Next Article in Journal
Candida auris Updates: Outbreak Evaluation through Molecular Assays and Antifungal Stewardship—A Narrative Review
Next Article in Special Issue
Molecular Biomarkers of Canine Reproductive Functions
Previous Article in Journal
Employing Bidirectional Two-Sample Mendelian Randomization Analysis to Verify the Potential of Polyunsaturated Fatty Acid Levels in the Prevention of Pancreatic Cancer
Previous Article in Special Issue
Multi Omics Applications in Biological Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cadmium Stress Signaling Pathways in Plants: Molecular Responses and Mechanisms

by
Valentina Vitelli
1,2,
Agnese Giamborino
2,
Andrea Bertolini
2,
Alessandro Saba
2 and
Andrea Andreucci
1,*
1
Department of Biology, University of Pisa, 56126 Pisa, Italy
2
Department of Surgical, Medical and Molecular Pathology and Critical Care Area, University of Pisa, 56126 Pisa, Italy
*
Author to whom correspondence should be addressed.
Curr. Issues Mol. Biol. 2024, 46(6), 6052-6068; https://doi.org/10.3390/cimb46060361
Submission received: 14 May 2024 / Revised: 6 June 2024 / Accepted: 10 June 2024 / Published: 14 June 2024
(This article belongs to the Special Issue Latest Review Papers in Molecular Biology 2024)

Abstract

:
Heavy metal (HM) pollution, specifically cadmium (Cd) contamination, is a worldwide concern for its consequences for plant health and ecosystem stability. This review sheds light on the intricate mechanisms underlying Cd toxicity in plants and the various strategies employed by these organisms to mitigate its adverse effects. From molecular responses to physiological adaptations, plants have evolved sophisticated defense mechanisms to counteract Cd stress. We highlighted the role of phytochelatins (PCn) in plant detoxification, which chelate and sequester Cd ions to prevent their accumulation and minimize toxicity. Additionally, we explored the involvement of glutathione (GSH) in mitigating oxidative damage caused by Cd exposure and discussed the regulatory mechanisms governing GSH biosynthesis. We highlighted the role of transporter proteins, such as ATP-binding cassette transporters (ABCs) and heavy metal ATPases (HMAs), in mediating the uptake, sequestration, and detoxification of Cd in plants. Overall, this work offered valuable insights into the physiological, molecular, and biochemical mechanisms underlying plant responses to Cd stress, providing a basis for strategies to alleviate the unfavorable effects of HM pollution on plant health and ecosystem resilience.

1. Introduction: Environmental Heavy Metal Contamination

Heavy metals (HMs) are elements characterized by an atomic number greater than 20, an atomic density exceeding 5 g/cm3, and metallic properties. Transition metals, like cadmium (Cd), mercury (Hg), and chromium (Cr), elements in the bottom left section of the periodic table like lead (Pb) and tin (Sn), and certain metalloids like arsenic (As) are essentially what are referred to as HMs. The latter can be broadly classified into two categories: essential and non-essential HMs. The first ones are necessary for living organisms to perform fundamental processes such as growth, metabolism, and organ development. Therefore, HMs, such as iron (Fe), zinc (Zn), copper (Cu), and cobalt (Co) function as cofactors that support redox and protein structure processes, making them crucial for plant growth and development. The non-essential ones are dangerous because they can bind to a variety of functional groups in biomolecules, including amino, carboxylic acid, and sulfur-containing groups [1]. Moreover, they compete with the plant for nutrient uptake; they are dangerous even at low concentrations as they can result in macronutrient deficiencies. Increased HM buildup results in acute toxicity, growth inhibition, and ultimately plant death [2]. In contrast to organic pollutants, HMs are non-biodegradable and tend to accumulate in living organisms [3]. The pervasive problem of HM pollution in the environment is a significant risk to public health worldwide. Reports indicate that 20 million hectares of China’s farmland, constituting roughly 20% of the nation’s total agricultural area, are affected by HM pollution, notably Cd, Pb, As, and Hg. Within this affected area, 14,000 hectares are particularly afflicted by Cd contamination [1,4]. Groundwater, agricultural products, and crops alike will soon suffer the far-reaching consequences of the presence of Cd in soil [5]. Pollution of the environment by HMs can result from both natural processes and human activities. Among the latter, the primary artificial causes of soil contamination include farming practices, urban sprawl, industrial activities, mining operations [6], and the expansion of transport infrastructure [7]. Industrial operations contribute to ecological pollution since HMs and their numerous chemical combinations are frequently employed as catalysts in chemical reactions. Transportation-wise, the emission of gases from tires, brakes, and gasoline fuel have been shown to be the main contributors of Pb and Cd emissions to the air due to increased traffic [8]. Concerning the mining activity, although there are no mines present nearby, agricultural goods cultivated in roadside fields have higher levels of HM contamination than crops grown in other parts [9]. Moreover, the mine environment is further polluted by soil-borne metal particles carried by wind or mineral transportation [10]. The soil pollution of HMs in close proximity to mines is also significantly influenced by climate [10,11]. The expanding trash mounds have a negative effect on the surrounding areas of mining and could potentially transform some agricultural regions into wasteland [12]. Besides, one of the primary causes of ecological contamination in these areas is the water utilized in the mining process. People living close to mines suffer from a variety of autoimmune and non-autoimmune disorders that can be fatal due to toxic HMs in the surrounding soil [12]. Furthermore, human-induced soil modifications, such as the standard agricultural methods used in different areas, cause crops to take in HMs. The most common causes of liquid-induced HM environmental pollution are the application of phosphate fertilizers and the irrigation of agricultural lands with dirty water [13].
One trace element that is not necessary for biological systems yet is present in large quantities in the environment is Cd. Due to its exceptional toxicity to all biological systems, Cd is an HM that is frequently released into the environment by both natural and man-made sources [14]. For these reasons, it is considered a major metal pollutant [15,16]. Cd and its compounds are soluble, making them more mobile than other metals in environmental matrices. They also have high bioavailability, which makes them easily absorbable by plants. Once inside the plant, Cd can interfere with important physiological processes [17,18]. Given that Cd is highly toxic, this poses a significant risk to human health, leading to cancer and extensive environmental harm [19]. According to numerous studies, dumps often contain Cd levels exceeding the permissible limit in soil [20,21,22,23]. With a global mean of 0.36 mg/kg, Cd usually occurs in soils at amounts ranging from 0.01 to 1 mg/kg [16] and can be found in soil water at up to 5 μg/L [24] and groundwater at up to 1 μg/L [25]. Increases in Cd concentrations in soils and groundwater, which are crucial for preserving a wholesome food supply and clean drinking water, can come from both natural and man-made sources. The WHO Guidelines for Drinking-Water Quality recommend a maximum value of 3 μg/L for Cd. The potential for Cd accumulation in plants to transfer to humans via the food chain, causing serious harm, is nowadays a considerable concern [26,27]. Due to competition with calcium (Ca) and other nutrients, chronic Cd poisoning, also known as itai-itai disease and initially identified in Japan in the early 20th century, results in renal tubular dysfunction, osteomalacia, and osteoporosis [28,29,30]. According to Khan et al. 2017 [30], Cd exposure is also linked to problems with glucose metabolism, lung and breast cancer, myocardial infarction, and heart failure.

2. Effects of Cd in Plants

Crops, agricultural goods, and groundwater are significantly affected by Cd [5], a trace element that is not necessary for plant growth, although it does accumulate in all sections of plants after absorption [2,31]. The first signs of Cd toxic effects on plants include browning of the roots, a decrease in the number of lateral roots, and the death of the tips of the roots due to inhibition of root growth. Furthermore, the accumulation of Cd could result in stem development retardation, terminal leaf chlorosis, and leaf curl [32,33]. By substituting the sulfhydryl groups of proteins and enzymes, Cd replaces the essential elements of enzyme active sites needed in biological reactions. It also modifies macromolecule conformation, causes protein denaturation, and deteriorates cell membranes. The plant defense system reacts to the entry of Cd into plant cells by accumulating proline [34,35,36], variations in the synthesis and levels of phytohormones [37,38,39], and the initiation of several signaling pathways, including the Ca-calmodulin pathway as well as the activation of various antioxidant enzymes [40,41].
Cd significantly increases the production of reactive oxygen species (ROS) after binding with proteins and enzyme complexes, which leads to oxidative stress [42,43]. Reactive nitrogen species (RNS) and ROS are byproducts of regular cellular metabolism and are known to play two distinct roles in biological systems [44]. High ROS levels cause the cell to arrest and eventually become necrotic or apoptotic, whereas low ROS levels accelerate the cell cycle’s progression [45]. The toxicity due to elevated ROS concentrations reflects the oxidation of proteins, DNA, and membrane lipids, as well as the breakdown of polyphenols, catecholamines, and thiols. Previous reports highlight ROS dual function in plant defense systems as a signaling molecule and as a source of toxicity [46,47]. According to multiple studies [48,49] these molecules are produced by the Fenton or Haber-Weiss reaction, respiratory burst oxidase homologs (also known as NADPH oxidases), and the inhibition of antioxidative enzymes, such as oxidases and peroxidases. The cell has evolved a sophisticated antioxidant system that may shield biological molecules from potentially hazardous, purely physiological, species-producing oxidative stress through both enzymatic and non-enzymatic methods. Proteins like catalase (CAT), glutathione peroxidase (GPx), ascorbic acid (Vitamin C), ϱ-tocopherol (Vitamin E), glutathione (GSH), carotenoids, flavonoids, and other antioxidant agents are part of the enzymatic defense system [50] (Figure 1).

3. Plant Responses to Cd Stress and Detoxification via PCn Production

Different plant species are affected by nutrient variance under Cd stress [15,51,52]. Certain molecular alterations control Cd mitigation, tolerance, and adaptation, including the overexpression of specific genes. It has been observed that natural resistance and macrophage proteins (NRAMPs), iron-regulated transporter-like proteins (IRTs), zinc-regulated transporter-like proteins (ZIPs), heavy metal ATPases (HMAs), and metal tolerance or transporter proteins (MTPs) play roles in the uptake, translocation, and sequestration of Cd within plants [53]. Jia et al. 2019 [54] revealed that a high concentration of Cd (6.5 µM Cd) increases the expression of genes (SlQUA1 and SlPME1) associated with pectin synthesis in the cell walls of tomatoes, leading to an increase in pectin methyl esterase activity and ion binding sites, which counteracts Cd stress. Additionally, Song, Feng, et al. 2017 [55] demonstrated that overexpression of AtFC1 in A. thaliana enhances Cd accumulation and tolerance under Cd exposure. Conversely, genes like HvIRT1 and HvNramp5 in barley [56] downregulate Cd uptake, while lower expression of HMA-A, HMA-B, IRT1, and ZIP1 in tobacco leads to reduced Cd accumulation [57]. Genes such as ZIP2, ZIP3, IRT1, HMA2, and HMA4 and CAX4, HMA3, MRP7, MTP3, and COPT5 are overexpressed in Pak choi (B. rapa) grown in sand artificially contaminated with 10 mg Cd Kg−1 [58]. Other studies on the response to Cd stress have been conducted in rape, A. thaliana, maize, rice, and other crops [59,60]. NcNramp1 was shown to be crucial for the process of plant absorption and accumulation of the ionic form Cd2+ when transcriptome technology was used to examine N. caerulescens under Cd stress in various genotypes [61].
As a result of the significant toxic impact on plant growth and development, plants have evolved a variety of tolerance and detoxification strategies to control the uptake, transportation, and accumulation of Cd and lessen its toxicity [62]. The plant root hairs, along with the cuticle and epidermis, produce protective tissue to keep Cd out [63]. They also release organic acids, including citric, malic, and lactic acids, which mix with Cd to form compounds that prevent Cd from passing through membranes. In order to reduce the quantity of Cd that migrates from the soil into the root, they also modify the pH and EH of the rhizosphere [64]. Furthermore, to reduce the toxicity of Cd to cell activity, Cd that does enter cells will be enriched in vacuoles. This enrichment occurs due to the ability of polysaccharides and proteins in the cell wall to bond with the metal and prevent it from entering the cell [65].
Plants, particularly the hyperaccumulator ones, react in different ways to the toxicity of HMs, including Cd. HM chelation utilizing peptide ligands called phytochelatins (PCn) is one of the novel defense mechanisms [66]. These peptides are primarily composed of (c-Glu-Cys)n-Gly, with (2 < n < 11), although other PCn structures such as (c-Glu-Cys)n-Glu, (c-Glu-Cys)n-Ser, and (c-Glu-Cys)n-c-Ala have also been identified [67,68]. PCn are synthesized from GSH by the activation of the enzyme phytochelatin synthase (PCS), a γ-EC dipeptidyl (trans)peptidase that is constitutively expressed in the plant cytosol [69,70,71]. Synthesized PCn can form a complex with Cd, referred to as a chelator-Cd complex. Intracellular Cd promotes PCS activity, and the subsequent metal-induced production of PCn allows intracellular Cd detoxification [72]. Indeed, the PCn complex aids in transporting Cd from the cytosol into the vacuole, where it becomes inactive [73,74] (Figure 2).
A study reports the exogenous presence of GSH to reduce Cd translocation to shoots and enhance Cd accumulation in B. campestris roots treated with 20 mM of Cd [40]. According to the authors, exogenous GSH promotes the synthesis of PCn by increasing Cd chelation and facilitating its translocation into the root vacuoles. In fact, recent studies [72,75,76,77] confirm that exposure to Cd raises GSH levels in addition to activating the constitutively expressed enzyme PCS in M. polymorpha (MpPCS), which significantly increases the amount of PCn generated overall. Particularly, a detailed gene expression analysis confirmed the constitutive transcription of MpPCS; indeed, Cd is able to induce the activity, but not the gene expression, of the enzyme [72]. As a result, PCn synthesis is essential for plant detoxification and tolerance to high Cd concentrations. Another notable instance is seen in rice plants, where PCn increases by 23% and 47%, respectively, in response to 1 and 2 mM Cd [78]. Similarly, B. auriculata treated with 125 mM of Cd showed similar outcomes [70]. In response to metallic micronutrients including Fe, Zn, Cu, and the hazardous element Cd, PCn activation has recently been found in tracheophytes (mostly in angiosperms), as well as bryophytes and charophytes [79,80,81]. Therefore, M. polymorpha may be a great model organism to study the cellular and molecular mechanisms involved in Cd detoxification. Studies examining how metal detoxifying mechanisms evolved from aquatic to terrestrial environments can greatly benefit from the study of bryophytes. Not to mention, these plants do not form strong hydrophobic barriers, have a very high surface/volume ratio, an elevated cation exchange capacity, and are hence more susceptible to unrestricted, seemingly uncontrolled metal absorption [18,82]. Consequently, bryophytes have been employed as a crucial biological monitoring system for metal environmental pollution because of their widespread distribution [82,83,84]. Additionally, a study conducted by Bellini et al. 2020 [85] examined the medium acidification of this liverwort and its thiol-peptide synthesis (GSH, PCn) following a 72 h exposure to Cd. The loss of selective permeability, a well-known property of biological membranes, can lead to swelling or shrinking of cell compartments due to HM-induced membrane damage or energy loss. Moreover, changes in chloroplast ultrastructure have been linked to HM concentrations in studies conducted both in the field and in vitro on bryophytes [83,86]. L. riparium has been explored to determine whether this moss could be suitable for biomonitoring in natural field conditions. Recent studies showed that Cd exposure severely affected the amount of photosynthesis, PCn, GSH, and nitrogen metabolism in L. riparium plants. Nitrogen (N2) and Cd metabolism are known to be closely correlated [87,88,89,90]. N2 is typically associated with the abiotic stress response because it competes with essential reductants required for antioxidant responses, particularly in the presence of metals and metalloids [89,91]. High N2 levels promote the uptake of Fe, Zn, Cu, Ca, Hg, and other cations in certain plants [90]. Instead, the activity of glutamate dehydrogenase, another important enzyme involved in abiotic stress, is reduced because it serves as a substitute route for assimilating N and detoxifies excess NH4 that is produced under stressful circumstances [91,92]. On the other hand, several major crops, such as rice, beans, peas, and corn, exhibited reduced glutathione synthase (GS) activity in response to HM uptake [93,94]. Research conducted by Maresca et al., 2022 [95] demonstrated that L. riparium has a natural tendency to increase the occurrence of GS2 proteins. Cd exposure causes the metal to be chelated in the roots and transferred to the shoots via GSH and PCn [90]. Maintaining this chelation mechanism, which controls the amount of GSH and non-protein thiols, as well as the expression of PCS and GSH synthase genes, requires a sufficient absorption of N2 [90].

4. Role of GSH in Mitigating Cd-Induced Abiotic Stress

Due to the implication of GSH in mitigating HM stress, which is linked to its involvement in PCn synthesis, it is relevant to deepen our understanding of the GSH response during abiotic stress. Two ATP-dependent enzymes, glutamate-cysteine ligase (GSH1, previously known as γ-glutamylcysteine synthetase) and GSH synthase (GSH2), sequentially synthesize GSH from its three amino acid building blocks [96]. The first enzyme, GSH1, conjugates Cys and Glu to produce the dipeptide γ-glutamylcysteine. The second step is catalyzed by GSH2, which joins Gly to γ-glutamylcysteine to generate GSH. Growth and viability are restored when cytosol-specific GSH2 is added to the seedling-lethal GSH2 mutant strains. Therefore, for normal plant growth, cytosolic production of GSH alone is sufficient [97]. Although both GSH1 and GSH2 are relatively widely expressed in plant tissues, biotic stressors such as methyl jasmonate and high light [98], as well as abiotic stresses like HMs [99,100], but not H2O2 or oxidative agents [99], increase transcriptional levels of both GSH1 and GSH2. Its homodimerization is therefore thought to increase enzyme stability by shielding sensitive regions [101]. Since the chloroplast is home to GSH1, the redox state of this organelle plays a crucial role in controlling the synthesis of GSH. Due to the inherent instability of oxygenic photosynthesis, chloroplasts are highly reactive to unfavorable environmental conditions, which subsequently provides quick responses to abiotic and biotic challenges [102]. The recognized function of GSH in maintaining intracellular redox homeostasis largely depends on the redox control of GSH biosynthesizing enzymes. Indeed, oxidative stress-induced metabolic circumstances create a demand for GSH and the production of its oxidated form, GSSG, while also promoting GSH synthesis through the post-translational activation of GSH1. Consequently, elevated GSH levels establish a reducing environment, triggering a regulatory feedback loop to slow down GSH synthesis and decrease GSH1 activity. The activity of the sulfate assimilation pathway also influences GSH synthesis [103]. Cys serves as the initial organic product of sulfur assimilation in plants, and thus, its availability plays a crucial role in determining GSH production [103,104]. Consequently, several regulatory mechanisms influence the GSH biosynthesis process. Various environmental conditions and stresses impact the coarse control of GSH1 and GSH2 expression. Additionally, precursor availability affects the system’s activity, and a redox mechanism that considers feedback information finely regulates GSH1 activity. Currently, it remains uncertain whether a redox-coupling mechanism is necessary for the transformation of GSH1 from its oxidized to reduced form. The redox state of GSH, along with its cellular concentration, plays a critical role in mediating and finely regulating metabolism and stress responses. The flavoenzyme NADPH-dependent GSH reductase (GR) efficiently regenerates GSSG to GSH and permits recurrent redox cycling. In order to sustain fundamental functions in the cytosol, mitochondria, and chloroplast stroma in the light, GR contributes to the maintenance of a negative EGSH (i.e., high GSH/GSSG). Gill et al. 2013 [105] conducted a thorough literature analysis about GR genes and found evidence that a variety of abiotic stressors, including exposure to HMs, including Cd, salt, drought, UV radiation, and cold temperatures, boost GR activity in many plant species. Moreover, several stress-tolerant plants exhibit significant GR activity [106]. Because of its effective ROS scavenging capacity, transgenic techniques to control GR activity in plants also demonstrate that increased GR activity significantly improves tolerance to oxidative stress brought on by a range of abiotic stimuli [105]. Two nuclear genes, GR1 (At3g24170) and GR2 (At3g54660), which both code for differently localized isoforms, are responsible for encoding GRs in A. thaliana. GR1 is more highly expressed in roots, while GR2 expression is more pronounced in photosynthetic tissues [98]. GR1 is localized in the cytoplasm, nucleus, and peroxisomes [107,108,109,110] while GR2 targets mitochondria as well as chloroplasts [108].

5. Transporters Involved in Cd Remediation

There are various forms of Cd in soil, but many of them are not absorbed by plants [111]. For Cd to be absorbed, it must be available for uptake, which depends on the plant species, the physicochemical conditions of the soil, and the speciation of the metals [112,113].
A number of technologies have been developed to lower soil Cd levels. Phytoremediation, used by plants to absorb Cd and store it in easily harvested aerial sections of the plant, is one of the approaches [114]. The choice of appropriate plants that can grow in Cd-polluted soils and accumulate Cd in aerial parts is crucial to the effectiveness of phytoremediation. Over the past ten years, Cd hyperaccumulating plants have received attention [114]. However, only a few plant species that can withstand high amounts of Cd have been identified, including A. halleri, N. caerulescens, and S. alfredii.
Meighan et al., 2011 [115] found that herbaceous Cd hyperaccumulators had limited relevance for phytoremediation in Cd-polluted fields due to their slow development and low biomass output. The use of Populus species has been suggested as an alternative for phytoremediation because these woody species can be managed in short rotation coppicing plantations and are known for their fast growth and deep rooting systems [116,117,118,119].
Metal uptake is significantly influenced by root features such as surface area, size, and shape [113,120,121,122,123,124]. Furthermore, plants with thin, hairy roots have demonstrated high absorption and accumulation of metals [125]. In most cases, bivalent forms of trace elements are absorbed [126]. The same transporters that carry Ca, Fe, Mg, Cu, and Zn must carry Cd across root cell walls [15]. By passive transport (diffusion), Cd can move from the soil to plant roots through the cell walls [120,127]. Furthermore, studies have shown that Cd uses nonspecific membrane transport proteins (IRTs and ZIPs) and metal pum** ATPases to facilitate active transport across the plasma membrane of root cells [128,129]. Different transporters are used by plants to sequester metal ions inside organelles, whether in a free ionic state or as chelates. Plants can store extra metals in vacuoles when they are in neutral chelated forms. Long-distance transportation also makes use of chelated metal forms. To lessen the harmful effects of metals, metal chelators can be proteins that bind to metals or other compounds, such as amino acids, their derivatives, or organic acids [130].
These metal chelators also exhibit antioxidant properties. By detoxifying metal ions, specific protein metal chelators like glutaredoxins (GRXs), metallothionins (MTs), and PCn are essential in reducing the effects of metal stress-mediated oxidative damage [130]. Thereby, according to Kumar et al., 2020 [131], the overexpression of glutaredoxin CaGRX in chickpeas decreases the absorption of metal ions and boosts the ROS scavenging mechanism, thus mitigating metal stress.
Transporters serve dual roles as receptors and transporters, known as transceptors, within the plant’s sensory apparatus aimed at preserving metal homeostasis [132,133,134,135]. Transceptors have the capability to discern the availability of metal ions in soil and their scarcity within the cytosol, thereby orchestrating the regulation of metal ion uptake through low- or high-affinity transporters. This intricate process leads to a substantial decrease in metal ions, effectively maintaining metal homeostasis within plant cells through the expulsion and exclusion mechanisms facilitated by transporters.
Among these transporters, Cd, Co, Fe, Mn, and Zn divalent ions are transported by the IRT1. Within root cells, IRT1 acts as a transceptor within sensing and signaling mechanisms, prioritizing the sensing of Fe above other ions. Additionally, the IRT1 regulatory domain senses additional ions such as Co, Mn, and Zn, thereby inhibiting IRT1 function [132]. In a different study, CBL-interacting protein kinase 23 (CIPK23) phosphorylates the metal-bound IRT1, leading to ubiquitination by the IRT1 degradation factor 1 (IDF1) protein. This ubiquitination targets IRT1 for breakdown in the vacuole, effectively halting further metal absorption [136].
Wu et al., 2020 [137] discovered in their intriguing study on A. thaliana that the use of hydrogen-rich water accelerates the uptake of Cd. Elevated levels of Cd result in hydrogen production, which in turn triggers a cytosolic influx of Ca. Moreover, the activation of respiratory burst oxidase homolog protein D (RBOHD) by Ca elevation results in the production of ROS, primarily H2O2. In the end, the induced-H2O2 can inhibit IRT1 transcripts, which will lessen the absorption of more Cd [137].
Another instance in which Cd stress is mitigated is shown in apples, where the kinase salt overly sensitive 2-like1 (SOS2L1) phosphorylates the malate transporter aluminum-activated malate transporter 14 (ALMT14), enhancing its functionality [49]. Otherwise, a rice natural resistance-associated macrophage protein 1 (OsNRAMP1) plasma-membrane-localized Fe transporter that is involved in absorbing Cd is overexpressed, increasing the amount of Cd stored in leaves. Additionally, it was found that the expression of OsNRAMP1 is correlated with the amount of Cd found across different rice cultivars. Compared to the hyperaccumulating cultivar Sasanishiki, it was expressed up to two times greater in the Cd hyperaccumulating rice cultivar Habataki [138]. According to Lang et al., 2011 [139], B. juncea exhibits Cd tolerance through the involvement of two cation efflux transporters, BjCET3 and BjCET4, whose overexpression results in increased Cd accumulation in the shoot without lowering plant biomass.
Otherwise, in A. thaliana, the gene AtNRAMP6 confers Cd tolerance by locating at the Golgi apparatus. By transferring the Fe contained in Golgi vesicles into the cytosol during Fe depletion, AtNRAMP6 overexpression plays a critical role in preserving intracellular Fe equilibrium [140]. Consequently, by compartmentalizing and redistributing metals to specific organelles, including vacuoles, the endoplasmic reticulum, chloroplasts, and mitochondria, Golgi apparatus-localized metal transporters are necessary to maintain metal homeostasis of Fe, Cd, and Mn.
Furthermore, plants utilize energy-dependent metal transport systems to expel or remove metals, primarily relying on HM ATPases (HMAs) and ABC (ATP-binding cassette) transporters. Typically, these transporters expel metals from root cells into the surrounding air or other sections of the plant, where they can accumulate harmlessly. The role of ABC transporters in the translocation and redistribution of metals has been extensively studied. Fu et al. 2019 [141] demonstrated that Cd treatment in rice promoted the expression of the G-type ABC transporter ABCG36. A loss-of-function mutant exhibited a higher Cd concentration in its roots, suggesting that ABCG36 facilitates Cd translocation from roots to shoots. An interesting study revealed that the P. tomentosa ABCG36 gene can be overexpressed to improve Cd tolerance in A. thaliana plants, suggesting that this transporter serves as a Cd extrusion pump [142]. In accordance with a different study, the A. thaliana ABC transporter pleiotropic drug resistance 8 (AtPDR8) acts as a plasma membrane-localized Cd efflux pump, pushing Cd out of the cytosol and onto the cell’s exterior, thus providing HM stress tolerance [143]. In certain plants, only C-type ABC transporters (ABCCs) localize on tonoplasts and contribute to metal tolerance. According to Brunetti et al. 2015 [144] and Song et al. 2010 [145], tonoplast-localized ABCCs in A. thaliana, such as AtABCC1, AtABCC2, and AtABCC3, facilitate the transport of PCn and its complexes with As, Cd, Mn, and Zn into the vacuole, thus enhancing metal tolerance.
A recent study compared wild-type (WT) P. alba clone Villafranca plants with trans-genic plants overexpressing aquaporin (aqua1) in a hydroponic treatment with 10μM Cd. PaHMA2, PaNRAMP1.3, PaNRAMP2, PaNRAMP3.1, PaNRAMP3.2, PaABCC9, and PaABCC13 gene expressions were evaluated, revealing that the majority of the genes analyzed exhibited a prompt transcriptional response in WT plants at 1 day post-treatment and adaptation at 60 days. Conversely, aqua1 plants showed a weaker transcriptional response along with a higher Cd content in the medial leaves. In conclusion, aqua1 overexpression in poplar enhanced Cd translocation, indicating that aqua1 plants are less sensitive to Cd. Given the significance of PaNRAMP3 in Cd compartmentalization, a different transcription of this gene in the WT line may account for this distinct response. Nevertheless, even after two months of treatment, Cd administration did not cause observable toxicity symptoms in WT and aqua1 plants, demonstrating the remarkable resistance of this poplar species to Cd [146].
As part of a redistribution strategy, the HMA family of P1B-ATPases also serves as metal pumps, transporting metals into the xylem, phloem, and apoplast for transfer to other plant components, including shoots and grains. Members of the HMA family are essential for the mobilization of metal ions in a variety of plants, including Cu, Cd, and Zn. HvHMA1 is essential for scavenging Cu and Zn from the chloroplast to the cytosol in barley, facilitating their transport to areas of need [147]. To prevent their intracellular accumulation, HMA2 in A. thaliana expels excess cytosolic Zn, Cd, and other ions into the apoplast [148]. In response to Zn deficiency in the shoot, HMA2 and metal transporter protein 2 (MTP2) are stimulated to redistribute Zn from the root to the shoot [149]. Additionally, when cultivated in environments with high concentrations of metals, the HMA2 mutant exhibits raised intracellular Cd and Zn levels.
AtHMA3, localized in tonoplasts, plays a role in storing Cd, Zn, Co, and Pb. According to Morel et al. 2009 [150], overexpression of AtHMA3 enhances resistance to metal stress and boosts Cd storage capacity. Furthermore, research by Sasaki et al. 2014 [151] observed that vacuolar sequestration of Cd in rice roots contributes to its accumulation. In a comparable way, GmHMA3 in soybeans is essential for Zn and Cd tolerance. Moreover, a study on rice showed how it has the capability to diminish the quantity of Cd present in its grains by adjusting the Zn-to-Cd ratio via the mutation of OsHMA2 [152].
Additionally, the endomembrane system includes the endoplasmic reticulum (ER), housing a spacious lumen utilized for the storage and utilization of metals for various purposes [153]. Within the ER membrane, there exist broad-specificity transporters for Cd, Cu, and Zn, which are occasionally observed to also localize on the plasma membrane. The endomembrane system’s continuity as an element of the secretory system may be the cause of this shared location. For example, OsZIP1 in rice is situated at both the ER and plasma membrane, primarily in roots where Cd, Cu, and zinc levels are elevated. All the different aforementioned transporters are summarized in Table 1.

6. Conclusions

HMs pose significant risks to plant health, causing toxicity, growth inhibition, and oxidative stress. Essential metals are necessary but can be harmful in excess, while non-essential metals like Cd disrupt critical functions. Plants employ various defense mechanisms, including the synthesis of PCn from GSH via PCS, which chelates Cd and facilitates its detoxification by sequestering it into vacuoles. Due to the implication of GSH in mitigating Cd stress, this review highlights the GSH role in response to abiotic stress.
Moreover, plant cells employ a variety of transporters to preserve metal homeostasis. Different transporters are used by plants to sequester metal ions inside organelles, whether in a free ionic state or as chelates. Transporters belonging to the IRT, NRAMP, and HMA subfamily play crucial roles in Cd absorption, translocation, and sequestration across various plant species; ATPases and ABC transporters facilitate the removal of metals from root cells, ensuring metal stress tolerance. Furthermore, the endomembrane system, particularly the ER, houses transporters for Cd, Cu, and Zn, contributing to metal storage and utilization within plant cells.
The limitations of this study include detailing only a few of the methods plants use for detoxification. A more comprehensive study would explore further mechanisms that could act additively or synergistically. It could distinguish mechanisms of avoidance, where the plant limits the entry of metals, such as the release of radical exudates or wall lignification, and mechanisms for damage repair [14,154], in addition to those that have been mentioned in greater depth.
However, this review highlights the importance of understanding the sources and pathways of Cd contamination in soil, emphasizing the need for comprehensive strategies to mitigate HM pollution. By elucidating the physiological, molecular, and biochemical mechanisms underlying plant responses to Cd stress, this work provides a basis for the development of effective strategies to safeguard plant health and ensure ecosystem resilience in the face of HM pollution.

Author Contributions

Writing, original draft preparation and work conceptualization, V.V.; critical revision of the work, V.V., A.G., A.B., A.S. and A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data provided in the current review are available from the corresponding author on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Jiang, Y.; Jiang, S.; Li, Z.; Yan, X.; Qin, Z.; Huang, R. Field Scale Remediation of Cd and Pb Contaminated Paddy Soil Using Three Mulberry (Morus alba L.) Cultivars. Ecol. Eng. 2019, 129, 38–44. [Google Scholar] [CrossRef]
  2. Bhattacharya, S. Protective Role of the Essential Trace Elements in the Obviation of Cadmium Toxicity: Glimpses of Mechanisms. Biol. Trace Elem. Res. 2022, 200, 2239–2246. [Google Scholar] [CrossRef]
  3. Yadav, M.; Gupta, R.; Sharma, R.K. Chapter 14-Green and Sustainable Pathways for Wastewater Purification. In Advances in Water Purification Techniques; Ahuja, S.B.T.-A., Ed.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 355–383. ISBN 978-0-12-814790-0. [Google Scholar]
  4. Lin, X.; Li, S.; Wei, Z.; Chen, Y.; Hei, L.; Wu, Q.-T. Indirect Application of Sludge for Recycling in Agriculture to Minimize Heavy Metal Contamination of Soil. Resour. Conserv. Recycl. 2021, 166, 105358. [Google Scholar] [CrossRef]
  5. Qin, X.; **a, Y.; Hu, C.; Yu, M.; Shabala, S.; Wu, S.; Tan, Q.; Xu, S.; Sun, X. Ionomics Analysis Provides New Insights into the Co-Enrichment of Cadmium and Zinc in Wheat Grains. Ecotoxicol. Environ. Saf. 2021, 223, 112623. [Google Scholar] [CrossRef]
  6. Gu, Y.-G.; Li, Q.; Fang, J.; He, B.; Fu, H.; Tong, Z. Identification of Heavy Metal Sources in the Reclaimed Farmland Soils of the Pearl River Estuary in China Using a Multivariate Geostatistical Approach. Ecotoxicol. Environ. Saf. 2014, 105, 7–12. [Google Scholar] [CrossRef]
  7. Ministry of Ecology and Environment of the People’s Republic of China. 2021 China Ecology and Environment Statement; Ministry of Ecology and Environment of the People’s Republic of China: Bei**g, China, 2022. [Google Scholar]
  8. Kim, E.; Hopke, P.; Edgerton, E. Improving Source Identification of Atlanta Aerosol Using Temperature Resolved Carbon Fractions in Positive Matrix Factorization. Atmos. Environ. 2004, 38, 3349–3362. [Google Scholar] [CrossRef]
  9. Fei, X.; Lou, Z.; **ao, R.; Ren, Z.; Lv, X. Source Analysis and Source-Oriented Risk Assessment of Heavy Metal Pollution in Agricultural Soils of Different Cultivated Land Qualities. J. Clean. Prod. 2022, 341, 130942. [Google Scholar] [CrossRef]
  10. **ao, R.; Ali, A.; Wang, P.; Li, R.; Tian, X.; Zhang, Z. Comparison of the Feasibility of Different Washing Solutions for Combined Soil Washing and Phytoremediation for the Detoxification of Cadmium (Cd) and Zinc (Zn) in Contaminated Soil. Chemosphere 2019, 230, 510–518. [Google Scholar] [CrossRef] [PubMed]
  11. Ekrami, E.; Pouresmaieli, M.; Hashemiyoon, E.; Noorbakhsh, N.; Mahmoudifard, M. Nanotechnology: A Sustainable Solution for Heavy Metals Remediation. Environ. Nanotechnol. Monit. Manag. 2022, 18, 100718. [Google Scholar] [CrossRef]
  12. Chen, L.; Beiyuan, J.; Hu, W.; Zhang, Z.; Duan, C.; Cui, Q.; Zhu, X.; He, H.; Huang, X. Phytoremediation of Potentially Toxic Elements (PTEs) Contaminated Soils Using Alfalfa (Medicago sativa L.): A Comprehensive Review. Chemosphere 2022, 293, 133577. [Google Scholar] [CrossRef]
  13. Chaoua, S.; Boussaa, S.; Gharmali, A.; Boumezzough, A. Impact of Irrigation with Wastewater on Accumulation of Heavy Metals in Soil and Crops in the Region of Marrakech in Morocco. J. Saudi Soc. Agric. Sci. 2018, 18, 429–436. [Google Scholar] [CrossRef]
  14. Sanita’ di Toppi, L.; Gabbrielli, R. Response to Cadmium in Higher Plants. Environ. Exp. Bot. 1999, 41, 105–130. [Google Scholar] [CrossRef]
  15. Ismael, M.A.; Elyamine, A.M.; Moussa, M.G.; Cai, M.; Zhao, X.; Hu, C. Cadmium in Plants: Uptake, Toxicity, and Its Interactions with Selenium Fertilizers. Metallomics 2019, 11, 255–277. [Google Scholar] [CrossRef] [PubMed]
  16. Kubier, A.; Wilkin, R.T.; Pichler, T. Cadmium in Soils and Groundwater: A Review. Appl. Geochem. J. Int. Assoc. Geochemistry Cosmochem. 2019, 108, 104388. [Google Scholar] [CrossRef] [PubMed]
  17. Clemens, S.; Aarts, M.; Thomine, S.; Verbruggen, N. Plant Science: The Key to Preventing Slow Cadmium Poisoning. Trends Plant Sci. 2012, 18, 92. [Google Scholar] [CrossRef] [PubMed]
  18. Stankovic, J.; Sabovljevic, A.; Sabovljevic, M. Bryophytes and Heavy Metals: A Review. Acta Bot. Croat. 2018, 77, 109–118. [Google Scholar] [CrossRef]
  19. Zhao, H.; Li, P.; He, X. Remediation of Cadmium Contaminated Soil by Modified Gangue Material: Characterization, Performance and Mechanisms. Chemosphere 2021, 290, 133347. [Google Scholar] [CrossRef]
  20. Alimba, C.G.; Gandhi, D.; Sivanesan, S.; Bhanarkar, M.D.; Naoghare, P.K.; Bakare, A.A.; Krishnamurthi, K. Chemical Characterization of Simulated Landfill Soil Leachates from Nigeria and India and Their Cytotoxicity and DNA Damage Inductions on Three Human Cell Lines. Chemosphere 2016, 164, 469–479. [Google Scholar] [CrossRef]
  21. Wang, R.; Lin, K.; Chen, H.; Qi, Z.; Liu, B.; Cao, F.; Chen, H.; Wu, F. Metabolome Analysis Revealed the Mechanism of Exogenous Glutathione to Alleviate Cadmium Stress in Maize (Zea mays L.) Seedlings. Plants 2021, 10, 105. [Google Scholar] [CrossRef]
  22. Wang, Y.; Hu, Y.; Xue, C.; Khan, A.; Zheng, X.; Cai, L. Risk Assessment of Lead and Cadmium Leaching from Solidified/Stabilized MSWI Fly Ash under Long-Term Landfill Simulation Test. Sci. Total Environ. 2022, 816, 151555. [Google Scholar] [CrossRef]
  23. Hölzle, I.; Somani, M.; Ramana, G.V.; Datta, M. Heavy Metals in Soil-like Material from Landfills–Resource or Contaminants? J. Clean. Prod. 2022, 369, 133136. [Google Scholar] [CrossRef]
  24. Alloway, B. Heavy Metals in Soils: Trace Metals and Metalloids in Soils and Their Bioavailability; Springer: Berlin/Heidelberg, Germany, 2013; ISBN 978-94-007-4469-1. [Google Scholar]
  25. Naseem, S.; Hamza; Nawaz-Ul-Huda, S.; Erum, B.; ul-Haq, Q. Geochemistry of Cd in Groundwater of Winder, Balochistan and Suspected Health Problems. Environ. Earth Sci. 2013, 71, 1683–1690. [Google Scholar] [CrossRef]
  26. Alengebawy, A.; Abdelkhalek, S.; Qureshi, S.; Wang, M.-Q. Heavy Metals and Pesticides Toxicity in Agricultural Soil and Plants: Ecological Risks and Human Health Implications. Toxics 2021, 9, 42. [Google Scholar] [CrossRef]
  27. Rehman, A.; Nazir, S.; Irshad, R.; Tahir, K.; Rehman, K.; Islam, R.; Wahab, Z. Toxicity of Heavy Metals in Plants and Animals and Their Uptake by Magnetic Iron Oxide Nanoparticles. J. Mol. Liq. 2020, 321, 114455. [Google Scholar] [CrossRef]
  28. Aoshima, K. Itai-Itai Disease: Renal Tubular Osteomalacia Induced by Environmental Exposure to Cadmium—Historical Review and Perspectives. Soil Sci. Plant Nutr. 2016, 62, 319–326. [Google Scholar] [CrossRef]
  29. Arain, M.B.; Kazi, T.G.; Baig, J.A.; Afridi, H.I.; Sarajuddin; Brehman, K.D.; Panhwar, H.; Arain, S.S. Co-Exposure of Arsenic and Cadmium through Drinking Water and Tobacco Smoking: Risk Assessment on Kidney Dysfunction. Environ. Sci. Pollut. Res. Int. 2015, 22, 350–357. [Google Scholar] [CrossRef] [PubMed]
  30. Khan, M.A.; Khan, S.; Khan, A.; Alam, M. Soil Contamination with Cadmium, Consequences and Remediation Using Organic Amendments. Sci. Total Environ. 2017, 601–602, 1591–1605. [Google Scholar] [CrossRef] [PubMed]
  31. Meysam Hoseini, S.; Zargari, F. Cadmium in Plants: A Review. Int. J. Farming Allied Sci. 2013, 2, 579–581. [Google Scholar]
  32. Li, Y.; Sun, M.; He, W.; Wang, H.; Pan, H.; Yang, Q.; Lou, Y.; Zhuge, Y. Effect of Phosphorus Supplementation on Growth, Nutrient Uptake, Physiological Responses, and Cadmium Absorption by Tall Fescue (Festuca arundinacea Schreb.) Exposed to Cadmium. Ecotoxicol. Environ. Saf. 2021, 213, 112021. [Google Scholar] [CrossRef]
  33. Meng, Y.; **g, H.; Huang, J.; Shen, R.; Zhu, X. The Role of Nitric Oxide Signaling in Plant Responses to Cadmium Stress. Int. J. Mol. Sci. 2022, 23, 6901. [Google Scholar] [CrossRef]
  34. Rady, M.; Elrys, A.; Aboelmaati, M.; Desoky, E.-S. Interplaying Roles of Silicon and Proline Effectively Improve Salt and Cadmium Stress Tolerance in Phaseolus Vulgaris Plant. Plant Physiol. Biochem. 2019, 139, 558–568. [Google Scholar] [CrossRef] [PubMed]
  35. Semida, W.; Hemida, K.; Rady, M. Sequenced Ascorbate-Proline-Glutathione Seed Treatment Elevates Cadmium Tolerance in Cucumber Transplants. Ecotoxicol. Environ. Saf. 2018, 154, 171–179. [Google Scholar] [CrossRef] [PubMed]
  36. Zouari, M.; Ben Ahmed, C.; Zorrig, W.; Elloumi, N.; Rabhi, M.; Delmail, D.; Ben Rouina, B.; Labrousse, P.; Ben Abdallah, F. Exogenous Proline Mediates Alleviation of Cadmium Stress by Promoting Photosynthetic Activity, Water Status and Antioxidative Enzymes Activities of Young Date Palm (Phoenix dactylifera L.). Ecotoxicol. Environ. Saf. 2016, 128, 100–108. [Google Scholar] [CrossRef] [PubMed]
  37. Shahid, M.; Javed, M.T.; Mushtaq, A.; Akram, M.S.; Mahmood, F.; Ahmed, T.; Noman, M.; Azeem, M. Microbe-Mediated Mitigation of Cadmium Toxicity in Plants. In Cadmium Toxicity and Tolerance in Plants; Elsevier: Amsterdam, The Netherlands, 2019; pp. 427–449. ISBN 9780128148648. [Google Scholar]
  38. Sidhu, G.P.S.; Bali, A.; Kumar, V.; Bhardwaj, R. Mitigating Cadmium Toxicity in Plants by Phytohormones. In Cadmium Toxicity and Tolerance in Plants: From Physiology to Remediation; Hasanuzzaman, M., Prasad, M.N.V., Fujita, M., Eds.; Elsevier: London, UK, 2019; pp. 375–396. ISBN 978-0-12-814864-8. [Google Scholar]
  39. Hameed, A.; Rasool, S.; Azooz, M.M.; Hossain, M.A.; Ahanger, M.A.; Ahmad, P. Chapter 24-Heavy Metal Stress: Plant Responses and Signaling. In Plant Metal Interaction; Ahmad, P., Ed.; Elsevier: Amsterdam, The Netherlands, 2016; pp. 557–583. ISBN 978-0-12-803158-2. [Google Scholar]
  40. Huang, Y.; Zhu, Z.; Wu, X.; Liu, Z.; Zou, J.; Chen, Y.; Su, N.; Cui, J. Lower Cadmium Accumulation and Higher Antioxidative Capacity in Edible Parts of Brassica campestris L. Seedlings Applied with Glutathione under Cadmium Toxicity. Environ. Sci. Pollut. Res. 2019, 26, 13235–13245. [Google Scholar] [CrossRef] [PubMed]
  41. Meng, Y.; Zhang, L.; Wang, L.; Zhou, C.; Shangguan, Y.; Yang, Y. Antioxidative Enzymes Activity and Thiol Metabolism in Three Leafy Vegetables under Cd Stress. Ecotoxicol. Environ. Saf. 2019, 173, 214–224. [Google Scholar] [CrossRef] [PubMed]
  42. Shahid, M.; Dumat, C.; Khalid, S.; Schreck, E.; **ong, T.; Niazi, N.K. Foliar Heavy Metal Uptake, Toxicity and Detoxification in Plants: A Comparison of Foliar and Root Metal Uptake. J. Hazard. Mater. 2017, 325, 36–58. [Google Scholar] [CrossRef] [PubMed]
  43. Faizan, M.; Alam, P.; Rajput, V.; Faraz, A.; Afzal, S.; Ahmed, S.; Yu, F.-Y.; Minkina, T.; Hayat, S. Nanoparticle Mediated Plant Tolerance to Heavy Metal Stress: What We Know? Sustainability 2023, 15, 1446. [Google Scholar] [CrossRef]
  44. Valko, M.; Rhodes, C.; Izakovic, M.; Mazur, M. Valko M, Rhodes CJ, Moncol J, Izakovic M, Mazur MFree Radicals, Metals and Antioxidants in Oxidative Stress-Induced Cancer. Chem. Biol. Interact. 2006, 160, 1–40. [Google Scholar] [CrossRef]
  45. Boonstra, J.; Post, J. Molecular Events Associated with Reactive Oxygen Species and Cell Cycle Progression in Mammalian Cells. Gene 2004, 337, 1–13. [Google Scholar] [CrossRef]
  46. Kumar, A.; Majeti, P. Plant-Lead Interactions: Transport, Toxicity, Tolerance, and Detoxification Mechanisms. Ecotoxicol. Environ. Saf. 2018, 166, 401–418. [Google Scholar] [CrossRef]
  47. Liu, Y.; Liu, L.; Qi, J.; Dang, P.; **a, T. Cadmium Activates ZmMPK3-1 and ZmMPK6-1 via Induction of Reactive Oxygen Species in Maize Roots. Biochem. Biophys. Res. Commun. 2019, 516, 747–752. [Google Scholar] [CrossRef]
  48. Berni, R.; Luyckx, M.; Xu, X.; Legay, S.; Sergeant, K.; Hausman, J.-F.; Lutts, S.; Cai, G.; Guerriero, G. Reactive Oxygen Species and Heavy Metal Stress in Plants: Impact on the Cell Wall and Secondary Metabolism. Environ. Exp. Bot. 2019, 161, 98–106. [Google Scholar] [CrossRef]
  49. Ma, Q.-J.; Sun, M.-H.; Lu, J.; Hu, D.-G.; Kang, H.; You, C.-X.; Hao, Y.-J. Phosphorylation of a Malate Transporter Promotes Malate Excretion and Reduces Cadmium Uptake in Apple. J. Exp. Bot. 2020, 71, 3437–3449. [Google Scholar] [CrossRef]
  50. Sharma, S.K.; Singh, D.; Pandey, H.; Jatav, R.B.; Singh, V.; Pandey, D. An Overview of Roles of Enzymatic and Nonenzymatic Antioxidants in Plant BT-Antioxidant Defense in Plants: Molecular Basis of Regulation; Aftab, T., Hakeem, K.R., Eds.; Springer Nature Singapore: Singapore, 2022; pp. 1–13. ISBN 978-981-16-7981-0. [Google Scholar]
  51. Metwally, A.; Safronova, V.I.; Belimov, A.A.; Dietz, K.-J. Genotypic Variation of the Response to Cadmium Toxicity in Pisum sativum L. J. Exp. Bot. 2005, 56, 167–178. [Google Scholar] [CrossRef] [PubMed]
  52. Nada, E.; Ferjani, B.A.; Ali, R.; Bechir, B.R.; Imed, M.; Makki, B. Cadmium-Induced Growth Inhibition and Alteration of Biochemical Parameters in Almond Seedlings Grown in Solution Culture. Acta Physiol. Plant. 2007, 29, 57–62. [Google Scholar] [CrossRef]
  53. Fan, W.; Liu, C.; Cao, B.; Qin, M.; Long, D.; **ang, Z.; Zhao, A. Genome-Wide Identification and Characterization of Four Gene Families Putatively Involved in Cadmium Uptake, Translocation and Sequestration in Mulberry. Front. Plant Sci. 2018, 9, 879. [Google Scholar] [CrossRef] [PubMed]
  54. Jia, H.; Wang, X.; Wei, T.; Zhou, R.; Muhammad, H.; Hua, L.; Ren, X.; Guo, J.; Ding, Y. Accumulation and Fixation of Cd by Tomato Cell Wall Pectin under Cd Stress. Environ. Exp. Bot. 2019, 167, 103829. [Google Scholar] [CrossRef]
  55. Song, J.; Feng, S.J.; Chen, J.; Zhao, W.T.; Yang, Z.M. A Cadmium Stress-Responsive Gene AtFC1 Confers Plant Tolerance to Cadmium Toxicity. BMC Plant Biol. 2017, 17, 187. [Google Scholar] [CrossRef]
  56. Astolfi, S.; Ortolani, M.R.; Catarcione, G.; Paolacci, A.R.; Cesco, S.; Pinton, R.; Ciaffi, M. Cadmium Exposure Affects Iron Acquisition in Barley (Hordeum vulgare) Seedlings. Physiol. Plant. 2014, 152, 646–659. [Google Scholar] [CrossRef]
  57. Bahmani, R.; Modareszadeh, M.; Kim, D.; Hwang, S. Overexpression of Tobacco UBQ2 Increases Cd Tolerance by Decreasing Cd Accumulation and Oxidative Stress in Tobacco and Arabidopsis. Environ. Exp. Bot. 2019, 166, 103805. [Google Scholar] [CrossRef]
  58. Yu, R.; Li, D.; Du, X.; **a, S.; Liu, C.; Shi, G. Comparative Transcriptome Analysis Reveals Key Cadmium Transport-Related Genes in Roots of Two Pak Choi (Brassica rapa L. ssp. chinensis) Cultivars. BMC Genom. 2017, 18, 587. [Google Scholar] [CrossRef] [PubMed]
  59. Rui, H.; Zhang, X.; Shinwari, K.; Zheng, L.; Shen, Z.G. Comparative Transcriptomic Analysis of Two Vicia sativa L. Varieties with Contrasting Responses to Cadmium Stress Reveals the Important Role of Metal Transporters in Cadmium Tolerance. Plant Soil 2018, 423, 241–255. [Google Scholar] [CrossRef]
  60. Shu, X.; Zhang, K.; Zhang, Q.; Wang, W. Ecophysiological Responses of Jatropha curcas L. Seedlings to Simulated Acid Rain under Different Soil Types. Ecotoxicol. Environ. Saf. 2019, 185, 109705. [Google Scholar] [CrossRef] [PubMed]
  61. Milner, M.; Mitani-Ueno, N.; Yamaji, N.; Yokosho, K.; Craft, E.; Fei, Z.; Ebbs, S.; Zambrano, M.; Kochian, L. Root and Shoot Transcriptome Analysis of Two Ecotypes of Noccaea Caerulescens Uncovers the Role of NcNramp1 in Cd Hyperaccumulation. Plant J. 2014, 78, 398–410. [Google Scholar] [CrossRef] [PubMed]
  62. Zhang, J.; Zhang, C.; Huang, S.; Chang, L.; Li, J.; Tang, H.; Dey, S.; Biswas, A.; Du, D.; Li, D.; et al. Key Cannabis Salt-Responsive Genes and Pathways Revealed by Comparative Transcriptome and Physiological Analyses of Contrasting Varieties. Agronomy 2021, 11, 2338. [Google Scholar] [CrossRef]
  63. Feng, Z.; Bartholomew, E.S.; Liu, Z.; Cui, Y.; Dong, Y.; Li, S.; Wu, H.; Ren, H.; Liu, X. Glandular Trichomes: New Focus on Horticultural Crops. Hortic. Res. 2021, 8, 158. [Google Scholar] [CrossRef] [PubMed]
  64. Chen, H.; Li, Y.; Ma, X.; Guo, L.; He, Y.; Ren, Z.; Kuang, Z.; Zhang, X.; Zhang, Z. Analysis of Potential Strategies for Cadmium Stress Tolerance Revealed by Transcriptome Analysis of Upland Cotton. Sci. Rep. 2019, 9, 86. [Google Scholar] [CrossRef] [PubMed]
  65. Wei, W.; Peng, H.; **e, Y.; Wang, X.; Huang, R.; Chen, H.; Ji, X. The Role of Silicon in Cadmium Alleviation by Rice Root Cell Wall Retention and Vacuole Compartmentalization under Different Durations of Cd Exposure. Ecotoxicol. Environ. Saf. 2021, 226, 112810. [Google Scholar] [CrossRef] [PubMed]
  66. Ghuge, S.A.; Nikalje, G.C.; Kadam, U.S.; Suprasanna, P.; Hong, J.C. Comprehensive Mechanisms of Heavy Metal Toxicity in Plants, Detoxification, and Remediation. J. Hazard. Mater. 2023, 450, 131039. [Google Scholar] [CrossRef]
  67. Ahmad, J.; Ali, A.A.; Baig, M.A.; Iqbal, M.; Haq, I.; Irfan Qureshi, M. Chapter 8-Role of Phytochelatins in Cadmium Stress Tolerance in Plants. In Cadmium Toxicity and Tolerance in Plants; Hasanuzzaman, M., Prasad, M.N.V., Fujita, M., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 185–212. ISBN 978-0-12-814864-8. [Google Scholar]
  68. Hendrix, S.; Schröder, P.; Keunen, E.; Huber, C.; Cuypers, A. Chapter Six-Molecular and Cellular Aspects of Contaminant Toxicity in Plants: The Importance of Sulphur and Associated Signalling Pathways. In Phytoremediation; Cuypers, A., Vangronsveld, J., Eds.; Academic Press: Cambridge, MA, USA, 2017; Volume 83, pp. 223–276. ISBN 0065-2296. [Google Scholar]
  69. Al Mahmud, J.; Hasanuzzaman, M.; Nahar, K.; Bhuyan, M.H.M.B.; Fujita, M. Insights into Citric Acid-Induced Cadmium Tolerance and Phytoremediation in Brassica juncea L.: Coordinated Functions of Metal Chelation, Antioxidant Defense and Glyoxalase Systems. Ecotoxicol. Environ. Saf. 2018, 147, 990–1001. [Google Scholar] [CrossRef]
  70. Peco, J.D.; Campos, J.A.; Romero-Puertas, M.C.; Olmedilla, A.; Higueras, P.; Sandalio, L.M. Characterization of Mechanisms Involved in Tolerance and Accumulation of Cd in Biscutella auriculata L. Ecotoxicol. Environ. Saf. 2020, 201, 110784. [Google Scholar] [CrossRef] [PubMed]
  71. Clemens, S.; Peršoh, D. Multi-Tasking Phytochelatin Synthases. Plant Sci. 2009, 177, 266–271. [Google Scholar] [CrossRef]
  72. Bellini, E.; Bandoni, E.; Giardini, S.; Sorce, C.; Spanò, C.; Bottega, S.; Fontanini, D.; Kola, A.; Valensin, D.; Bertolini, A.; et al. Glutathione and Phytochelatins Jointly Allow Intracellular and Extracellular Detoxification of Cadmium in the Liverwort Marchantia Polymorpha. Environ. Exp. Bot. 2023, 209, 105303. [Google Scholar] [CrossRef]
  73. Huang, L.; Li, W.C.; Tam, N.F.Y.; Ye, Z. Effects of Root Morphology and Anatomy on Cadmium Uptake and Translocation in Rice (Oryza sativa L.). J. Environ. Sci. 2019, 75, 296–306. [Google Scholar] [CrossRef] [PubMed]
  74. Zhao, F.-J.; Wang, P. Arsenic and Cadmium Accumulation in Rice and Mitigation Strategies. Plant Soil 2020, 446, 1–21. [Google Scholar] [CrossRef]
  75. Bellini, E.; Varotto, C.; Borsò, M.; Rugnini, L.; Bruno, L.; Di Toppi, L.S. Eukaryotic and Prokaryotic Phytochelatin Synthases Differ Less in Functional Terms than Previously Thought: A Comparative Analysis of Marchantia Polymorpha and Geitlerinema Sp. PCC 7407. Plants 2020, 9, 914. [Google Scholar] [CrossRef] [PubMed]
  76. Li, M.; Barbaro, E.; Bellini, E.; Saba, A.; di Toppi, L.S.; Varotto, C. Ancestral Function of the Phytochelatin Synthase C-Terminal Domain in Inhibition of Heavy Metal-Mediated Enzyme Overactivation. J. Exp. Bot. 2020, 71, 6655–6669. [Google Scholar] [CrossRef] [PubMed]
  77. Li, M.; Leso, M.; Buti, M.; Bellini, E.; Bertoldi, D.; Saba, A.; Larcher, R.; Sanità di Toppi, L.; Varotto, C. Phytochelatin Synthase De-Regulation in Marchantia Polymorpha Indicates Cadmium Detoxification as Its Primary Ancestral Function in Land Plants and Provides a Novel Visual Bioindicator for Detection of This Metal. J. Hazard. Mater. 2022, 440, 129844. [Google Scholar] [CrossRef]
  78. Bhuyan, M.H.M.B.; Parvin, K.; Mohsin, S.M.; Mahmud, J.A.; Hasanuzzaman, M.; Fujita, M. Modulation of Cadmium Tolerance in Rice: Insight into Vanillic Acid-Induced Upregulation of Antioxidant Defense and Glyoxalase Systems. Plants 2020, 9, 188. [Google Scholar] [CrossRef]
  79. Degola, F.; De Benedictis, M.; Petraglia, A.; Massimi, A.; Fattorini, L.; Sorbo, S.; Basile, A.; Sanita’ di Toppi, L. A Cd/Fe/Zn-Responsive Phytochelatin Synthase Is Constitutively Present in the Ancient Liverwort Lunularia cruciata (L.) Dumort. Plant Cell Physiol. 2014, 55, 1884–1891. [Google Scholar] [CrossRef]
  80. Petraglia, A.; De Benedictis, M.; Degola, F.; Pastore, G.; Calcagno, M.; Ruotolo, R.; Mengoni, A.; Sanita’ di Toppi, L. The Capability to Synthesize Phytochelatins and the Presence of Constitutive and Functional Phytochelatin Synthases Are Ancestral (Plesiomorphic) Characters for Basal Land Plants. J. Exp. Bot. 2014, 65, 1153–1163. [Google Scholar] [CrossRef] [PubMed]
  81. Fontanini, D.; Andreucci, A.; Castiglione, M.; Basile, A.; Sorbo, S.; Petraglia, A.; Degola, F.; Bellini, E.; Bruno, L.; Varotto, C.; et al. The Phytochelatin Synthase from Nitella Mucronata (Charophyta) Plays a Role in the Homeostatic Control of Iron(II)/(III). Plant Physiol. Biochem. 2018, 127, 88–96. [Google Scholar] [CrossRef] [PubMed]
  82. Bellini, E.; Betti, C.; Sanità di Toppi, L. Responses to Cadmium in Early-Diverging Streptophytes (Charophytes and Bryophytes): Current Views and Potential Applications. Plants 2021, 10, 770. [Google Scholar] [CrossRef] [PubMed]
  83. Maresca, V.; Sorbo, S.; Loppi, S.; Funaro, F.; Del Prete, D.; Basile, A. Biological Effects from Environmental Pollution by Toxic Metals in the “Land of Fires” (Italy) Assessed Using the Biomonitor Species Lunularia cruciata L. (Dum). Environ. Pollut. 2020, 265, 115000. [Google Scholar] [CrossRef] [PubMed]
  84. Fasani, E.; Li, M.; Varotto, C.; Furini, A.; DalCorso, G. Metal Detoxification in Land Plants: From Bryophytes to Vascular Plants. STATE of the Art and Opportunities. Plants 2022, 11, 237. [Google Scholar] [CrossRef] [PubMed]
  85. Bellini, E.; Maresca, V.; Betti, C.; Castiglione, M.R.; Fontanini, D.; Capocchi, A.; Sorce, C.; Borsò, M.; Bruno, L.; Sorbo, S.; et al. The Moss Leptodictyum riparium Counteracts Severe Cadmium Stress by Activation of Glutathione Transferase and Phytochelatin Synthase, but Slightly by Phytochelatins. Int. J. Mol. Sci. 2020, 21, 1583. [Google Scholar] [CrossRef] [PubMed]
  86. Esposito, S.; Loppi, S.; Monaci, F.; Paoli, L.; Vannini, A.; Sorbo, S.; Maresca, V.; Fusaro, L.; Karam, E.; Lentini, M.; et al. In-Field and in-Vitro Study of the Moss Leptodictyum riparium as Bioindicator of Toxic Metal Pollution in the Aquatic Environment: Ultrastructural Damage, Oxidative Stress and HSP70 Induction. PLoS ONE 2018, 13, e0195717. [Google Scholar] [CrossRef] [PubMed]
  87. Landi, S.; Esposito, S. Nitrate Uptake Affects Cell Wall Synthesis and Modeling. Front. Plant Sci. 2017, 8, 1376. [Google Scholar] [CrossRef] [PubMed]
  88. Singh, S.; Prasad, S.M. Effects of 28-Homobrassinoloid on Key Physiological Attributes of Solanum Lycopersicum Seedlings under Cadmium Stress: Photosynthesis and Nitrogen Metabolism. Plant Growth Regul. 2017, 82, 161–173. [Google Scholar] [CrossRef]
  89. Lentini, M.; De Lillo, A.; Paradisone, V.; Liberti, D.; Landi, S.; Esposito, S. Early Responses to Cadmium Exposure in Barley Plants: Effects on Biometric and Physiological Parameters. Acta Physiol. Plant. 2018, 40, 178. [Google Scholar] [CrossRef]
  90. Yang, Y.; **ong, J.; Tao, L.; Cao, Z.; Tang, W.; Zhang, J.; Yu, X.; Fu, G.; Zhang, X.; Lu, Y. Regulatory Mechanisms of Nitrogen (N) on Cadmium (Cd) Uptake and Accumulation in Plants: A Review. Sci. Total Environ. 2020, 708, 135186. [Google Scholar] [CrossRef]
  91. Ben Azaiez, F.E.; Ayadi, S.; Capasso, G.; Landi, S.; Paradisone, V.; Jallouli, S.; Hammami, Z.; Chamekh, Z.; Zouari, I.; Trifa, Y.; et al. Salt Stress Induces Differentiated Nitrogen Uptake and Antioxidant Responses in Two Contrasting Barley Landraces from MENA Region. Agronomy 2020, 10, 1426. [Google Scholar] [CrossRef]
  92. Jallouli, S.; Ayadi, S.; Landi, S.; Capasso, G.; Santini, G.; Chamekh, Z.; Zouari, I.; Ben Azaiez, F.E.; Trifa, Y.; Esposito, S. Physiological and Molecular Osmotic Stress Responses in Three Durum Wheat (Triticum Turgidum ssp Durum) Genotypes. Agronomy 2019, 9, 550. [Google Scholar] [CrossRef]
  93. Lee, H.J.; Abdula, S.E.; Jang, D.W.; Park, S.-H.; Yoon, U.-H.; Jung, Y.J.; Kang, K.K.; Nou, I.S.; Cho, Y.-G. Overexpression of the Glutamine Synthetase Gene Modulates Oxidative Stress Response in Rice after Exposure to Cadmium Stress. Plant Cell Rep. 2013, 32, 1521–1529. [Google Scholar] [CrossRef] [PubMed]
  94. Saini, S.; Kaur, N.; Pati, P.K. Phytohormones: Key Players in the Modulation of Heavy Metal Stress Tolerance in Plants. Ecotoxicol. Environ. Saf. 2021, 223, 112578. [Google Scholar] [CrossRef]
  95. Maresca, V.; Bellini, E.; Landi, S.; Capasso, G.; Cianciullo, P.; Carraturo, F.; Pirintsos, S.; Sorbo, S.; Sanità di Toppi, L.; Esposito, S.; et al. Biological Responses to Heavy Metal Stress in the Moss Leptodictyum riparium (Hedw.) Warnst. Ecotoxicol. Environ. Saf. 2022, 229, 113078. [Google Scholar] [CrossRef]
  96. Meister, A. Glutathione Metabolism. Methods Enzymol. 1995, 251, 3–7. [Google Scholar] [CrossRef] [PubMed]
  97. Pasternak, M.; Lim, B.; Wirtz, M.; Hell, R.; Cobbett, C.S.; Meyer, A.J. Restricting Glutathione Biosynthesis to the Cytosol Is Sufficient for Normal Plant Development. Plant J. 2008, 53, 999–1012. [Google Scholar] [CrossRef] [PubMed]
  98. Belin, C.; Bashandy, T.; Cela, J.; Delorme-Hinoux, V.; Riondet, C.; Reichheld, J.P. A Comprehensive Study of Thiol Reduction Gene Expression under Stress Conditions in Arabidopsis thaliana. Plant. Cell Environ. 2015, 38, 299–314. [Google Scholar] [CrossRef]
  99. **ang, C.; Oliver, D.J. Glutathione Metabolic Genes Coordinately Respond to Heavy Metals and Jasmonic Acid in Arabidopsis. Plant Cell 1998, 10, 1539–1550. [Google Scholar] [CrossRef]
  100. Wachter, A.; Wolf, S.; Steininger, H.; Bogs, J.; Rausch, T. Differential Targeting of GSH1 and GSH2 Is Achieved by Multiple Transcription Initiation: Implications for the Compartmentation of Glutathione Biosynthesis in the Brassicaceae. Plant J. 2005, 41, 15–30. [Google Scholar] [CrossRef] [PubMed]
  101. Yang, Y.; Lenherr, E.D.; Gromes, R.; Wang, S.; Wirtz, M.; Hell, R.; Peskan-Berghöfer, T.; Scheffzek, K.; Rausch, T. Plant Glutathione Biosynthesis Revisited: Redox-Mediated Activation of Glutamylcysteine Ligase Does Not Require Homo-Dimerization. Biochem. J. 2019, 476, 1191–1203. [Google Scholar] [CrossRef] [PubMed]
  102. Watson, S.J.; Sowden, R.G.; Jarvis, P. Abiotic Stress-Induced Chloroplast Proteome Remodelling: A Mechanistic Overview. J. Exp. Bot. 2018, 69, 2773–2781. [Google Scholar] [CrossRef] [PubMed]
  103. Kopriva, S.; Rennenberg, H. Control of Sulphate Assimilation and Glutathione Synthesis: Interaction with N and C Metabolism. J. Exp. Bot. 2004, 55, 1831–1842. [Google Scholar] [CrossRef] [PubMed]
  104. Strohm, M.; Jouanin, L.; Kunert, K.J.; Pruvost, C.; Polle, A.; Foyer, C.H.; Rennenberg, H. Regulation of Glutathione Synthesis in Leaves of Transgenic Poplar (Populus tremula X P. alba) Overexpressing Glutathione Synthetase. Plant J. 1995, 7, 141–145. [Google Scholar] [CrossRef]
  105. Gill, S.S.; Anjum, N.A.; Hasanuzzaman, M.; Gill, R.; Trivedi, D.K.; Ahmad, I.; Pereira, E.; Tuteja, N. Glutathione and Glutathione Reductase: A Boon in Disguise for Plant Abiotic Stress Defense Operations. Plant Physiol. Biochem. 2013, 70, 204–212. [Google Scholar] [CrossRef] [PubMed]
  106. Sairam, R.K.; Saxena, D.C. Oxidative Stress and Antioxidants in Wheat Genotypes: Possible Mechanism of Water Stress Tolerance. J. Agron. Crop Sci. 2000, 184, 55–61. [Google Scholar] [CrossRef]
  107. Romero-Puertas, M.C.; Corpas, F.J.; Sandalio, L.M.; Leterrier, M.; Rodríguez-Serrano, M.; Del Río, L.A.; Palma, J.M. Glutathione Reductase from Pea Leaves: Response to Abiotic Stress and Characterization of the Peroxisomal Isozyme. New Phytol. 2006, 170, 43–52. [Google Scholar] [CrossRef] [PubMed]
  108. Marty, L.; Siala, W.; Schwarzländer, M.; Fricker, M.D.; Wirtz, M.; Sweetlove, L.J.; Meyer, Y.; Meyer, A.J.; Reichheld, J.-P.; Hell, R. The NADPH-Dependent Thioredoxin System Constitutes a Functional Backup for Cytosolic Glutathione Reductase in Arabidopsis. Proc. Natl. Acad. Sci. USA 2009, 106, 9109–9114. [Google Scholar] [CrossRef]
  109. Kataya, A.R.A.; Reumann, S. Arabidopsis Glutathione Reductase 1 Is Dually Targeted to Peroxisomes and the Cytosol. Plant Signal. Behav. 2010, 5, 171–175. [Google Scholar] [CrossRef]
  110. Delorme-Hinoux, V.; Bangash, S.A.K.; Meyer, A.J.; Reichheld, J.-P. Nuclear Thiol Redox Systems in Plants. Plant Sci. 2016, 243, 84–95. [Google Scholar] [CrossRef] [PubMed]
  111. Liu, L.; Li, W.; Song, W.; Guo, M. Remediation Techniques for Heavy Metal-Contaminated Soils: Principles and Applicability. Sci. Total Environ. 2018, 633, 206–219. [Google Scholar] [CrossRef]
  112. Rizwan, M.; Ali, S.; Zia ur Rehman, M.; Rinklebe, J.; Tsang, D.C.W.; Bashir, A.; Maqbool, A.; Tack, F.M.G.; Ok, Y.S. Cadmium Phytoremediation Potential of Brassica Crop Species: A Review. Sci. Total Environ. 2018, 631–632, 1175–1191. [Google Scholar] [CrossRef] [PubMed]
  113. Luo, Z.-B.; He, J.; Polle, A.; Rennenberg, H. Heavy Metal Accumulation and Signal Transduction in Herbaceous and Woody Plants: Paving the Way for Enhancing Phytoremediation Efficiency. Biotechnol. Adv. 2016, 34, 1131–1148. [Google Scholar] [CrossRef] [PubMed]
  114. Krämer, U. Metal Hyperaccumulation in Plants. Annu. Rev. Plant Biol. 2010, 61, 517–534. [Google Scholar] [CrossRef] [PubMed]
  115. Meighan, M.M.; Fenus, T.; Karey, E.; MacNeil, J. The Impact of EDTA on the Rate of Accumulation and Root/Shoot Partitioning of Cadmium in Mature Dwarf Sunflowers. Chemosphere 2011, 83, 1539–1545. [Google Scholar] [CrossRef] [PubMed]
  116. Di Lonardo, S.; Capuana, M.; Arnetoli, M.; Gabbrielli, R.; Gonnelli, C. Exploring the Metal Phytoremediation Potential of Three Populus alba L. Clones Using an in Vitro Screening. Environ. Sci. Pollut. Res. 2011, 18, 82–90. [Google Scholar] [CrossRef] [PubMed]
  117. Marmiroli, M.; Visioli, G.; Maestri, E.; Marmiroli, N. Correlating SNP Genotype with the Phenotypic Response to Exposure to Cadmium in Populus spp. Environ. Sci. Technol. 2011, 45, 4497–4505. [Google Scholar] [CrossRef] [PubMed]
  118. Pietrini, F.; Zacchini, M.; Iori, V.; Pietrosanti, L.; Bianconi, D.; Massacci, A. Screening of Poplar Clones for Cadmium Phytoremediation Using Photosynthesis, Biomass and Cadmium Content Analyses. Int. J. Phytoremediation 2010, 12, 105–120. [Google Scholar] [CrossRef]
  119. Wu, F.; Yang, W.; Zhang, J.; Zhou, L. Cadmium Accumulation and Growth Responses of a Poplar (Populus deltoids x Populus nigra) in Cadmium Contaminated Purple Soil and Alluvial Soil. J. Hazard. Mater. 2010, 177, 268–273. [Google Scholar] [CrossRef]
  120. Redjala, T.; Zelko, I.; Sterckeman, T.; Legué, V.; Lux, A. Relationship between Root Structure and Root Cadmium Uptake in Maize. Environ. Exp. Bot. 2011, 71, 241–248. [Google Scholar] [CrossRef]
  121. Thakur, S.; Singh, L.; Wahid, Z.; Siddiqui, M.; Atnaw, S.; Md Din, M.F. Plant-Driven Removal of Heavy Metals from Soil: Uptake, Translocation, Tolerance Mechanism, Challenges, and Future Perspectives. Environ. Monit. Assess. 2016, 188, 206. [Google Scholar] [CrossRef] [PubMed]
  122. Kumar Yadav, K.; Gupta, N.; Kumar, A.; Reece, L.M.; Singh, N.; Rezania, S.; Ahmad Khan, S. Mechanistic Understanding and Holistic Approach of Phytoremediation: A Review on Application and Future Prospects. Ecol. Eng. 2018, 120, 274–298. [Google Scholar] [CrossRef]
  123. Laghlimi, M.; Baghdad, B.; EL Hadi, H.; Bouabdli, A. Phytoremediation Mechanisms of Heavy Metal Contaminated Soils: A Review. Open J. Ecol. 2015, 5, 375–388. [Google Scholar] [CrossRef]
  124. Fahr, M.; Laplaze, L.; Bendaou, N.; Hocher, V.; EL Mzibri, M.; Bogusz, D.; Smouni, A. Effect of Lead on Root Growth. Front. Plant Sci. 2013, 4, 175. [Google Scholar] [CrossRef] [PubMed]
  125. Gupta, N.; Yadav, K.K.; Kumar, V.; Kumar, S.; Chadd, R.P.; Kumar, A. Trace Elements in Soil-Vegetables Interface: Translocation, Bioaccumulation, Toxicity and Amelioration–A Review. Sci. Total Environ. 2019, 651, 2927–2942. [Google Scholar] [CrossRef] [PubMed]
  126. Gupta, N.; Ram, H.; Kumar, B. Mechanism of Zinc Absorption in Plants: Uptake, Transport, Translocation and Accumulation. Rev. Environ. Sci. Bio/Technol. 2016, 15, 89–109. [Google Scholar] [CrossRef]
  127. Kim, R.-Y.; Yoon, J.-K.; Kim, T.-S.; Yang, J.; Owens, G.; Kim, K.-R. Bioavailability of Heavy Metals in Soils: Definitions and Practical Implementation—A Critical Review. Environ. Geochem. Health 2015, 37, 1041–1061. [Google Scholar] [CrossRef] [PubMed]
  128. Sebastian, A.; Prasad, M.N. V Exogenous Citrate and Malate Alleviate Cadmium Stress in Oryza Sativa L.: Probing Role of Cadmium Localization and Iron Nutrition. Ecotoxicol. Environ. Saf. 2018, 166, 215–222. [Google Scholar] [CrossRef] [PubMed]
  129. Wu, Z.; Zhao, X.; Sun, X.; Tan, Q.; Tang, Y.; Nie, Z.; Hu, C. Xylem Transport and Gene Expression Play Decisive Roles in Cadmium Accumulation in Shoots of Two Oilseed Rape Cultivars (Brassica napus). Chemosphere 2015, 119, 1217–1223. [Google Scholar] [CrossRef]
  130. Hall, J.L. Cellular Mechanisms for Heavy Metal Detoxification and Tolerance. J. Exp. Bot. 2002, 53, 1–11. [Google Scholar] [CrossRef]
  131. Kumar, A.; Dubey, A.; Kumar, V.; Ansari, M.; Narayan, S.M.; Kumar, S.; Meenakshi; Pandey, V.; Shirke, P.; Pande, V.; et al. Over-Expression of Chickpea Glutaredoxin (CaGrx) Provides Tolerance to Heavy Metals by Reducing Metal Accumulation and Improved Physiological and Antioxidant Defence System. Ecotoxicol. Environ. Saf. 2020, 192, 110252. [Google Scholar] [CrossRef] [PubMed]
  132. Cointry, V.; Vert, G. The Bifunctional Transporter-receptor IRT1 at the Heart of Metal Sensing and Signaling. New Phytol. 2019, 223, 1173–1178. [Google Scholar] [CrossRef] [PubMed]
  133. Narayan, O.; Verma, N.; Jogawat, A.; Dua, M.; Johri, A. Role of Sulphate Transporter (PiSulT) of Endophytic Fungus Serendipita Indica in Plant Growth and Development. bioRxiv 2020. [Google Scholar] [CrossRef]
  134. Narayan, O.; Verma, N.; Jogawat, A.; Dua, M.; Johri, A. Sulfur Transfer from the Endophytic Fungus Serendipita Indica Improves Maize Growth and Requires the Sulfate Transporter SiSulT. Plant Cell 2021, 33, 1268–1285. [Google Scholar] [CrossRef] [PubMed]
  135. Yadav, B.; Jogawat, A.; Lal, S.K.; Lakra, N.; Mehta, S.; Shabek, N.; Narayan, O. Plant Mineral Transport Systems and the Potential for Crop Improvement. Planta 2021, 253, 45. [Google Scholar] [CrossRef] [PubMed]
  136. Dubeaux, G.; Julie, N.; Zelazny, E.; Vert, G. Metal Sensing by the IRT1 Transporter-Receptor Orchestrates Its Own Degradation and Plant Metal Nutrition. Mol. Cell 2018, 69, 953–964.e5. [Google Scholar] [CrossRef] [PubMed]
  137. Wu, Q.; Huang, L.; Su, N.; Shabala, L.; Wang, H.; Huang, X.; Wen, R.; Yu, M.; Cui, J.; Shabala, S. Calcium-Dependent Hydrogen Peroxide Mediates Hydrogen-Rich Water-Reduced Cadmium Uptake in Plant Roots. Plant Physiol. 2020, 183, 1331–1344. [Google Scholar] [CrossRef] [PubMed]
  138. Takahashi, R.; Ishimaru, Y.; Senoura, T.; Shimo, H.; Ishikawa, S.; Arao, T.; Nakanishi, H.; Nishizawa, N. The OsNRAMP1 Iron Transporter Is Involved in Cd Accumulation in Rice. J. Exp. Bot. 2011, 62, 4843–4850. [Google Scholar] [CrossRef]
  139. Lang, M.; Hao, M.; Fan, Q.; Wang, W.; Mo, S.; Zhao, W.; Zhou, J. Functional Characterization of BjCET3 and BjCET4, Two New Cation-Efflux Transporters from Brassica juncea L. J. Exp. Bot. 2011, 62, 4467–4480. [Google Scholar] [CrossRef]
  140. Li, J.; Wang, Y.; Zheng, L.; Li, Y.; Zhou, X.; Li, J.; Gu, D.; Xu, E.; Lu, Y.; Chen, X.; et al. The Intracellular Transporter AtNRAMP6 Is Involved in Fe Homeostasis in Arabidopsis. Front. Plant Sci. 2019, 10, 1124. [Google Scholar] [CrossRef] [PubMed]
  141. Fu, S.; Lu, Y.; Zhang, X.; Yang, G.; Chao, D.; Wang, Z.; Shi, M.; Chen, J.; Chao, D.-Y.; Li, R.; et al. The ABC Transporter ABCG36 Is Required for Cadmium Tolerance in Rice. J. Exp. Bot. 2019, 70, 5909–5918. [Google Scholar] [CrossRef] [PubMed]
  142. Wang, H.; Liu, Y.; Peng, Z.; Li, J.; Huang, W.; Liu, Y.; Wang, X.; **e, S.; Sun, L.; Han, E.; et al. Ectopic Expression of Poplar ABC Transporter PtoABCG36 Confers Cd Tolerance in Arabidopsis thaliana. Int. J. Mol. Sci. 2019, 20, 3293. [Google Scholar] [CrossRef] [PubMed]
  143. Kim, D.Y.; Bovet, L.; Maeshima, M.; Martinoia, E.; Lee, Y. The ABC Transporter AtPDR8 Is a Cadmium Extrusion Pump Conferring Heavy Metal Resistance. Plant J. 2007, 50, 207–218. [Google Scholar] [CrossRef] [PubMed]
  144. Brunetti, P.; Zanella, L.; De Paolis, A.; Di Litta, D.; Cecchetti, V.; Falasca, G.; Barbieri, M.; Altamura, M.M.; Costantino, P.; Cardarelli, M. Cadmium-Inducible Expression of the ABC-Type Transporter AtABCC3 Increases Phytochelatin-Mediated Cadmium Tolerance in Arabidopsis. J. Exp. Bot. 2015, 66, 3815–3829. [Google Scholar] [CrossRef]
  145. Song, W.-Y.; Park, J.; Mendoza Cozatl, D.; Grotemeyer, M.; Shim, D.; Hörtensteiner, S.; Geisler, M.; Weder, B.; Rea, P.; Rentsch, D.; et al. Arsenic Tolerance in Arabidopsis Is Mediated by Two ABCC-Type Phytochelatin Transporters. Proc. Natl. Acad. Sci. USA 2010, 107, 21187–21192. [Google Scholar] [CrossRef] [PubMed]
  146. Neri, A.; Traversari, S.; Andreucci, A.; Francini, A.; Sebastiani, L. The Role of Aquaporin Overexpression in the Modulation of Transcription of Heavy Metal Transporters under Cadmium Treatment in Poplar. Plants 2021, 10, 54. [Google Scholar] [CrossRef] [PubMed]
  147. Mikkelsen, M.; Pedas, P.; Stokholm, M.; Vincze, E.; Mills, R.; Borg, S.; Møller, A.; Schjoerring, J.; Williams, L.; Baekgaard, L.; et al. Barley HvHMA1 Is a Heavy Metal Pump Involved in Mobilizing Organellar Zn and Cu and Plays a Role in Metal Loading into Grains. PLoS ONE 2012, 7, e49027. [Google Scholar] [CrossRef]
  148. Eren, E.; Argüello, J. Arabidopsis HMA2, a Divalent Heavy Metal-Transporting PIB-Type ATPase, Is Involved in Cytoplasmic Zn2+ Homeostasis. Plant Physiol. 2004, 136, 3712–3723. [Google Scholar] [CrossRef]
  149. Sinclair, S.; Senger, T.; Talke, I.; Cobbett, C.; Haydon, M.; Kraemer, U. Systemic Upregulation of MTP2- and HMA2-Mediated Zn Partitioning to the Shoot Supplements Local Zn Deficiency Responses of Arabidopsis. Plant Cell 2018, 30, 2463–2479. [Google Scholar] [CrossRef]
  150. Morel-Rouhier, M.; Crouzet, J.; Gravot, A.; Auroy, P.; Leonhardt, N.; Vavasseur, A.; Richaud, P. AtHMA3, a P1B-ATPase Allowing Cd/Zn/Co/Pb Vacuolar Storage in Arabidopsis. Plant Physiol. 2008, 149, 894–904. [Google Scholar] [CrossRef] [PubMed]
  151. Sasaki, A.; Yamaji, N. Overexpression of OsHMA3 Enhances Cd Tolerance and Expression of Zn Transporter Genes in Rice. J. Exp. Bot. 2014, 65, 6013–6021. [Google Scholar] [CrossRef] [PubMed]
  152. Satoh-Nagasawa, N.; Mori, M.; Nakazawa, N.; Kawamoto, T.; Nagato, Y.; Sakurai, K.; Takahashi, H.; Watanabe, A.; Akagi, H. Mutations in Rice (Oryza sativa) Heavy Metal ATPase 2 (OsHMA2) Restrict the Translocation of Zinc and Cadmium. Plant Cell Physiol. 2012, 53, 213–224. [Google Scholar] [CrossRef] [PubMed]
  153. De caroli, M.; Furini, A.; Dalcorso, G.; Rojas, M.; Di Sansebastiano, G. Endomembrane Reorganization Induced by Heavy Metals. Plants 2020, 9, 482. [Google Scholar] [CrossRef]
  154. Sanita’ di Toppi, L.; Bruno, L.; Bruni, R.; Ligrone, R.; Ferrarese, A.; Paoli, L.; Lingua, G. Interazioni Piante-Ambiente; PICCIN: Padova, Italy, 2018; ISBN 978-88-299-2870-5. [Google Scholar]
Figure 1. Toxic effects of Cd in plants. Induction of oxidative stress leads to ROS generation, resulting in cell membrane damage, protein oxidation, and degradation, with effects on gene expression. Cellular responses to Cd entry include activation of the defense system as well as chelation, sequestration, and translocation of the HM.
Figure 1. Toxic effects of Cd in plants. Induction of oxidative stress leads to ROS generation, resulting in cell membrane damage, protein oxidation, and degradation, with effects on gene expression. Cellular responses to Cd entry include activation of the defense system as well as chelation, sequestration, and translocation of the HM.
Cimb 46 00361 g001
Figure 2. The Cd-PCn complex. Cd influx into cells induces overexpression of PCS and involves glutathione in PC2 formation, resulting in the synthesis of other PCn. This production of PCn leads to cadmium chelation, and the complex is transported to the vacuole for detoxification or removed from the cell via membrane transporters and efflux pumps.
Figure 2. The Cd-PCn complex. Cd influx into cells induces overexpression of PCS and involves glutathione in PC2 formation, resulting in the synthesis of other PCn. This production of PCn leads to cadmium chelation, and the complex is transported to the vacuole for detoxification or removed from the cell via membrane transporters and efflux pumps.
Cimb 46 00361 g002
Table 1. Main features of transporters implicated in Cd tolerance.
Table 1. Main features of transporters implicated in Cd tolerance.
SubfamilyNameSubcellular
Localization
Substrate
IRTIRT1Plasma membraneFe, Co, Mn, Zn, Cd
SOSSOS2L1 *Plasma membraneMalate and malic acid
ALMTALMT14 *Plasma membraneMalate and malic acid
CETBjCET3Plasma membraneZn, Co, Ni, Cd
BjCET4
RAMPOsNRAMP1Plasma membraneFe, Cd
AtNRAMP6GolgiFe, Mn, Cd
ABCG (PDR)PtABCG36Plasma membranePb, Cd
AtPDR8
ABCCAtABCC1TonoplastPCn and its complexes with As, Mn, Zn, Cd
AtABCC2
AtABCC3
HMAHvHMA1ChloroplastZn, Cd
PaHMA2Plasma membrane
AtHMA2Zn, Cd
AtHMA3TonoplastZn, Co, Pb, Cd
ZIPOsZIP1Endoplasmic reticulum and plasma membraneZn, Cu, Cd
* Not directly involved in Cd transport. Cd presence enhances their functionality, increasing Cd tolerance.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vitelli, V.; Giamborino, A.; Bertolini, A.; Saba, A.; Andreucci, A. Cadmium Stress Signaling Pathways in Plants: Molecular Responses and Mechanisms. Curr. Issues Mol. Biol. 2024, 46, 6052-6068. https://doi.org/10.3390/cimb46060361

AMA Style

Vitelli V, Giamborino A, Bertolini A, Saba A, Andreucci A. Cadmium Stress Signaling Pathways in Plants: Molecular Responses and Mechanisms. Current Issues in Molecular Biology. 2024; 46(6):6052-6068. https://doi.org/10.3390/cimb46060361

Chicago/Turabian Style

Vitelli, Valentina, Agnese Giamborino, Andrea Bertolini, Alessandro Saba, and Andrea Andreucci. 2024. "Cadmium Stress Signaling Pathways in Plants: Molecular Responses and Mechanisms" Current Issues in Molecular Biology 46, no. 6: 6052-6068. https://doi.org/10.3390/cimb46060361

Article Metrics

Back to TopTop