygenasesare involved in oxylipin biosynthesis. Enzymatically synthesized oxylipin, the jasmonic acid (JA) and a precursor molecule 12-oxo-phytodienoic acid (OPDA) were shown to accumulate in response to pathogen infection [98]. Jasmonate-dependent tolerance to heavy metals is mediated by defensins, small cysteine-rich proteins present in plant cells [99, 100]. Oxylipins are involved in stress signal transduction, regulate stress-induced gene expression, and interact with other signaling pathways in plant cells, including signaling pathways of the plant hormones auxin, gibberellin, ethylene, and abscisic acid (ABA) [101, 102]. However, the role of oxylipins in plant adaptation and defense mechanism to abiotic stress is less studied. Protective action of oxylipins at toxic levels of heavy metals is considered to activate detoxification processes. Non-enzymatic reaction of reactive oxygen species with lipidic substances results in hydroxy fatty acids which are biologically active oxylipins and play important role in protective action of heavy metal stress [103]. Spontaneously formed oxylipins are called phytoprostanes. Pretreatment of tobacco plants with phytoprostanes results in reduction of cell death in response to copper sulfate stress [104]. In roots of tobacco seedlings Al-induced accumulation of 2-alkenals formed from fatty acid hydroperoxides was detected [105]. ROS generated during aluminum stress formed toxic aldehydes in reaction with fatty acids and caused severe root growth inhibition. Removal of 2-alkenals from tissue through overexpression of 2-alkenal reductase reduced Al phytotoxicity. Stimulation the expression of a number of stress responsive by phytoprostanes [106] makes oxylipins promising tools for improving stress tolerance of plants to heavy metals.

Advertisement

7. Conclusion

Heavy metal stress on plants in contaminated soils due to increased anthropogenetic activities led to an intensive research on detoxification and sequestering mechanisms of plants to understand and manipulate defense mechanisms developed in plants to tolerate stress by metal ions. The accumulation of heavy metals in plant organs results in the damage of the structural components and disruption of cell metabolic processes leading to plant growth retardation. Increased biosynthesis of various plant biomolecules to confer tolerance during toxic action of heavy metals is an intrinsic ability of plants. Induced formation of low-molecular weight amino acids, peptides or proteines as chelators such as proline (Pro), glutathione (GSH), phytochelatins (PCs) or metallothioneins (MTs) under heavy metal stress enhances metal binding and detoxification capability of plants. In addition, proline and GSH related enzymes such as GSH reductase, GSH peroxidases and glutathione S-transferases are also key components of the antioxidant defense system in the cells to scavenge reactive oxygen species (ROS). Protective action of oxidized fatty acids oxylipins at toxic levels of heavy metals is considered to activate detoxification processes as signaling molecules.

Exploring and manipulating genes induced under heavy metal stress make possible to develop transgenic plants with enhanced detoxication properties for phytoremediation technologies in polluted soils. The use of hyperaccumulator plants [107] which can accumulate heavy metals in the leaves at 100-fold higher concentration than normal plant species can be the alternative solution to clean-up polluted soils.

References

  1. 1. Rauser WE. Structure and function of metal chelators produced by plants: The case for organic acids, amino acids, phytin and metallothioneins. Cell Biochemistry and Biophysics. 1999;31:19-48. DOI: 10.1007/BF02738153
  2. 2. Hasanuzzaman M, Nahar K, Anee TI, Fujita M. Glutathione in plants: Biosynthesis and physiological role environmental stress tolerance. Physiology and Molecular Biology of Plants. 2017;23:249-268. DOI: 10.1007/s12298-017-0422-2
  3. 3. Noctor G, Mhamdi A, Chaouch S, Han Z, Neukermans J, Marquez-Garcia B, et al. Glutathione in plants: An integrated overview. Plant, Cell and Environment. 2012;35:454-484. DOI: 10.1111/j.1365-3040.2011.02400.x
  4. 4. Sharma SS, Dietz KJ. The significance of amino acids and amino acid-derived molecules in plant responses and adaptation to heavy metal stress. Journal of Experimental Botany. 2006;57:711-726. DOI: 10.1093/jxb/erj073
  5. 5. Ghanta S, Chattopadhyay S. Glutathione as a signaling molecule - another challenge to pathogens. Plant Signaling and Behavior. 2011;6:783-788. DOI: 10.4161/psb.6.6.15147
  6. 6. Zechmann B, Muller M. Subcellular compartmentation of glutathione in dicotyledonous plants. Protoplasma. 2010;246:15-24. DOI: 10.1007/s00709-010-0111-2
  7. 7. Zeng X, Qiu D, Hu R, Zhang M. Glutathione Transporters in Plants. In: Hossain M, Mostofa M, Diaz-Vivancos P, Burritt D, Fujita M, Tran LS, editors. Glutathione in Plant Growth, Development, and Stress Tolerance. Cham: Springer; 2017. pp. 359-372. DOI: 10.1007/978-3-319-66682-2_16
  8. 8. Mhamdi A, Hager J, Chaouch S, Queval G, Han Y, Taconnat L, et al. Arabidopsis glutathione reductase plays a crucial role in leaf responses to intracellular H2O2 and in ensuring appropriate gene expression through both salicylic acid and jasmonic acid signaling pathways. Plant Physiology. 2010;153:1144-1160. DOI: 10.1104/pp.110.153767
  9. 9. Vanacker H, Carver TLW, Foyer CH. Early H2O2 accumulation in mesophyll cells leads to induction of glutathione during the hyper-sensitive response in the barley powdery mildew interaction. Plant Physiology. 2000;123:1289-1300. DOI: 10.1104/pp.123.4.1289
  10. 10. Szalai G, Kellos T, Galiba G, Kocsy G. Glutathione as an antioxidant and regulatory molecule in plants under abiotic stress conditions. Plant Growth Regulators. 2009;28:66-80. DOI: 10.1007/s00344-008-9075-2
  11. 11. Mahmood Q , Ahmad R, Kwak SS, Rashid A, Anjum NA. Ascorbate and glutathione: Protectors of plants in oxidative stress. In: Mahmood Q , Ahmad R, Kwak SS, Rashid A, Anjum NA, editors. Ascorbate–glutathione pathway and stress tolerance in plants. Berlin: Springer; 2010. pp. 209-229. DOI: 10.1007/978-90-481-9404-9_7
  12. 12. Anjum NA, Ahmad I, Mohmood I, Pacheco M, Duarte AC, Pereira E, et al. Modulation of glutathione and its related enzymes in plants responses to toxic metals and metalloids—A review. Environmental and Experimental Botany. 2012;75:307-324. DOI: 10.1016/j.envexpbot.2011.07.002
  13. 13. Iqbal N, Masood A, Nazar R, Syeed S, Khan NA. Photosynthesis, growth and antioxidant metabolism in mustard (Brassica juncea L.) cultivars differing in cadmium tolerance. Agricultural Sciences in China. 2010;9:519-527. DOI: 10.1016/S1671-2927(09)60125-5
  14. 14. Chen F, Wang F, Wu F, Mao W, Zhang G, Zhou M. Modulation of exogenous glutathione in antioxidant defense system against Cd stress in the two barley genotypes differing in Cd tolerance. Plant Physiology and Biochemistry. 2010;48:663-672. DOI: 10.1007/s12011-011-9121-y
  15. 15. Aravind P, Prasad MNV. Modulation of cadmium-induced oxidative stress in Ceratophyllum demersum by zinc involves ascorbate–glutathione cycle and glutathione metabolism. Plant Physiology and Biochemistry. 2005;43:107-116. DOI: 10.1016/j.plaphy.2005.01.002
  16. 16. Ivanov AA. Role of glutathione in enhancing metal hyperaccumulation in plants. In: Hasanuzzaman M, MNV P, editors. Handbook of Bioremediation. Physiological, Molecular and Biotechnological Interventions. London: Academic Press; 2020. pp. 115-152. DOI: 10.1016/B978-0-12-819382-2.00008-9
  17. 17. Gill SS, Tuteja N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiology and Biochemistry. 2010;48:909-930. DOI: 10.1016/j.plaphy.2010.08.016
  18. 18. Anjum NA, Umar S, Iqbal M, Khan NA. Cadmium causes oxidative stress in moongbean [Vigna radiata (L.) Wilczek] by affecting antioxidant enzyme systems and ascorbate-glutathione cycle metabolism. Russian Journal of Plant Physiology. 2011;58:92-99. DOI: 10.1134/S1021443710061019
  19. 19. Fornazier RF, Ferreira RR, Vitoria AP, Molina SMG, Lea PJ, Azevedo RA. Effects of cadmium on antioxidant enzyme activities in sugar cane. Biologia Plantarum. 2002;45:91-97. DOI: 10.1023/A:1015100624229
  20. 20. Aravind P, Prasad MNV. Cadmium and zinc interactions in a hydroponic system using Ceratophylum demersum L.: Adaptive ecophysiology, biochemistry and molecular toxicology. Brazilian Journal of Plant Physiology. 2005;17:3-20. DOI: 10.1590/S1677-04202005000100002
  21. 21. Verma S, Dubey RS. Lead toxicity induces lipid peroxidation and alters the activities of antioxidant enzymes in growing rice plants. Plant Science. 2003;164:645-655. DOI: 10.1016/S0168-9452(03)00022-0
  22. 22. Bela K, Horváth E, Gallé Á, Szabados L, Tari I, Csiszár J. Plant glutathione peroxidases: Emerging role of the antioxidant enzymes in plant development and stress responses. Journal of Plant Physiology. 2015;176:192-201. DOI: 10.1016/j.jplph.2014.12.014
  23. 23. Leon AM, Palma JM, Corpas FJ, Gomez M, Romero-Puertas MC, Chatterjee D, et al. Antioxidant enzymes in cultivars of pepper plants with different sensitivity to cadmium. Plant Physiology and Biochemistry. 2002;40:813-820. DOI: 10.1016/s0981-9428(02)01444-4
  24. 24. Dixit V, Pandey V, Shyam R. Differential antioxidative responses to cadmium in roots and leaves of pea (Pisum sativum L. cv. Azad). Journal of Experimental Botany. 2001;52:1101-1109. DOI: 10.1093/JEXBOT/52.358.1101
  25. 25. Baccouch S, Chaoui A, Ferjani EE. Nickel toxicity induces oxidative damage in Zea mays roots. Journal of Plant Nutrition. 2001;24:1085-1097. DOI: 10.1080/01904169809365425
  26. 26. Hartikainen H, Kue TL, Piironem V. Selenium as an antioxidant and prooxidant in rye grass. Plant and Soil. 2000;225:193-200. DOI: 10.1023/A:1026512921026
  27. 27. Marrs KA. The functions and regulation of glutathione S-transferases in plants. Annual Review of Plant Physiology and Plant Molecular Biology. 1996;47:127-158. DOI: 10.1146/annurev.arplant.47.1.127
  28. 28. Frova C. Glutathione transferases in the genomics era: New insights and perspectives. Biomolecular Engineering. 2006;23:149-169. DOI: 10.1016/j.bioeng.2006.05.020
  29. 29. Townsend D, Tew K. The role of glutathione-S-transferase in anti-cancer drug resistance. Oncogene. 2003;22:7369-7375. DOI: 10.1038/sj.onc.1206940
  30. 30. Edwards R, Dixon DP, Walbot V. Plant glutathione S-transferases: Enzymes with multiple functions in sickness and in health. Trends in Plant Science. 2000;5:193-198. DOI: 10.1016/S1360-1385(00)01601-0
  31. 31. Cummins I, Cole DJ, Edwards R. A role for glutathione transferases functioning as glutathione peroxidases in resistance to multiple herbicides in black-grass. The Plant Journal. 1999;18:285-292. DOI: 10.1046/j.1365-313X.1999.00452.x
  32. 32. Iannelli MA, Pietrini F, Fiore L, Petrilli L, Massacci A. Antioxidant response to cadmium in Phragmites australis plants. Plant Physiology and Biochemistry. 2002;40:977-982. DOI: 10.1016/S0981-9428(02)01455-9
  33. 33. Halusková L, Valentovicová K, Huttová J, Mistrik I, Tamás L. Effect of abiotic stresses on glutathione peroxidase and glutathione S-transferase activity in barley root tips. Plant Physiology and Biochemistry. 2009;47:1069-1074. DOI: 10.1016/j.plaphy.2009.08.003
  34. 34. Marss KA, Walbot V. Expression and RNA splicing of the maize glutathione S-transferase of wheat bronze2 gene is regulated by cadmium and other stresses. Plant Physiology. 1997;47:127-158
  35. 35. Ha S-B, Smith AP, Howden R, Dietrich WM, Bugg S, O'Connell MJ, et al. Phytochelatin synthase genes from Arabidopsis and the yeast Schizosaccharomyces pombe. The Plant Cell. 1999;11:1153-1164. DOI: 10.1105/tpc.11.6.1153
  36. 36. Grill E, Winnacker EL, Zenk MH. Phytochelatins: The principal heavy-metal complexing peptides of higher plants. Science. 1985;230:674-676. DOI: 10.1073/pnas.86.18.6838
  37. 37. Zenk MH. Heavy metal detoxification in higher plants: A review. Gene. 1996;179:21-30. DOI: 10.1016/s0378-1119(96)00422-2
  38. 38. Dennis KK, Uppal K, Liu KH, Ma C, Liang B, Go Y-M, et al. Phytochelatin database: A resource for phytochelatin complexes of nutritional and environmental metals. Database. 2019;2019:1-9. DOI:10.1093/database/baz083
  39. 39. Hirata K, Tsuji N, Miyamoto K. Biosynthetic regulation of phytochelatins, heavy metal-binding peptides. Journal of Bioscience and Bioengineering. 2005;100:593-599. DOI: 10.1093/database/baz083
  40. 40. Gusmao R, Arino C, Díaz-Cruz JM, Esteban M. Electrochemical survey of the chain length influence in phytochelatins competitive binding by cadmium. Analytical Biochemistry. 2010;406:61-69. DOI: 10.1016/j.ab.2010.06.034
  41. 41. Song W-Y, Mendoza-Cózatl DG, Lee Y, Schroeder JI, Ahn S-N, Lee H-S, et al. Phytochelatin–metal(loid) transport into vacuoles shows different substrate preferences in barley and Arabidopsis. Plant Cell Environment. 2014;37:1192-1201. DOI: 10.1111/pce.12227
  42. 42. Brunetti P, Zanella L, Proia A, De Paolis A, Falasca G, Altamura MM, et al. Cadmium tolerance and phytochelatin content of Arabidopsis seedlings over-expressing the phytochelatin synthase gene AtPCS1. Journal of Experimental Botany. 2011;62:5509-5519. DOI: 10.1093/jxb/err228
  43. 43. Kotrba P, Macek T, Ruml T. Heavy metal-binding peptides and proteins in plants. A review. Collection of Czechoslovak Chemical Communications. 1999;64:1057-1086. DOI: 10.1002/CHIN.199942309
  44. 44. Thangavel P, Long S, Minocha R. Changes in phytochelatins and their biosynthetic intermediates in red spruce (Picea rubens Sarg.) cell suspension cultures under cadmium and zinc stress. Plant Cell Tissue and Organ Culture. 2007;88:201-216. DOI: 10.1007/s11240-006-9192-1
  45. 45. Cobbett SS. Heavy metal detoxification in plants: phytochelatin biosynthesis and function. IUBMB Life. 2001;51:183-188. DOI: 10.1080/152165401753544250
  46. 46. Fidalgo F, Azenha M, Silva AF, de Sousa A, Santiago A, Ferraz P, et al. Copper-induced stress in Solanum nigrum L. and antioxidant defense system response. Food and Energy Security. 2013;2:70-80. DOI: 10.1002/fes3.20
  47. 47. Batista BL, Nigar M, Mestrot A, Rocha BA, Barbosa F Jr, He PA, et al. Identification and quantification of phytochelatins in roots of rice to long-term exposure: Evidence of individual role on arsenic accumulation and translocation. The Journal of Experimental Botany. 2014;65:1467-1479. DOI: 10.1093/jxb/eru018
  48. 48. Schmöger MEV, Oven M, Grill E. Detoxification of arsenic by phytochelatins in plants. Plant Physiology. 2000;122:93-801. DOI: 10.1104/pp.122.3.793
  49. 49. Zhang Z-C, Chen B-X, Qiu B-S. Phytochelatin synthesis plays a similar role in shoots of the cadmium hyperaccumulator Sedum alfredii as in non-resistant plants. Plant, Cell and Environment. 2010;33:1248-1255. DOI: 10.1111/j.1365-3040.2010.02144.x
  50. 50. Hentz S, McComb J, Miller G, Begonia M, Begonia G. Cadmium uptake, growth and phytochelatin contents of Triticum aestivum in response to various concentrations of cadmium. World Environment. 2012;2:44-50. DOI: 10.5923/j.env.20120203.05
  51. 51. Inouhe M. Phytochelatins. Brazilian Journal of Plant Physiology. 2005;17:65-78. DOI: 10.1590/S1677-04202005000100006
  52. 52. Rauser WE, Meuwly P. Retention of cadmium in roots of maize seedlings. Role of complexation by phytochelatins and related thiol peptides. Plant Physiology. 1995;109:195-202. DOI: 10.1104/pp.109.1.195
  53. 53. Shah K, Nongkynrih JM. Metal hyperaccumulator and bioremediation. Biologia Plantarum. 2007;51:618-634. DOI: 10.1007/s10535-007-0134-5
  54. 54. Mendoza-Cozatl DG, Jobe TO, Hauser F, Schroeder JL. Long-distance transport, vacuolar sequestration, tolerance, and transcriptional responses induced by cadmium and arsenic. Current Opinion in Plant Biology. 2011;14:554-562. DOI: 10.1016/j.pbi.2011.07.004
  55. 55. Scheidegger C, Suter MJ-F, Behra R, Sigg L. Characterization of lead–phytochelatin complexes by nano-electrospray ionization mass spectrometry. Frontiers in Microbiology. 2012;3:41. DOI: 10.3389/fmicb.2012.00041
  56. 56. Scarano G, Morelli E. Characterization of cadmium- and lead- phytochelatin complexes formed in a marine microalga in response to metal exposure. Biometals. 2002;15:145-151. DOI: 10.1023/a:1015288000218
  57. 57. Joshi R, Pareek A, Singla-Pareek SL. Plant Metallothioneins: Classification, Distribution, Function, and Regulation. In: Ahmad P, editor. Plant Metal Interaction. Emerging Remediation Techniques. Amsterdan: Elsevier; 2016. pp. 239-261. DOI: 10.1016/B978-0-12-803158-2.00009-6
  58. 58. Leszczyszyn OI, Imam HT, Blindauer CA. Diversity and distribution of plant metallothioneins: A review of structure, properties and functions. Metallomics. 2013;5:1146-1169. DOI: 10.1039/c3mt00072a
  59. 59. Grennan AK. Metallothioneins, a diverse protein family. Plant Physiology. 2011;155:1750-1751. DOI: 10.1104/pp.111.900407
  60. 60. Cobbett C, Goldsbrough PB. Phytochelatins and metallothioneins: Roles in heavy metal detoxification and homeostasis. Annual Review of Plant Biology. 2002;53:159-182. DOI: 10.1146/annurev.arplant.53.100301.135154
  61. 61. Verkleij JAC, Sneller FEC, Schat H. Metallothioneins and phytochelatins: ecophysiological aspects. In: Abrol YP, Ahmad A, editors. Sulphur in Plants. Dordrecht, Netherlands: Springer; 2003. pp. 163-176. DOI: 10.1104/pp.111.900407
  62. 62. Blindauer CA, Leszczyszyn OI. Metallothioneins: Unparalleled diversity in structures and functions for metal ion homeostasis and more. Natural Product Report. 2010;27:720. DOI: 10.1039/b906685n
  63. 63. Freisinger E. Plant MT – long neglected members of the metallothionein superfamily. Dalton Transactions. 2008;(47):6663-6667. DOI: 10.1039/b809789e
  64. 64. Hassinen VH, Tervahauta AI, Schat H, Karenlampi SO. Plant metallothioneins – metal chelators with ROS scavenging activity? Plant Biology. 2011;13:225-232. DOI: 10.1111/j.1438-8677.2010.00398.x
  65. 65. Emamverdian A, Ding Y, Mokhberdoran F, Xie Y. Heavy metal stress and some mechanisms of plant defense response. The Scientific World Journal. 2015;2015:1-18. DOI: 10.1155/2015/756120
  66. 66. Guo W-J, Meetam M, Goldsbrough PB. Examining the specific contributions of individual Arabidopsis metallothioneins to copper distribution and metal tolerance. Plant Physiology. 2008;146:1697-1706. DOI: 10.1104/pp.108.11578
  67. 67. Murphy A, Taiz L. Comparison of metallothionein gene expression and nonprotein thiols in ten Arabidopsis ecotypes (correlation with copper tolerance). Plant Physiology. 1995;109:945-954. DOI: 10.1104/pp.109.3.945
  68. 68. Zimeri AM, Dhankher OP, McCaig B, Meagher RB. The plant MT1 metallothioneins are stabilized by binding cadmium and are required for cadmium tolerance and accumulation. Plant Molecular Biology. 2005;58:839-855. DOI: 10.1007/s11103-005-8268-3
  69. 69. Yang Z, Chu C. Towards understanding plant response to heavy metal stress, in Abiotic Stress in Plants—Mechanisms and Adaptations. Shanghai, China: InTech; 2011. pp. 59-78. DOI: 10.5772/24204
  70. 70. Klaassen CD, Liu J, Chodhuri S. Metallothionein: An intracellular protein to protect against cadmium toxicity. Annual Review of Pharmacology and Toxicology. 1999;39:267-294. DOI: 10.1146/annurev.pharmtox.39.1.267
  71. 71. Winge DR, Miklossy K-A. Differences in the polymorphic forms of metallothionein. Archives of Biochemistry and Biophysics. 1982;214:80-88. DOI: 10.1016/0003-9861(82)90010-8
  72. 72. Maret W. Zinc and sulfur: A critical biological partnership. Biochemistry. 2004;43:3301-3309. DOI: 10.1021/bi036340p
  73. 73. Tommey AM, Shi J, Lindsay WP, Urwin PE, Robinson NJ. 1Expression of the pea gene PsMT A in E. coli: Metal-binding properties of the expressed protein. FEBS Letters. 1991;292:48-52. DOI: 10.1016/0014-5793(91)80831-M
  74. 74. Liang X, Zhang L, Natarajan SK, Becker DF. Proline mechanisms of stress survival. Antioxidants and Redox Signaling. 2013;19:998-1011. DOI: 10.1089/ars.2012.5074
  75. 75. Farago ME, Mullen WA. Plants which accumulate metals. Part IV. A possible copper–proline complex from the roots of Armeria maritima. Inorganica Chimica Acta. 1979;32:L93-L94. DOI: 10.1016/S0020-1693(00)91627-X
  76. 76. Siripornadulsil S, Traina S, Verma DPS, Sayre RT. Molecular mechanisms of proline-mediated tolerance to toxic heavy metals in transgenic microalgae. The Plant Cell. 2002;14:2837-2847. DOI: 10.1105/tpc.004853
  77. 77. Matysik J, Alia BB, Mohanty P. Molecular mechanisms of quenching of reactive oxygen species by proline under stress in plants. Current Science. 2002;82:525-532
  78. 78. Choudhary M, Jetley UK, Abash Khan M, Zutshi S, Fatma T. Effect of heavy metal stress on proline, malondialdehyde, and superoxide dismutase activity in the cyanobacterium Spirulina platensis-S5. Ecotoxicology and Environmental Safety. 2007;66:204-209. DOI: 10.1016/j.ecoenv.2006.02.002
  79. 79. Sharma SS, Schat H, Vooijs R. In vitro alleviation of heavy metal-induced enzyme inhibition by proline. Phytochemistry. 1998;49:1531-1535. DOI: 10.1016/s0031-9422(98)00282-9
  80. 80. Perrin DD. Stability constants of metal-ion complexes. In: Part B: Organic Ligands. 2. Suppl. Z. 2. Oxford UK/Elmsford, NY, USA: Pergamon Press; 1979
  81. 81. Delauney AJ, Verma DPS. Proline biosynthesis and osmoregulation in plants. The Plant Journal. 1993;4:215-223. DOI: 10.1046/j.1365-313X.1993.04020215.x
  82. 82. Verbruggen N, Villarroel R, Van Montagu M. Osmoregulation of a pyrroline-5-carboxylate reductase gene in Arabidopsis thaliana. Plant Physiology. 1993;103:771-781. DOI: 10.1104/pp.103.3.771
  83. 83. Karimi LN, Khanahmadi M, Moradi B. Accumulation and phytotoxicity of lead in Cynara scolymus. Indian Journal of Science and Technology. 2012;5:3634-3641. DOI: 10.17485/ijst%2F2012%2Fv5i11%2F30653
  84. 84. Gohari M, Habib-Zadeh AR, Khayat M. Assessing the intensity of tolerance to lead and its effect on amount of protein and proline in root and aerial parts of two varieties of rape seed (Brassica napus L.). Journal of Basic and Applied Scientific Research. 2012;2:935-938
  85. 85. John R, Ahmad P, Gadgil K, Sharma S. Heavy Metal Toxicity: Effect on plant growth, biochemical parameters and metal accumulation by Brassica juncea L. International Journal of Plant Production. 2009;3:65-76
  86. 86. Lesko K, Simon-Sarkadi L. Effect of cadmium stress on amino acid and polyamine content of wheat seedlings. Periodica Polytechnica: Chemical Engineering. 2002;46:65-71
  87. 87. Handique GK, Handique AK. Proline accumulation in lemongrass (Cymbopogon flexuosus Stapf.) due to heavy metal stress. Journal of Environmental Biology. 2009;30:299-302
  88. 88. Nikolic N, Kojic D, Pilipovic A, Pajevic S, Krstic B, Borisev M, et al. Responses of hybrid poplar to cadmium stress: photosynthetic characteristics, cadmium and proline accumulation, and antioxidant enzyme activity. Acta Biologica Cracoviensia Series Botanica. 2008;50:95-103
  89. 89. Ruscitti M, Arango M, Ronco M, Beltrano J. Inoculation with mycorrhizal fungi modifies proline metabolism and increases chromium tolerance in pepper plants (Capsicum annuum L.). Brazilian. Journal of Plant Physiology. 2011;23:15-25. DOI: 10.1590/S1677-04202011000100004
  90. 90. Zengin FK, Kirbag S. Effects of copper on chlorophyll, proline, protein and abscisic acid level of sunflower (Helianthus annuus L.) seedlings. Journal of Environmental Biology. 2007;28:561-566
  91. 91. Theriappan P, Gupta AK, Dhasarrathan P. Accumulation of proline under salinity and heavy metal stress in cauliflowern seedlings. Journal of Applied Sciences and Environmental Management. 2011;15:251-255. DOI: 10.4314/jasem.v15i2.68497
  92. 92. Pant PP, Tripathi AK, Dwivedi V. Effect of heavy metals on some biochemical parameters of sal (Shorea robusta) seedling at nursery level, Doon Valley. India. Journal of Agricultural Science. 2011;2:45-45. DOI: 10.1080/09766898.2011.11884667
  93. 93. Vinod K, Awasthi G, Chauhan PK. Cu and Zn tolerance and responses of the biochemical and physiochemical system of wheat. Journal of Stress Physiology and Biochemistry. 2012;8:203-213
  94. 94. Saradhi A, Saradhi PP. Proline accumulation under heavy metal stress. Journal of Plant Physiology. 1991;138:554-558. DOI: 10.1016/S0176-1617(11)80240-3
  95. 95. Eckard N. Oxylipin signaling in plant stress responses. The Plant Cell. 2008;20:495-497. DOI: 10.1105/tpc.108.059485
  96. 96. Savchenko TV, Zastrijnaja OM, Klimov VV. Oxylipins and plant abiotic stress resistance. Biochemistry (Moscow). 2014;79:362-375. DOI: 10.1134/S0006297914040051
  97. 97. Howe GA, Schilmiller AL. Oxylipin metabolism in response to stress. Current Opinion in Plant Biology. 2002;5:230-236. DOI: 10.1016/s1369-5266(02)00250-9
  98. 98. Block A, Schmelz E, Jones JB, Klee HJ. Coronatine and salicylic acid: The battle between Arabidopsis and Pseudomonas for phytohormone control. Molecular Plant Pathology. 2005;6:79-83. DOI: 10.1111/j.1364-3703.2004.00265.x
  99. 99. Thomma BP, Cammue BP, Thevissen K. Plant defensins. Planta. 2002;216:193-202. DOI: 10.1007/s00425-002-0902-6
  100. 100. Mirouze M, Sels J, Richard O, Czernic P, Loubet S, Jacquier A, et al. A putative novel role for plant defensins: A defensin from the zinc hyperaccumulating plant, Arabidopsis halleri, confers zinc tolerance. The Plant Journal. 2006;47:329-342. DOI: 10.1111/j.1365-313X.2006.02788.x
  101. 101. Lopez MA, Vicente J, Kulasekaran S, Vellosillo T, Martinez M, Irigoyen ML, et al. Antagonistic role of 9-lipoxygenase-derived oxylipins and ethylene in the control of oxidative stress, lipid peroxidation and plant defence. The Plant Journal. 2011;67:447-458. DOI: 10.1111/j.1365-313X.2011.04608.x
  102. 102. Yang DL, Yao J, Mei CS, Tong XH, Zeng LJ, Li Q , et al. Plant hormone jasmonate prioritizes defense over growth by interfering with gibberellin signaling cascade. Proceedings of the National Academy of Sciences USA. 2012;109:E1192-E1200. DOI: 10.1073/pnas.1201616109
  103. 103. Sattler SE, Mene-Saffrane L, Farmer EE, Krischke M, Mueller MJ, Della Penna D. Nonenzymatic lipid peroxidation reprograms gene expression and activates defense markers in Arabidopsis tocopherol deficient mutants. The Plant Cell. 2006;18:3706-3720. DOI: 10.1105/tpc.106.044065
  104. 104. Setlik I, Allakhverdiev SI, Nedbal L, Setlikova E, Klimov VV. Three types of photosystem II photoinactivation: I Damaging processes on the acceptor side. Photosynthesis Research. 1990;23:39-48. DOI: 10.1007/BF00030061
  105. 105. Yin L, Mano J, Wang S, Tsuji W, Tanaka K. The involvement of lipid peroxide-derived aldehydes in aluminum toxicity of tobacco roots. Plant Physiology. 2010;152:1406-1417. DOI: 10.1104/pp.109.151449
  106. 106. Thoma I, Loeffler C, Sinha AK, Gupta M, Krischke M, Steffan B, et al. Cyclopentenone isoprostanes induced by reactive oxygen species trigger defense gene activation and phytoalexin accumulation in plants. The Plant Journal. 2003;34:363-375. DOI: 10.1046/j.1365-313X.2003.01730.x
  107. 107. Rascio N, Navari-Izzo F. Heavy metal hyperaccumulating plants: How and why do they do it? And what makes them so interesting? Plant Science. 2011;180:169-181. DOI: 10.1016/j.plantsci.2010.08.016

Written By

Istvan Jablonkai

Submitted: 19 July 2021 Reviewed: 21 December 2021 Published: 21 March 2022