Heavy metal hyperaccumulating plants - PSGSC

35 downloads 0 Views 322KB Size Report
also metalloids that can be toxic to both plants and animals even at very low ... be deleted from the list if this trait is unconfirmed by experimenta- tion under ...
Plant Science 180 (2011) 169–181

Contents lists available at ScienceDirect

Plant Science journal homepage: www.elsevier.com/locate/plantsci

Review

Heavy metal hyperaccumulating plants: How and why do they do it? And what makes them so interesting? Nicoletta Rascio a,∗ , Flavia Navari-Izzo b a b

Department of Biology, University of Padova, via U. Bassi 58/B, I-35121 Padova, Italy Department of Chemistry and Agricultural Biotechnologies, University of Pisa, via San Michele degli Scalzi 2, I-56124 Pisa, Italy

a r t i c l e

i n f o

Article history: Received 26 May 2010 Received in revised form 25 August 2010 Accepted 26 August 2010 Available online 15 September 2010 Keywords: Heavy metal uptake Heavy metal translocation Heavy metal detoxification/sequestration Hyperaccumulators Phytomining Phytoremediation

a b s t r a c t The term “hyperaccumulator” describes a number of plants that belong to distantly related families, but share the ability to grow on metalliferous soils and to accumulate extraordinarily high amounts of heavy metals in the aerial organs, far in excess of the levels found in the majority of species, without suffering phytotoxic effects. Three basic hallmarks distinguish hyperaccumulators from related non-hyperaccumulating taxa: a strongly enhanced rate of heavy metal uptake, a faster root-to-shoot translocation and a greater ability to detoxify and sequester heavy metals in leaves. An interesting breakthrough that has emerged from comparative physiological and molecular analyses of hyperaccumulators and related non-hyperaccumulators is that most key steps of hyperaccumulation rely on different regulation and expression of genes found in both kinds of plants. In particular, a determinant role in driving the uptake, translocation to leaves and, finally, sequestration in vacuoles or cell walls of great amounts of heavy metals, is played in hyperaccumulators by constitutive overexpression of genes encoding transmembrane transporters, such as members of ZIP, HMA, MATE, YSL and MTP families. Among the hypotheses proposed to explain the function of hyperaccumulation, most evidence has supported the “elemental defence” hypothesis, which states that plants hyperaccumulate heavy metals as a defence mechanism against natural enemies, such as herbivores. According to the more recent hypothesis of “joint effects”, heavy metals can operate in concert with organic defensive compounds leading to enhanced plant defence overall. Heavy metal contaminated soils pose an increasing problem to human and animal health. Using plants that hyperaccumulate specific metals in cleanup efforts appeared over the last 20 years. Metal accumulating species can be used for phytoremediation (removal of contaminant from soils) or phytomining (growing plants to harvest the metals). In addition, as many of the metals that can be hyperaccumulated are also essential nutrients, food fortification and phytoremediation might be considered two sides of the same coin. An overview of literature discussing the phytoremediation capacity of hyperaccumulators to clean up soils contaminated with heavy metals and the possibility of using these plants in phytomining is presented. © 2010 Elsevier Ireland Ltd. All rights reserved.

Contents 1. 2. 3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . What are heavy metal hyperaccumulator plants? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . How do plants hyperaccumulate heavy metals? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Heavy metal uptake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Root-to-shoot translocation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Detoxification/sequestration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Why did plants evolve hyperaccumulation of heavy metals? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. The “elemental defence” hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. The “joint effects” hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

∗ Corresponding author. Tel.: +39 049 8276278; fax: +39 049 8276260. E-mail address: [email protected] (N. Rascio). 0168-9452/$ – see front matter © 2010 Elsevier Ireland Ltd. All rights reserved. doi:10.1016/j.plantsci.2010.08.016

170 170 171 171 172 173 174 174 175

170

5.

6.

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

Why do hyperaccumulators attract so much interest? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. Potential application in phytoremediation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. Transfer of hyperaccumulation traits to rapidly growing species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3. Potential application in phytomining . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusions and future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction From a chemical point of view, the term heavy metal is strictly ascribed to transition metals with atomic mass over 20 and specific gravity above 5. In biology, “heavy” refers to a series of metals and also metalloids that can be toxic to both plants and animals even at very low concentrations. Here the term “heavy metals” will be for these potentially phytotoxic elements. Some of these heavy metals, such as As, Cd, Hg, Pb or Se, are not essential, since they do not perform any known physiological function in plants. Others, such as Co, Cu, Fe, Mn, Mo, Ni and Zn, are essential elements required for normal growth and metabolism of plants. These latter elements can easily lead to poisoning when their concentration rises to supra-optimal values. Heavy metal phytotoxicity may result from alterations of numerous physiological processes caused at cellular/molecular level by inactivating enzymes, blocking functional groups of metabolically important molecules, displacing or substituting for essential elements and disrupting membrane integrity. A rather common consequence of heavy metal poisoning is the enhanced production of reactive oxygen species (ROS) due to interference with electron transport activities, especially that of chloroplast membranes [1,2]. This increase in ROS exposes cells to oxidative stress leading to lipid peroxidation, biological macromolecule deterioration, membrane dismantling, ion leakage, and DNA-strand cleavage [3–5]. Plants resort to a series of defence mechanisms that control uptake, accumulation and translocation of these dangerous elements and detoxify them by excluding the free ionic forms from the cytoplasm (Fig. 1). One commonly employed strategy lies in hindering the entrance of heavy metals into root cells through entrapment in the apoplastic environment by binding them to exuded organic acids [6] or to anionic groups of cell walls [7,8]. Most of the heavy metals that do enter the plant are then kept in root cells, where they are detoxified by complexation with amino acids, organic acids or metal-binding peptides and/or sequestered into vacuoles [9]. This greatly restricts translocation to the above-ground organs thus protecting the leaf tissues, and particularly the metabolically active photosynthetic cells from heavy metal damage. A further defence mechanism generally adopted by heavy metal-exposed plants is enhancement of cell antioxidant systems which counteracts oxidative stress [4,10]. It is interesting to notice that there are plants that survive, grow and reproduce on natural metalliferous soils as well as on sites polluted with heavy metals as a result of anthropogenic activities. The majority of species that tolerate heavy metal concentrations that are highly toxic to the other plants behave as “excluders” (Fig. 1), relying on tolerance and even hypertolerance strategies helpful for restricting metal entrance. They retain and detoxify most of the heavy metals in the root tissues, with a minimized translocation to the leaves whose cells remain sensitive to the phytotoxic effects [9]. Nevertheless, a number of hypertolerant species, defined as “hyperaccumulators”, exhibit an opposite behaviour as far as heavy metal uptake and distribution in the plant is concerned (Fig. 1). 2. What are heavy metal hyperaccumulator plants? The term “hyperaccumulator” was coined [11] for plants (Fig. 1) that, differently from the excluder plants, actively take up exceed-

175 175 176 176 177 178

ingly large amounts of one or more heavy metals from the soil. Moreover, the heavy metals are not retained in the roots but are translocated to the shoot and accumulated in aboveground organs, especially leaves, at concentrations 100–1000-fold higher than those found in non-hyperaccumulating species. They show no symptoms of phytotoxicity [12,13]. Although a distinct feature, hyperaccumulation also relies on hypertolerance, an essential key property allowing plants to avoid heavy metal poisoning, to which hyperaccumulator plants are as sensitive as non-hyperaccumulators [14]. About 450 angiosperm species have been identified so far as heavy metal (As, Cd, Co, Cu, Mn, Ni, Pb, Sb, Se, Tl, Zn,) hyperaccumulators, accounting for less than 0.2% of all known species. However, new reports of this kind of plants continue to accrue [15–18], so that it is conceivable that many yet unidentified hyperaccumulators may occur in nature. On the other hand, some species classified as hyperaccumulators on the basis of field samples might be deleted from the list if this trait is unconfirmed by experimentation under controlled conditions [19]. For instance, the finding that in a number of cuprophytes the Cu and Co hyperaccumulation by field samples was actually due to leaf surface contamination has

Fig. 1. Mechanisms involved in heavy metal hypertolerance and heavy metal distribution in an excluder non-hyperaccumulator (left) and a hyperaccumulator (right) plant. (1) Heavy metal binding to the cell walls and/or cell exudates, (2) root uptake, (3) chelation in the cytosol and/or sequestration in vacuoles, (4) root-to-shoot translocation. The spots indicate the plant organ in which the different mechanisms occur and the spot sizes the level of each of them. According to the elemental defence hypothesis the high heavy metal concentrations make hyperaccumulator leaves poisonous to herbivores.

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

led to a critical re-examination of the Cu/Co hyperaccumulators [20,21]. The hyperaccumulator species are distributed in a wide range of distantly related families, showing that the hyperaccumulation trait has evolved independently more than once under the spur of selective ecological factors. The evolutionary reasons that gave rise to hyperaccumulating plants are unknown and still under debate. However, a series of hypotheses have been proposed. They will be discussed in the following section. Heavy metal hyperacumulators do occur on metal-rich soils in both tropical and temperate zones. They are found in vegetations from regions of South Africa, New Caledonia, Latin America as well as of North America and Europe [22]. Initially the term hyperaccumulator referred to plants able to accumulate more than 1 mg g−1 Ni (dry weight) in the shoot, an exceptionally high heavy metal concentration considering that in vegetative organs of most plants Ni toxicity starts from 10 to 15 ␮g g−1 . Threshold values were successively provided to define the hyperaccumulation of each other heavy metal, based on its specific phytotoxicity. According to such a criterion hyperaccumulators are plants that, when growing on native soils, concentrate >10 mg g−1 (1%) Mn or Zn, >1 mg g−1 (0.1%) As, Co, Cr, Cu, Ni, Pb, Sb, Se or Tl, and >0.1 mg g−1 (0.01%) Cd in the aerial organs, without suffering phytotoxic damage [23]. Ni is hyperaccumulated by the greatest number of taxa (more than 75%), while a low number of hyperaccumulators (only 5 species to date) has been found for Cd, which is one of the most toxic heavy metals. Ni is also the metal that has been shown to reach the highest concentration in a plant. This occurs in Sebertia acuminata (Sapotaceae), a tree endemic to the serpentine soil from New Caledonia, which accumulates up to 26% Ni (dry mass) in its latex [24]. About 25% of discovered hyperaccumulators belong to the family of Brassicaceae and, in particular, to genera Thlaspi and Alyssum. These also include the highest number of Ni hyperaccumulating taxa [25]. Zn hyperaccumulators are less numerous and include Arabidopsis halleri and species of Thlaspi, among Brassicaceae [22], and Sedum alfredii (Crassulaceae) [26]. A. halleri and S. alfredii, together with Thlaspi caerulescens and T. praecox, are the four recognized species that, besides Zn, hyperaccumulate Cd. Recently Solanum nigrum (Solanaceae) has been noticed as the fifth Cd hyperaccumulator [16]. Species hyperaccumulating Se are distributed in genera of different families, among which Fabaceae, Asteraceae, Rubiaceae, Brassicaceae, Scrophulariaceae and Chenopodiaceae [27]. Besides some angiosperms, such as the Brassicaceae Isatis cappadocica and Hesperis persica [18,28], a number of brake ferns belonging to the genus Pteris have also been found to hyperaccumulate As [29,30]. Most hyperaccumulators are endemic to metalliferous soils behaving as “strict metallophytes”, whereas some “facultative metallophytes” can live also on non-metalliferous ones, although are more prevalent on metal-enriched habitats [31]. Furthermore, there are species that include both metallicolous and non-metallicolous populations. In some of these, such as in Zn hyperaccumulators A. halleri and T. caerulescens, the hyperaccumulation is a constitutive trait at the species level, being found in all populations [32,33]. In others, such as in the Zn hyperaccumulator S. alfredii and in Cd hyperaccumulators, this trait, instead, is not constitutive at the species level, but only confined to metallicolous populations [23,34,35].

3. How do plants hyperaccumulate heavy metals? The degree of hyperaccumulation of one or more heavy metals can vary significantly in different species or also in populations and ecotypes of the same species [36,37]. However, hyperaccumulation depends on three basic hallmarks that distinguish hyperaccumula-

171

tors from related non-hyperaccumulator taxa. These common traits are: a much greater capability of taking up heavy metals from the soil; a faster and effective root-to-shoot translocation of metals; and a much greater ability to detoxify and sequester huge amounts of heavy metals in the leaves (Fig. 1). Significant progress in understanding the mechanisms governing metal hyperaccumulation has been made in the last decade through comparative physiological, genomic, and proteomic studies of hyperaccumulators and related non-hyperaccumulator plants. A great number of studies are on T. caerulescens and A. halleri, which have become model plants [38–40]. A very interesting feature revealed by this research is that most key steps in hyperaccumulation do not rely on novel genes, but depend on genes common to hyperaccumulators and nonhyperaccumulators, that are differently expressed and regulated in the two kinds of plants [23]. 3.1. Heavy metal uptake Comparative studies have revealed that the enhanced Zn uptake into T. caerulescens and A. halleri roots, in comparison to congener non-hyperaccumulator species, can be attributed to the constitutive overexpression of some genes belonging to the ZIP (Zinc-regulated transporter Iron-regulated transporter Proteins) family, coding for plasma membrane located cation transporters [41] (Fig. 2). Moreover, the expression of these ZIP genes (ZTN1 and ZTN2 in T. caerulescens and ZIP6 and ZIP9 in A. halleri), that in non-hyperaccumulating plants is Zn-regulated [42] and occurs at detectable levels only under Zn deficiency, in hyperaccumulators is irrespective of Zn supply still persisting at high Zn availability [41,43]. The decreasing uptake of Cd by roots supplied with increasing Zn concentration, found in Cd/Zn hyperaccumulator A. halleri and in most ecotypes of T. caerulescens, clearly demonstrates that Cd influx is largely due to Zn transporters (Fig. 2), with a strong preference for Zn over Cd [44]. Surprisingly, in plants of the Ganges ecotype of T. caerulescens, which exhibit an exceptionally high ability to hyperaccumulate Cd in aerial tissues, Cd uptake is not inhibited by Zn, thus suggesting the presence in root cells of a specific and efficient independent Cd transport system [45]. The supposed existence of a transporter specific to this metal, regarded as unessential, raises the question as to whether Cd might play some physiological roles in that T. caerulescens accession. In shoots of the Ganges plants a positive correlation between Cd concentration and carbonic anhydrase activity has been found [46]. The only physiological function of this heavy metal had previously been noticed in the marine diatom Thalassiosira weissglogii owing to its finding in the active metal-binding site of a peculiar Cd-containing carbonic anhydrase [47,48]. Specific transporters for Ni hyperaccumulation have not yet been recognized. However, the preference of Zn over Ni by some Zn/Ni hyperaccumulators supplied with the same concentration of both heavy metals strongly suggests that a Zn transport system (Fig. 2) might also be employed in Ni entrance into roots [49]. Considerable evidence exists that As can enter plant roots as arsenate via transporters of the chemical analogue phosphate [50] (Fig. 2). In root cells of As hyperaccumulator Pteris vittata plasma membranes have a higher density of phosphate/arsenate transporters than non-hyperaccumulator P. tremula, plausibly due to constitutive gene overexpression [51]. Furthermore, the enhanced As uptake by the hyperaccumulating fern depends on the higher affinity for arsenate by the phosphate/arsenate transport systems [52] as well as on the plant’s ability to increase As bioavailability in the rhizosphere by reducing pH via root exudation of large amounts of dissolved organic carbon [53]. The pH decrease, in fact, enhances the water soluble As that can be taken up by the roots [53,54]. The chemical similarity between sulphate and selenate accounts for the root uptake of Se in this form through high-affinity sulphate

172

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

Fig. 2. A scheme showing transport systems constitutively overexpressed and/or with enhanced affinity to heavy metals, which are though to be involved in uptake, root-toshoot translocation and heavy metal sequestration traits of hyperaccumulator plants. (CAX = Cation Exchangers; CDF = Cation Diffusion Facilitators; FDR3 = a member of the Multidrug and Toxin Efflux family; HM = Heavy Metals; HMA = Heavy Metal transporting ATPases; NA = Nicotinamine; NIP = Nodulin 26-like Intrinsic Proteins; P = Phosphate transporters; S = Sulphate transporters; YSL = Yellow Strip 1-Like Proteins; ZIP = Zinc-regulated transporter Iron-regulated transporter Proteins). For details and references see the text.

transporters (Fig. 2), whose activity is regulated by the S status of the plant [55,56]. In Se hyperaccumulators, such as Astragalus bisulcatus (Fabaceae) and Stanleya pinnata (Brassicaceae), the Se/S ratios in shoots are much higher than in non-hyperaccumulator sister species. This supports the idea of a role in this increased Se uptake of one or more sulphate transporters, which may have acquired a Se-specificity, becoming independent of the plant S status [57]. 3.2. Root-to-shoot translocation Differently from non-hyperaccumulator plants, which retain in root cells most of the heavy metal taken up from the soil, detoxifying them by chelation in the cytoplasm or storing them into vacuoles, hyperaccumulators rapidly and efficiently translocate these elements to the shoot via the xylem (Fig. 1). This entails, of course, the heavy metal availability for xylem loading, which derives from a low sequestration into and a ready efflux out of the vacuoles, plausibly due to specific features of root cell tonoplast [58]. As a matter of fact the amount of Zn sequestered into cell root vacuoles is 2–3-fold lower and the Zn efflux out of vacuoles almost twice as fast in the hyperaccumulators T. caerulescens [58] and S. alfredii [59] than in non-hyperaccumulating relatives. A lower sequestration into root vacuoles accounts also for the enhanced As translocation in hyperaccumulator compared with non-hyperaccumulator species of Pteris [52]. Constitutively large quantities of small organic molecules are present in hyperaccumulator roots that can operate as metalbinding ligands. However, the involvement of different chelators in hyperaccumulation strategies has not been quite established yet. The role of organic acids, mainly malate and citrate, as lig-

ands in the root cells is particularly controversial, due to their low association constants with metals that makes complexation negligible at cytosolic pH values. They may be relevant only within the acidic vacuolar environment [60]. A key role in heavy metal hyperaccumulation seems to be played by free amino acids, such as histidine and nicotinamine, which form stable complexes with bivalent cations [61]. Free histidine (His) is regarded as the most important ligand involved in Ni hyperaccumulation [61]. In roots of the Ni hyperaccumulator Alyssum lesbiacum, as compared with the non-hyperaccumulator A. montanum, the constitutive overexpression of the TP-PRT1 gene (encoding the ATP-phosphoribosyl transferase enzyme committed in the first step of the biosynthetic pathway) leads to a larger endogenous pool of His, which favours the Ni xylem loading as a Ni-His complex [62,63]. The high concentrations of His in roots of different Ni hyperaccumulating Thlaspi species suggests that the amino acid may operate in the same way in other hyperaccumulators [31]. Moreover, in hyperaccumulators, but not in non-hyperaccumulators, the Ni–His complexation, besides the involvement in sustaining the Ni release into the xylem, plays an essential role in preventing the heavy metal entrapment in root cell vacuoles, thus keeping it in the cytosol, in a detoxified form available for translocation [23,64]. Genes encoding enzymes of the nicotinamine biosynthetic pathway are overexpressed in roots of the Zn/Ni hyperaccumulators T. caerulescens and A. halleri which contain 3-fold higher amount of nicotinamine than roots of congener non-hyperaccumulating species [43,65,66]. However, in T. caerulescens, enhanced nicotinamine synthesis and nicotinamine–metal chelation show a positive correlation with Ni hyperaccumulation [67], whereas in A. halleri they are involved in Zn hyperaccumulation, with a possible role for the cytosolic nicotinamine–Zn complexes also in keeping metal

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

ions in detoxified form disposable for loading into xylem vessels [43,68]. A large body of evidence indicates that fast and efficient root-to-shoot translocation of large amounts of heavy metals in hyperaccumulator plants relies on enhanced xylem loading by a constitutive overexpression of genes coding for transport systems common to non-hyperaccumulators. The P1B -type ATPases, a class of proteins, also named HMAs (Heavy Metal transporting ATPases), are of particular importance (Fig. 2). They operate in heavy metal transport and play a role in metal homeostasis and tolerance [69]. Genes encoding bivalent cation transporters belonging to HMAs (among which HMA4) are overexpressed in roots and shoots of Zn/Cd hyperaccumulators T. caerulescens and A. halleri [66,70–72]. Moreover, the HMA4 expression is up-regulated when these plants are exposed to high levels of Cd and Zn, whereas it is down-regulated in non-hyperaccumulator relatives [70]. The overexpression of HMA4 supports a role of the HMA4 protein (which belongs to the Zn/Co/Cd/Pb HMA subclass and is localized at xylem parenchyma plasma membranes) in Cd and Zn efflux from the root symplasm into the xylem vessels, necessary for shoot hyperaccumulation. This role is upheld by QTL analysis showing co-localization of a major QTL for Zn and Cd tolerance with the HMA4 gene in A. halleri [73–75]. Interestingly, it has been demonstrated that the HMA4 activity positively affects other candidate genes for hyperaccumulation. In fact, the increased expression of HMA4 enhances the expression of genes belonging to the ZIP family, implicated in heavy metal uptake. This strongly suggests that the root-to-shoot translocation acts as a driving force of the hyperaccumulation, by creating a permanent metal deficiency response in roots [72]. The MATE (Multidrug And Toxin Efflux) family of small organic molecule transporters seems to be another kind of transport proteins that are active in heavy metal translocation in hyperaccumulator plants (Fig. 2). FDR3, a gene encoding a member of this family, is constitutively overexpressed in roots of T. caerulescens and A. halleri [66,76]. The FDR3 protein, which is localized at root pericycle plasma membranes, usually operates in the xylem influx of citrate, which is required as a ligand for Fe homeostasis and transport [77], but its overexpression in hyperaccumulators suggests that FDR3 might also play a role in translocation of other metals, such as Zn [78]. Moreover, evidence exists for the involvement in heavy metal translocation by YSL (Yellow Strip1-Like) family members (Fig. 2), which mediate the loading into and unloading out of xylem of nicotinamine–metal chelates [79]. Three genes (TcYSL3, TcYSL5 and YSL7), are constitutively overexpressed in roots and shoots of T. caerulescens where the YLS proteins do participate in vascular loading and translocation of nicotinamine–metal (especially nicotinamine–Ni) complexes [80]. The transport system involved in xylem loading of Ni–His complexes occurring in hyperaccumulator roots, has not yet been elucidated. Comparative analyses between the Ni hyperaccumulator Thlaspi goesingense and the non-hyperaccumulator T. arvense have revealed that, under non-toxic conditions, both species display similar root-to-shoot Ni transport rates [81]. The authors conclude that the hyperaccumulation ability of T. goesingense depends on a very efficient Ni detoxification and/or sequestration mechanisms, much more than on enhanced heavy metal translocation. The greater arsenic translocation to the shoot in hyperaccumulator P. vittata, as compared with non-hyperaccumulator ferns, occurs principally as arsenite, which accounts for over 90% of the As in the xylem sap [82]. This is because in the roots of hyperaccumulating ferns most of arsenate (AsV ) is quickly reduced to arsenite (AsIII ) by the activity of an exclusive glutathione dependent arsenate reductase [83]. The remaining arsenate can be loaded into xylem by phosphate transporters, while the efflux toward the vascular tissues of the predominant arsenite requires differ-

173

ent transport systems, which have yet to be identified. However, some evidence supports aquaglyceroporins of the NIP (Nodulin 26-like Intrinsic Proteins) subfamily as the most likely candidates (Fig. 2). These plasma membrane proteins, that specifically operate in arsenite transport in mammals [84], also mediate arsenite transport in plants [85,86]. Thus a high expression of such proteins might conceivably account for the arsenite transfer from root cell cytoplasm to xylem vessels in As hyperaccumulators [87]. Most of Se taken up by root cells of Se hyperaccumulators remains as selenate. Thus, its root-to-shoot translocation occurs through sulphate transport systems [88] (Fig. 2). Whether the long-distance xylem transport of heavy metals can occur in free ionic forms or through metal complexation with organic acids is still controversial. Most of Zn and Cd, for instance, is present as free hydrated cations in the xylem sap of T. caerulescens and A. halleri [89,90], and only one-third of Ni is bound to citrate in the xylem of hyperaccumulator Stackhousia tryoni (Celastraceae) [91]. Conversely, almost all Ni is complexed with citrate and other organic acids in the latex of the extreme Ni hyperaccumulator S. acuminata [92]. 3.3. Detoxification/sequestration Great efficiency in detoxification and sequestration is a key property of hyperaccumulators which allows them to concentrate huge amounts of heavy metals in above-ground organs without suffering any phytotoxic effect. This exceptionally high heavy metal accumulation becomes even more astonishing bearing in mind that it principally occurs in leaves where photosynthesis, essential for plant survival, is accomplished, and that the photosynthetic apparatus is a major target for most of these contaminants. The preferential heavy metal detoxification/sequestration do occur in locations, such as epidermis [93–96], trichomes [97] and even cuticle [98], where they do least damage to the photosynthetic machinery. In many cases heavy metals are also excluded from both subsidiary and guard cells of stomata [99–101]. This may preserve the functional stomatal cells from metal phytotoxic effects. The detoxifying/sequestering mechanisms in aerial organs of hyperaccumulators consist mainly in heavy metal complexation with ligands and/or in their removal from metabolically active cytoplasm by moving them into inactive compartments, mainly vacuoles and cell walls (Fig. 1). Comparative transcriptome analyses between hyperaccumulator and related non-hyperaccumulator species have demonstrated that also the sequestration trait relies, at least in part, on constitutive overexpression of genes that, in this case, encode proteins operating in heavy metal transfer across the tonoplast and/or plasma membrane and involved in excluding them from cytoplasm. CDF (Cation Diffusion Facilitator) family members, also named MTPs (Metal Transporter Proteins), which mediate bivalent cation efflux from the cytosol, are important candidates (Fig. 2). MTP1, a gene encoding a protein localized at tonoplast, is highly overexpressed in leaves of Zn/Ni hyperaccumulators [102–105]. It has been suggested that MTP1, besides the role in Zn tolerance, may also play a role in enhancing Zn accumulation. The Zn transport into the vacuole, in fact, may initiate a systemic Zn deficiency response that includes the enhancement of the heavy metal uptake and translocation via the increased expression of ZIP transporters in hyperaccumulator plants [105]. MTP members also mediate the Ni vacuolar storage in T. goesingense shoots [106]. Moreover, the finding that MTP1 is localized at both vacuolar and plasma membrane suggests that it can also operate in Zn and Ni efflux from cytoplasm to cell wall [103]. The overexpression of HMA3, coding for a vacuolar P1B -ATPase, plausibly involved in Zn compartmentation, and that of CAX genes encoding members of a cation exchanger family that seems to mediate Cd sequestration (Fig. 2), have been noticed in T.

174

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

caerulescens and A. halleri and supposed to be involved in heavy metal hyperaccumulation [107,108]. Storage of As as inorganic arsenite in vacuoles is a key mechanism found in fronds of hyperaccumulator ferns, although the transport system located at the tonoplast has not been identified yet [87]. Small ligands, such as organic acids, have a major role as detoxifying factors. Such ligands may be instrumental in preventing the persistence of heavy metals as free ions in the cytoplasm and even more in enabling their entrapment in vacuoles where the metal–organic acid chelates are primarily located. Citrate, for instance, is the main ligand of Ni in leaves of T. goesingense [109], while citrate and acetate bind Cd in leaves of S. nigrum [16]. Moreover, most Zn in A. halleri and Cd in T. caerulescens are complexed with malate [89,110]. The heavy metal detoxification in hyperaccumulators, in contrast with tolerant non-hyperaccumulator plants, does not rely on high molecular mass ligands, such as phytochelatins [111,112], likely because of the excessive sulphur amounts and the prohibitive metabolic cost that a massive synthesis of this kind of chelators would require [113]. Overexpression of antioxidation-related genes [114], as well as enhanced synthesis of glutathione (GSH) as pivotal antioxidant molecule [108], do occur, instead, in hyperaccumulators, as a strategy to reinforce the cell antioxidant system and cope with the risk of ROS rise due to heavy metal stress. The major detoxification strategy in Se hyperaccumulators is to get rid of selenoaminoacids, mainly selenocysteine (SeCys), derived from selenate assimilation in leaf chloroplasts. Selenoaminoacids are misincorporated in proteins instead of sulphur amino acids, resulting in Se toxicity. This detoxification occurs through methylation of Se-Cys to the harmless non-protein amino acid methylselenocysteine in a reaction catalyzed by a selenocysteine methyltransferase, which is constitutively expressed and activated only in leaves of hyperaccumulator species [115].

4. Why did plants evolve hyperaccumulation of heavy metals? The discovery of a class of plants that concentrate exceptionally high amounts of normally toxic heavy metals in leaves has attracted considerable interest, and challenged biologists to find reasons for this unusual behaviour by providing answers to the question: why do some plants do it? In other words: what functions does hyperaccumulation perform in these plants and what are the benefits and the adaptive values of metal hyperaccumulation? A variety of hypotheses have been proposed to explain the role of high elemental concentrations in leaves [116], namely: metal tolerance/disposal, drought resistance, interference with neighbouring plants, and defence against natural enemies. According to the tolerance/disposal hypothesis, the peculiar hyperaccumulation pattern would allow plants to take heavy metals away from the roots by sequestering them in tolerant leaf tissues. This eliminates them from the plant body by shedding the high-metal aerial organ. Another postulated explanation is that large amounts of heavy metals might increase plant drought resistance, with a waterconserving role in the cell walls or acting as osmolytes inside the cells. These hypotheses, however, are hardly supported by experimental evidence, so that their validation deserves further investigation. The interference hypothesis, also termed “elemental allelopathy”, suggests, instead, that perennial hyperaccumulator plants may interfere with neighbouring plants through enrichment of metal in the surface soil under their canopies. This gives rise to a high-metal leaf litter that prevents the establishment of less metal tolerant species. One group has measured higher Ni levels in the surface soil under the canopy of hyperaccumulator S. acuminata

than under that of non-hyperaccumulator species [117]. However, another group has questioned elemental allelopathy finding that high-Ni leaves shed by Alyssum murale do not create a “toxic zone” around the Ni hyperaccumulator and do not inhibit seed germination of competing plants [118]. The lack of allelopathic effect is probably due to the fact that most Ni released from the leaf biomass does not remain in a soluble and phytoavailable form, but is rapidly bound to soil constituents thus becoming unable to affect neighbouring plants. Also the hypothesis of elemental allelopathy does not have satisfactory experimental verification yet [119]. 4.1. The “elemental defence” hypothesis The hypothesis which has attracted most attention suggests that the high heavy metal concentrations in aerial tissues may function as a self-defence strategy evolved in hyperaccumulator plants against some natural enemies, such as herbivores (Fig. 1) and pathogens. This “elemental defence” hypothesis has been widely tested, gaining much supporting evidence, although some tests have led to contradictory responses. Some recent studies, for instance, confirm the defensive function of Ni [120], Cd [121], Zn [122], As [123] and Se [124] while others [125,126] seem to counter the heavy metal involvement in plant defence. Despite the numerous reports regarding this popular hypothesis, more information is required since few taxa have been tested, the majority of studies have focused on Brassicaceae and only some elements (Ni, Zn, Cd, As, Se) have been examined. Moreover, the defensive effects have been analyzed mostly in laboratory conditions and have considered only one or a few selected herbivores, rather than being tested in the field where hyperaccumulators have to face an array of natural enemies [127]. The mix of contrasting conclusions reported in literature about the effectiveness of heavy metals as defence elements might depend on different experimental conditions or heavy metal concentrations used in each study, as well as on the ability of certain herbivores to overcome the plant defence [127]. Heavy metals, actually, may provide protection against a broad range of enemies that the plant encounters in natural situations, whereas some others may be able to feed on a hyperaccumulator species despite its elemental composition. Mechanisms enabling herbivores to circumvent the heavy metal defence might be “avoidance” which leads an herbivore to selectively eat only low-metal tissues of the plant and “dietary dilution”, which consists of lowering overall metal ingestion by eating both high-metal and low-metal tissues [127]. Another mechanism deserving a particular interest is “tolerance”, in which physiological adaptations allow specialist herbivores to withstand a high-metal diet, thus disarming the elemental defence of the plant [128,129]. The bug Melanotrichus boydi, for instance, prefers to feed on the Ni hyperaccumulator Streptanthus polygaloides [130] and a strain of the moth Plutella xylostella feeds on the Se hyperaccumulator S. pinnata without suffering from the high-Se diet [131]. Heavy metals can act against herbivore through their toxicity, but this does not safeguard the plant from undergoing damage before poisoning the enemy. Thus a more effective defence from herbivore attack should be through feeding deterrence. Experimental evidence exists that some herbivores prefer to eat low-Zn T. caerulescens [132] and low-Ni Senecio coronatus [133] when offered a choice between plants containing either low or high-metal concentrations. Deterrent effects have also been shown for Cd [121], As [123] and Se [134]. This ability to avoid feeding on plants with high heavy metal levels might support the view that herbivores have a “taste for metals”, although no information exists on how they might do it. However, since the metal treatment will strongly affect the plants’ metabolome, it might be that herbivores do not directly perceive metals in their food, but rather metal-induced metabolites.

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

175

4.2. The “joint effects” hypothesis Chemical defence of plants from enemy attack can also involve a variety of organic (secondary metabolite) compounds. However, elemental defence offers some advantages over organic defence [127]: the toxic elements are not synthesized by the plant but taken up from the soil thus making the elemental defence less metabolically expensive than the organic one; the inorganic elements cannot be biochemically degraded by most herbivores (although some specialistic herbivores can chelate or sequester them), so that this counter defence mechanism of the enemies is prevented. The rather high cost of secondary metabolite biosynthesis accounts for a “trade off” hypothesis, in which the metal-based defence of hyperaccumulator plants may have evolved to reduce levels of energy-demanding organic defences. Lower levels of defensive glucosinolates found in Ni hyperaccumulator S. polygaloides [135] and in Zn hyperaccumulator T. caerulescens [136], when compared with congener non-hyperaccumulator species, support the view of a trade off between metal hyperaccumulation and secondary metabolite synthesis. However, the trade off hypothesis remains somewhat controversial. Differences in concentration of specific glucosinolates, but not in their total amount between Ni hyperaccumulating and non-hyperaccumulating plants, have been measured [120], and it has been suggested that glucosinolates rather than Zn are involved as deterrents in antiherbivore defences of T. caerulescens populations [136]. Joint defensive effects may actually exist between elemental and organic plant compounds, which may act in concert with each other and enhance plant defence overall [127]. Heavy metal and several defensive organic metabolites operate additively against an herbivore enemy in Ni hyperaccumulator S. polygaloides [137]. This new joint effects hypothesis may justify the simultaneous presence of elemental and organic defences as well as the hyperaccumulation of more than one heavy metal in the same plant. Joint effects between heavy metals, besides that between a heavy metal and an organic chemical, have been highlighted [137]. However also this new interesting idea of a defensive enhancement achieved in hyperaccumulators through joint effects of elemental and organic plant defences needs to be supported by future investigation. 5. Why do hyperaccumulators attract so much interest? Besides their ecological and physiological interest, hyperaccumulator plants have received considerable attention due to the possibility of exploiting their accumulation traits for practical applications, in particular to develop technologies for phytoremediation of heavy metal contaminated soils or for mining valuable metals from mineralized sites. 5.1. Potential application in phytoremediation The last two decades have seen the emergence of eco-friendly soil remediation techniques, collectively known as phytoremediation, that utilize plant species. The use of plants to remediate polluted soils is seen as having great promise compared to conventional, civil-engineering methods and several recent comprehensive reviews summarising the most important aspects of soil metal phytoremediation are available [14,138–142]. The naturally occurring heavy metal hyperaccumulator plants which, when growing in metal-enriched habitats, can accumulate 100–1000fold higher levels of metals than normal plants are excellent candidates for phytoextraction (Fig. 3), as these plants take up from the soil two or three orders of magnitude more metals than plant species growing on uncontaminated soils. Chaney et al. [143] are the first to have proposed the exploitation of heavy metal hyperaccumulator plants to clean up polluted

Fig. 3. Phytoremediation and phytomining of heavy metals rich soils by using plants which hyperaccumulate these metals in above-ground organs. The harvesting of the aerial part of the plants leads to the disposal of the huge amounts of toxic heavy metals removed from the soil or to the recovery of the valuable metals taken up.

sites. However, hyperaccumulators have been later believed to have limited potential of phytoremediation because most of them are metal selective, have not been found for all elements of interest, can be used in their natural habitats only, and, above all, have small biomass, shallow root systems and slow growth rates, which limit the speed of metal removal [144,145]. In addition, there is no knowledge about the agronomics, genetics, breeding potential, and disease spectrum of these plants. This is the case for many hyperaccumulator plants including the Zn/Cd hyperaccumulator T. caerulescens, which gives a maximum of 2 tons ha−1 of shoot dry matter. Although the annual biomass yield is an important trait for phytoremediation, the ability to hyperaccumulate and hypertolerate metals is of greater importance than high biomass [14]. Pot and field studies have shown that the hyperaccumulator T. caerulescens grown as a crop can attain as high as 5 tons ha−1 by breeding to increase the combination of yield and shoot metal concentration [146]. Furthermore, the recycling of shoot metals may provide added value to the ash from metal hyperaccumulators, so that there is no need to pay for disposal of the plants. Various species of Thlaspi are known to hyperaccumulate more than one metal. Predominantly, Thlaspi grows on Ni contaminated sites and accumulates about 3% of its dry matter as metal but T. caerulescens can accumulate Cd, Ni, Zn and also Pb. As a hyperaccumulator of Cd and Zn it could remove as much as 60 kg Zn ha−1 and 8.4 Kg Cd ha−1 [147]; T. goesingese and T. ochroleucum hyperaccumulate Ni and Zn while T. rotundifolium hyperaccumulates Ni, Pb and Zn [148]. The brake fern P. vittata, which produces a relatively large biomass under favourable climate conditions, accumulates (from relatively low As concentration in the soil) 22 g As kg−1 in the frond dry weight, with a bioconcentration factor of 87 and a removal of 26% of the soil’s initial As ([29,139] and literature cited therein). These results suggest that phytoremediation of at least moderately Ascontaminated sites is feasible. Although Pb is largely immobile in soil and its extraction rate is limited by solubility and diffusion to the root surface, common buckwheat (Fagopyrum esculentum, Polygonaceae), the first known Pb hyperaccumulator species with high biomass, can accumulate up to 4.2 mg g−1 dry weight of Pb in the shoots [149]. Amending the soil with the biodegradable methylglycine diacetic acid (MGDA) resulted in a 5-fold increase in the Pb shoot concentration. This relevant finding qualifies this species as an excellent candidate for remediating Pb-contaminated soils. Phytolacca acinosa (Phytolaccaceae), a plant that grows rapidly and has substantial biomass, has been considered to have potential for use in phytoremediation. The plant can accumulate 19.3 g Mn kg−1

176

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

dry weight when grown on Mn-rich soils [150]. The efficiency of Alyssum serpyllifolium subsp. lusitanicum for use in phytoextraction of polymetal-contaminated soils has been examined [151]. The plants have been grown on soils contaminated with Cr, Cu, Pb and Zn. The results suggest that A. serpyllifolium can be suitable for phytoextraction in polymetal-polluted soils, provided that Cu concentrations are not phytotoxic. However, with the hyperaccumulators available, decades are needed to clean up contaminated sites. It has been calculated that to decrease Zn concentration from 440 to 300 mg Zn kg−1 soil nine croppings of T. caerulescens would be required [152] and 28 years of cultivation of this plant would be needed to remove 2100 mg Zn kg−1 from a soil [153]. T. caerulescens is useful for moderately Zn- and Cd-contaminated soils but would take far too long on highly contaminated ones. It also appears that season length, method of sowing seeds and soil pH have effects on the Zn and Cd extraction capacity of T. caerulescens from the soil [154,155]. The efficiency of phytoextraction, besides biomass production, depends on the metal bioconcentration factor (plant to soil concentration ratio) and for Zn and Cd it decreases log-linearly with soil metal concentration [156]. Moreover, the phytoremediation potential differs between different population of T. caerulescens. The southern French ecotype showed a higher ability to accumulate Cd than Zn: the different uptake of Cd and Zn shows that there are basic differences in the mechanism of accumulation of both metals in hyperaccumulators [45,154]. Thus, increased selection for traits of interest may help to improve the phytoremediation capacity of hyperaccumulators. A crop of T. caerulescens or A. halleri could, after cropping, remove decades-worth of Cd accumulation from pastures that have been treated with Cd-rich phosphate fertilizers [157]. Small-scale field experiments have also been conducted with Alyssum bertolonii and Berkheya coddii, fast-growing Ni hyperaccumulators [158,159]. The combination of high biomass and high-Ni content, together with its easy propagation and culture as well as its tolerance to cool climate conditions, should render B. coddii suitable for Ni-phytoremediation. For moderate Ni contamination two crops of B. coddii would be sufficient to reduce the metal concentration to well below the EU guidelines and for A. bertolonii, which has a lower biomass, from 5 to 10 annual crops would be needed. A 2year field study has been conducted to determine the efficiency of the fern P. vittata on As removal at an As-contaminated site. Approximately 19.3 g of As have been removed from the soil and it has been estimated that 8 years would be needed to completely remediate the soil in order to meet the residential site and/or commercial site requirements [160,161]. However, some estimates are based on achieving a soil clean up goal of 40 mg As kg−1 from an average of 82 mg As kg−1 [161], while that of others [160] contains on average 190 mg kg−1 . The presence of Pb in one soil [161] may have hindered the ability of P. vittata from removing As from it. It has been demonstrated that T. caerulescens, although being able to mine the soil metals more efficiently than non-accumulator plants, exhibits the same capacity of non-accumulators to increase metal availability in the rhizosphere, so that the use of amendments to raise metal solubility has been suggested. A number of possible amendments such as ethylene diamine tetracetic acid (EDTA), nitrilotriacetic acid (NTA) and citric acid have been tested under field conditions and seem to have no effect on increasing metal content, but may actually decrease biomass production of hyperaccumulators, thus reducing their potential of phytoextraction [139,154,158,159,162]. Recently, the effects of polycyclic aromatic hydrocarbons (PAHs) on the extractability of Ni by A. lesbiacum have been investigated. Plant growth is negatively affected by PAHs; however, there is no significant effect on the phytoextraction of Ni per unit biomass of shoot, indicating that A. lesbiacum might be effective in phytoextracting Ni from marginally PAHcontaminated soils [163]. No hyperaccumulators of radionuclides

have been reported so far, thus it is unlikely that the hyperaccumulation strategy is possible for these contaminants In summary, despite intensive research very few field studies or commercial operations that demonstrate successful phytoextraction by hyperaccumulators have been realized, so it could be considered a valuable tool only for Ni and As, while for the other metals the technology still appears to be far from the practice. At the moment phytoextraction with hyperaccumulators is an option to decontaminate soils with low to moderate metal concentrations while for highly contaminated soils it should be considered as a long-term remediation process. 5.2. Transfer of hyperaccumulation traits to rapidly growing species A promising biotechnological approach for enhancing the potential for metal phytoextraction, may be to improve the hyperaccumulator growth rate through selective breeding, or by the transfer of metal hyperaccumulation genes to high biomass species. In an effort to correct for small sizes of hyperaccumulator plants, somatic hybrids have been generated between T. caerulescens and Brassica napus. High biomas hybrid selected for Zn tolerance are capable of accumulating Zn level that would have been toxic to B. napus [164] indicating that the transfer of the metal hyperaccumulating phenotype is feasible. Somatic hybrids from T. caerulescens and B. juncea are also able to remove significant amounts of Pb [165]. T. caerulescens has also been used as source of genes for developing plant species better suited for phytoremediation of metal contaminated soils [166]. Bioengineered plants tolerant to the presence of toxic levels of metals like Cd [167], Zn, Cr, Cu, Pb [168], As [169] and Se [170] have been reported. A combination of transporter genes has also been used in rapidly growing plant species leading to promising results [169,171–173]. Transgenic B. juncea, grown either in hydroponic or in soils, shows higher uptake of Se and enhanced Se tolerance than the wild species [174,175]. To engineer Se tolerance the selenocysteine methyltransferase (SMT) gene has been transferred from the Se hyperaccumulator A. bisulcatus to Se-non-tolerant B. juncea. SMT transgenic plants of B. juncea grown in a contaminated soil accumulate 60% more Se than the wild-type ([176] and literature reported therein). The transgenic plant approach has shown to be promising, but only very few studies have been performed till now under field conditions [176]. Moreover, it has to be considered that tolerance and accumulation of heavy metals and thus phytoextraction potential of a given plant are controlled by many genes, so that genetic manipulations to improve these traits in fast-growing plants will require to change the expression levels in a number of genes, and to cross them to determine the number of genes involved and their characteristics. Functions and regulations of genes involved in metal hyperaccumulation, uptake, root-to-shoot translocation, detoxification/sequestration mechanisms need to be fully understood to render transgenic approach not far to solve the problem. 5.3. Potential application in phytomining Phytomining (a subset of phytoextraction) aims to generate revenue by recovering marketable amounts of metals from plant biomass (bio-ores) [177] through the use of plants to mine valuable heavy metals from contaminated or mineralized soils (Fig. 3). A pioneer phytomining study has been carried out using the Ni hyperaccumulator S. polygaloides [178]: a yield of 100 kg ha−1 of sulphur-free Ni could be obtained after moderate application of fertilizers. The removal of Ni from soil using phytomining is viable in principle, since there are many hyperaccumulator plants, such as Alyssum spp. and B. coddii, fulfilling the criterion of achieving shoot Ni concentrations higher than 10 g kg−1 on a dry matter basis and

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

producing more than 10,000 kg ha −1 per year [177]. A. bertolonii can also accumulate 10 mg Ni g−1 dry matter from serpentine soils [179]. Experiments have been carried out on the potential use of this hyperaccumulator plant in phytomining of serpentine soils. In the field trial plants of A. bertolonii have been fertilized with N + P + K over a period of 2 years. Fertilization increases biomass 3-fold without dilution of Ni concentration in the fertilized plants. It has been concluded that A. bertolonii, with a biomass after fertilization of about 13,500 kg ha−1 , or other species of Alyssum might be used for phytomining [158]. In another field trial B. coddii, with an unfertilized biomass of 12,000 kg ha−1 , has been reported as one of the best candidates for phytomining of Ni with applied fertilizers and adequate moisture, after which a biomass of 22,000 kg ha−1 and a high-Ni concentration has been achieved [159,162,177,180]. The potential of this species for phytomining has also been evaluated and a yield of 100 kg ha−1 of Ni should be achievable at many sites worldwide [180]. On the basis of biomass, the highNi concentration in the harvestable parts of the plants and the additional money obtained from the energy of combustion either of the Ni hyperaccumulator S. polygaloides or A. bertolonii, it has been concluded that the return to a farmer growing a “crop of nickel” would be comparable, or even superior, to that obtained for a crop of wheat [158,181]. Furthermore, if the above plants are used for phytoremediation of Ni polluted soils as a result of industrial activity, it would surely be of economic benefit considering the very high costs of conventional extraction methods and of storage of the toxic materials. Commercial phytomining technologies employing Alyssum Ni hyperaccumulator species have been developed [100]. However, hyperaccumulator plants might realistically also be expected to be used for Au, Tl, Co and U as well. Each has a high world price for the target metal and plants might extract from soils or mine tailing containing concentrations of the metals at a level uneconomic for conventional extraction techniques. Iberis intermedia and Biscutella (Brassicaceae) have been proposed for phytomining of Tl [180]. For other less valuable metals (Pb, Cd, 137 Cs, Cu, Se) phytomining will never emerge as a profitable agricultural industry. Notwithstanding all these promising field studies and the reported advantages over conventional mining [162], up to now there is no report of successful commercial phytomining operations. The potential limiting factors to the commercialization of phytomining have been investigated [182] and it has been concluded that is only attractive when applied to a contaminated site and might be usefully combined with conventional mining. In conclusion, phytomining with high-biomass hyperaccumulators could have economic advantages over traditional mining techniques, especially in cases where the extracted metals are biomining targets, have economic value and the energy of combustion of biomass can be sold. In addition, as bio-ore is practically sulphurfree its smelting does not contribute to acid rain. At the moment there is need to develop methods to recover and market the metals. Despite the large number of hyperaccumulators found to date, there is insufficient information on the distribution of these species or their uptake mechanisms so that they can be properly utilized in phytomining. Neither are their agronomical properties, such as fertilizer requirements, soil pH management, weed control and water requirements, adequately known. Notwithstanding these limitations it is clear that the commercialization of phytomining using high-biomass hyperaccumulator plants depends essentially on the metal concentration of the plant, its annual biomass production and the world price of the target metal.

6. Conclusions and future directions The problem of heavy metal pollution is continuously worsening due to a series of human activities, leading to intensification of

177

the research dealing with the phytotoxicity of these contaminants and with the mechanisms used by plants to counter their harmful effects. Great interest has been gained by the behaviour of hyperaccumulator plants growing on metalliferous soils, which accumulated heavy metals in leaves at concentrations several 100-fold higher than other plants. Aims of studying these heavy metal hyperaccumulator species has been to highlight physiological and molecular mechanisms underlying the hyperaccumulation ability, to discover the adaptive functions performed by hyperaccumulation in these plants and to explore the possibility of using them as tools to remove metals from contaminated or natural metal-rich soils. However, in spite of important progress made in recent years by the numerous studies accomplished, the complexity of hyperaccumulation is far from being understood and several aspects of this astonishing feature still await explanation. Hyperaccumulator plants, which are widespread on metal soils in both tropical and temperate zones of all the continents, belong to several unrelated families. This shows that the hyperaccumulation capability has been evolved more than once, although its adaptive value is still under debate. The recent idea that heavy metals would provide an elemental defence to the plant through joint effects with organic defence compounds requires much experimental investigation. More elements and a larger number of hyperaccumulator species need to be examined to validate the hypothesis of defensive effects of heavy metals. Furthermore, the investigations need to move from laboratory to field settings to provide realistic information about elemental defences in natural environments, where a plant can be exposed to a plethora of herbivores with different feeding modes, as well as to pathogens and parasites. Considerable attention has been given to the possibility of using hyperaccumulators for phytoremediation/phytomining of contaminated or natural metal-rich soils. However, more extensive research under field conditions for longer durations is required taking also into account that a specific phytoextraction prescription, due to the different site-specific conditions, cannot be applied to every site, even if with the same chemical composition. It is of pivotal importance to increase the understanding of hyperaccumulator-based remedial mechanisms because they will be able to provide clues for optimizing the effectiveness of phytoextraction with appropriate agronomic practices. In addition, knowledge acquired on genes involved in hyperaccumulation mechanisms will open the opportunity to use biotechnology to transfer specific genes to high-biomass promising species. Moreover, much research is still needed on rhizosphere and soil microbial composition under field conditions, in order to identify micro-organisms associated with metal solubility or precipitation. There is also an urgent need to find and characterize other hyperaccumulators, to cultivate them and better assess agronomic practices and management to enhance plant growth and metal uptake by selective breeding and gene manipulation. Even then, metal uptake might pose environmental risks, unless the biomass produced during the phytoremediation process could be rendered economical by burning it to produce bio-ore or converting it into bioenergy. However it is only matter of time before the commercialization of phytoextraction using high-biomass hyperaccumulator plants becomes widespread, considering that not only will it remediate contaminated sites but will generate income from agricultural lands otherwise not utilized. Last but not least, it has to be pointed out the interest in the potential exploiting of hyperaccumulators as a rich genetic resource to develop engineered plants with enhanced nutritional value for improving public health [183] or for contending with widespread mineral deficiencies in human vegetarian diets [184]. The strategies of food crop biofortification are still in infancy; however their paramount importance for the world’s population makes

178

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

this an exciting line of future research in the field of essential elements hyperaccumulation.

References [1] C. Pagliano, et al., Evidence for PSII-donor-side damage and photoinhibition induced by cadmium treatment on rice (Oryza sativa L.), J. Photochem. Photobiol. B: Biol. 84 (2006) 70–78. [2] N. La Rocca, C. Andreoli, G.M. Giacometti, N. Rascio, I. Moro, Responses of the Antarctic microalga Koliella antartica (Trebouxiophyceae, Chlorophyta) to cadmium contamination, Photosynthetica 47 (2009) 471–479. [3] M.F. Quartacci, E. Cosi, F. Navari-Izzo, Lipids and NADPH-dependent superoxide production in plasma membrane vesicles from roots of wheat grown under copper deficiency and excess, J. Exp. Bot. 152 (2001) 67–75. [4] F. Navari-Izzo, M.F. Quartacci, C. Pinzino, F. Dalla Vecchia, C. Sgherri, Thylakoid-bound and stromal antioxidative enzymes in wheat treated with excess of copper, Physiol. Plant. 104 (1998) 630–638. [5] F. Navari-Izzo, C. Pinzino, M.F. Quartacci, C. Sgherri, Superoxide and hydroxyl radical generation, and superoxide dismutase in PSII membrane fragments from wheat, Free Radic. Res. 31 (1999) S3–9. [6] T. Watanabe, M. Osaki, Mechanism of adaptation to high aluminium condition in native plant species growing in acid soils: a review, Commun. Soil Sci. Plant Anal. 33 (2002) 1247–1260. [7] F. Dalla Vecchia, et al., Morphogenetic, ultrastructural and physiological damages suffered by submerged leaves of Elodea canadensis exposed to cadmium, Plant Sci. 168 (2005) 329–338. [8] N. Rascio, et al., Metal accumulation and damage in rice (c.v. Vialone nano) seedlings exposed to cadmium, Environ. Exp. Bot. 62 (2008) 267–278. [9] J.L. Hall, Cellular mechanisms for heavy metal detoxification and tolerance, J. Exp. Bot. 53 (2002) 1–11. [10] C. Sgherri, E. Cosi, F. Navari-Izzo, Phenols and antioxidative status of Raphanus sativus grown in copper excess, Physiol. Plant. 118 (2003) 21–28. [11] R.R. Brooks, J. Lee, R.D. Reeves, T. Jaffré, Detection of nickeliferous rocks by analysis of herbarium specimens of indicator plants, J. Geochem. Explor. 7 (1977) 49–57. [12] N. Rascio, Metal accumulation by some plants growing on zinc-mine deposits, Oikos 29 (1977) 250–253. [13] R.D. Reeves, Hyperaccumulation of trace elements by plants, in: J.L. Morel, G. Echevarria, N. Goncharova (Eds.), Phytoremediation of Metal-Contaminated Soils, NATO Science Series: IV: Earth and Environmental Sciences, Springer, NY, 2006, pp. 1–25. [14] R.L. Chaney, et al., Phytoremediation of soil metals, Curr. Opin. Biotechnol. 8 (1997) 279–284. [15] B. Robinson, et al., Arsenic hyperaccumulation by aquatic macrophytes in the Taupo Volcanic Zone, New Zealand, Environ. Exp. Bot. 58 (2006) 206–215. [16] R. Sun, Q. Zhou, C. Jin, Cadmium accumulation in relation to organic acids in leaves of Solanum nigrum L. as a newly found cadmium hyperaccumulator, Plant Soil 285 (2006) 125–134. [17] P. Venkatachalam, A. Jain, S. Sahi, K. Raghothama, Molecular cloning and characterization of phosphate (Pi) responsive genes in Gulf ryegrass (Lolium multiflorum L.): a Pi hyperaccumulator, Plant Mol. Biol. 69 (2009) 1–21. [18] N. Karimi, S.M. Ghaderian, A. Raab, J. Feldmann, A.A. Meharg, An arsenicaccumulating, hypertolerant brassica, Isatis cappadocica, New Phytol. 184 (2009) 41–47. [19] M.R. Macnair, The hyperaccumulation of metals by plants, Adv. Bot. Res. 40 (2003) 63–105. [20] M.-P. Faucon, M. Ngoy Shutcha, P. Meerts, Revisiting copper and cobalt concentrations in supposed hyperaccumulators from SC Africa: influence of washing and metal concentrations in soil, Plant Soil 301 (2007) 29–36. [21] M.-P. Faucon, et al., Soil influence on Cu and Co uptake and plant size in the cuprophytes Crepidorhopalon perennis and C. tenuis (Scrophulariaceae) in SC Africa, Plant Soil 317 (2009) 201–212. [22] A.J.M. Baker, R.R. Brooks, Terrestrial higher plants which hyperaccumulate metallic elements–a review of their distribution, ecology and phytochemistry, Biorecovery 1 (1989) 81–126. [23] N Verbruggen, C. Hermans, H. Schat, Molecular mechanisms of metal hyperaccumulation in plants, New Phytol. 181 (2009) 759–776. [24] S. Sagner, et al., Hyperaccumulation, complexation and distribution of nickel in Sebestia acuminata, Phytochemistry 47 (1998) 339–347. [25] R.R. Brooks (Ed.), Plants that Hyperaccumulate Heavy Metals, CAB International, Wallingford, UK, 1998, p. 380. [26] X.E. Yang, et al., Cadmium tolerance and hyperaccumulation in a new Znhyperaccumulating plant species (Sedum alfredii Hance), Plant Soil 259 (2004) 181–189. [27] R.D. Reeves, A.J.M. Baker, Metal-accumulating plants, in: I. Raskin, B.D. Ensley (Eds.), Phytoremediation of Toxic Metals: Using Plants to Clean up the Environment, John Wiley & Sons, 2000, pp. 193–229. [28] N. Karimi, S.M. Ghaderian, H. Maroofi, H. Schat, Analysis of arsenic in soil and vegetation of a contaminated area in Zarshuran, Iran, Int. J. Phytoremed. 12 (2010) 159–173. [29] L.Q. Ma, K.M. Komar, C. Tu, W. Zhang, Y. Cai, A fern that hyperaccumulates arsenic, Nature 409 (2001) 579. [30] H.B. Wang, et al., Uptake and accumulation of arsenic by 11 Pteris taxa from southern China, Environ. Pollut. 145 (2007) 225–233.

[31] A.G.L. Assunc¸ão, H. Schat, M.G.M. Aarts, Thlaspi caerulescens, an attractive model species to study heavy metal hyperaccumulation in plants, New Phytol. 159 (2003) 351–360. [32] V. Bert, M.R. Macnair, P. De Laguerie, P. Saumitou-Laprade, D. Petit, Zinc tolerance and accumulation in metallicolous and nonmetallicolous populations of Arabidopsis halleri (Brassicaceae), New Phytol. 146 (2000) 225–233. [33] J. Escarré, et al., Zinc and cadmium hypraccumulation by Thlaspi caerulescens from metalliferous and nonmetalliferous sites in Mediterranean area: implication for phytoremediation, New Phytol. 145 (2000) 429–437. [34] X.E. Yang, et al., Dynamics of zinc uptake and accumulation in the hyperaccumulating and nonhyperaccumulating ecotypes of Sedum alfredii Hance, Plant Soil 284 (2006) 109–119. [35] V. Bert, et al., Genetic basis of Cd tolerance and hyperaccumulation in Arabidopsis halleri, Plant Soil 249 (2003) 9–18. [36] N. Roosens, N. Verbruggen, P. Meerts, P. Ximenez-Embun, J.A.C. Smith, Natural variation in cadmium tolerance and its relationship to metal hyperaccumulation for seven populations of Thlaspi caerulescens from Western Europe, Plant Cell Environ. 26 (2003) 1657–1672. [37] D.M. Deng, et al., Zinc and cadmium accumulation and tolerance in populations of Sedum alfredii, Environ. Pollut. 147 (2007) 381–386. [38] M.J. Milner, L.V. Kochian, Investigating heavy-metal hyperaccumulation using Thlaspi caerulescens as a model system, Ann. Bot. 102 (2008) 3–13. [39] U. Krämer, Metal hyperaccumulation in plants, Annu. Rev. Plant Biol. 61 (2010) 517–534. [40] H. Frérot, et al., Genetic architecture of zinc hyperaccumulation in Arabidopsis halleri: the essential role of QTLx environment interactions, New Phytol. 187 (2010) 355–367. [41] A.G.L. Assunc¸ão, et al., Elevated expression of metal transporter genes in three accessions of the metal hyperaccumulator Thlaspi caerulescens, Plant Cell Environ. 24 (2001) 217–226. [42] A.G.L. Assunc¸ão, et al., Arabidopsis thaliana transcription factors bZIP19 and bZIP23 regulate the adaptation to zinc deficiency, Proc. Natl. Acad. Sci. U.S.A. 107 (2010) 10296–10301. [43] M. Weber, E. Harada, C. Vess, E. von Roepenack-Lahaye, S. Clemens, Comparative microarray analysis of Arabidopsis thaliana and Arabidopsis halleri root identifies nicotinamine synthase, a ZIP transporter and other genes as potential metal hyperaccumulation factors, Plant J. 37 (2004) 269–281. [44] F.J. Zhao, R.E. Hamon, E. Lombi, M.J. McLaughlin, S.P. McGrath, Characteristics of cadmium uptake in two contrasting ecotypes of the hyperaccumulator Thlaspi caerulescens, J. Exp. Bot. 53 (2002) 535–543. [45] E. Lombi, F.J. Zhao, S.P. McGrath, S.D. Young, G.A. Sacchi, Physiological evidence for a high-affinity cadmium transporter highly expressed in a Thlaspi caerulescens ecotype, New Phytol. 149 (2001) 53–60. [46] M.-Q. Liu, Does cadmium play a physiological role in the hyperaccumulator Thlaspi caerulescens? Chemosphere 71 (2008) 1276–1283. [47] T.W. Lane, F.M.M. Morel, A biological function for cadmium in marine diatoms, Proc. Natl. Acad. Sci. U.S.A. 97 (2000) 4627–4631. [48] T.W. Lane, et al., A cadmium enzyme for a marine diatom, Nature 435 (2005) 42. [49] A.G.L. Assunc¸ão, P. Bleeker, W.M. Ten Bookum, R. Vooijs, H. Schat, Intraspecific variation of metal preference patterns for hyperaccumulation in Thlaspi caerulescens: evidence for binary metal exposures, Plant Soil 303 (2008) 289–299. [50] A.A. Meharg, J. Hartley-Whitaker, Arsenic uptake and metabolism in arsenic resistant and non-resistant plant species, New Phytol. 154 (2002) 29–42. [51] N. Caille, F.J. Zhao, S.P. McGrath, Comparison of root absorption, translocation and tolerance of arsenic in the hyperaccumulator Pteris vittata and the nonhyperaccumulator Pteris tremula, New Phytol. 165 (2005) 755–761. [52] C.Y. Poynton, J.W.W. Huang, M.J. Blaylock, L.V. Kochian, M.P. Ellass, Mechanisms of arsenic hyperaccumulation in Pteris species: root As influx and translocation, Planta 219 (2004) 1080–1088. [53] M.I. Gonzaga, L.Q. Ma, J.A. Santos, M.I. Matias, Rhizosphere characteristics of two arsenic hyperaccumulating Pteris ferns, Sci. Total Environ. 407 (2009) 4711–4716. [54] W.J. Fitz, W.W. Wenzel, Arsenic transformations in the soil–rhizosphere–plant system: fundamentals and potential application to phytoremediation, J. Biotechnol. 99 (2002) 259–278. [55] N. Shibagaki, et al., Selenate-resistant mutants of Arabidopsis thaliana identify Sultr1;2, a sulfate transporter required for efficient transport of sulfate into roots, Plant J. 29 (2002) 475–486. [56] M.Y. Hirai, et al., Global expression profiling of sulphur-starved Arabidopsis by DNA microarray reveals the role of O-acetyl-l-serine as a general regulator of gene expression in response to sulphur nutrition, Plant J. 33 (2003) 651–663. [57] M.L. Galeas, L.H. Zhang, J.L. Freeman, M. Wegner, E.A.H. Pilon-Smits, Seasonal fluctuations of selenium and sulphur accumulation in selenium hyperaccumulators and related nonaccumulators, New Phytol. 173 (2007) 517–525. [58] M.M. Lasat, N.S. Pence, D.F. Garvin, S.D. Abbs, L.V. Kochian, Molecular physiology of zinc transport in the Zn hyperaccumulator Thlaspi caerulescens, J. Exp. Bot. 51 (2000) 71–79. [59] X. Yang, et al., Zinc compartmentation in root, transport into xylem, and adsorption into leaf cells in the hyperaccumulating species of Sedum alfredii Hance, Planta 224 (2006) 185–195. [60] M.J. Haydon, C.S. Cobbett, Transporters of ligands for essential metal ions in plants, New Phytol. 174 (2007) 499–506. [61] D.L. Callahan, A.J.M. Baker, S.D. Kolev, A.G. Wedd, Metal ion ligands in hyperaccumulating plants, J. Biol. Inorg. Chem. 11 (2006) 2–12.

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181 [62] L. Kerkeb, U. Krämer, The role of free histidine in xylem loading of nickel in Alyssum lesbiacum and Brassica juncea, Plant Physiol. 131 (2003) 716–724. [63] R.A. Ingle, S.T. Mugford, J.D. Rees, M.M. Campbell, J.A.C. Smith, Constitutively high expression of the histidine biosynthetic pathway contributes to nickel tolerance in hyperaccumulator plants, Plant Cell 17 (2005) 2089–2106. [64] K.H. Richau, et al., Chelation by histidine inhibits the vacuolar sequestration of nickel in roots of the hyperaccumulator Thlaspi caerulescens, New Phytol. 183 (2009) 106–116. [65] S. Mari, et al., Root to shoot long-distance circulation of nicotinammine–nickel chelates in the metal hyperaccumulator Thlaspi caerulescens, J. Exp. Bot. 57 (2006) 4111–4122. [66] I.N. Talke, M. Hanikenne, U. Krämer, Zinc-dependent global transcriptional control, transcriptional deregulation, and higher gene copy number for genes in metal homeostasis of the hyperaccumulator Arabidopsis halleri, Plant Physiol. 142 (2006) 148–167. [67] V. Vacchina, et al., Speciation of nickel in a hyperaccumulating plant by high performance liquid chromatography-inductively coupled plasma mass spectrometry and electrospray ms/ms assisted by cloning using yeast complementation, Anal. Chem. 75 (2003) 2740–2745. [68] M. Becher, I.N. Talke, L. Krall, U. Krämer, Cross-species microarray transcript profiling reveals high constitutive expression of metal homeostasis genes in shoots of the zinc hyperaccumulator Arabidopsis halleri, Plant J. 37 (2004) 251–268. [69] K.B. Axelsen, M.G. Palmgren, Inventory of the superfamily of P-Type ion pumps in Arabidopsis, Plant Physiol. 126 (1998) 696–706. [70] A. Papoyan, L.V. Kochian, Identification of Thlaspi caerulescens genes that may be involved in heavy metal hyperaccumulation and tolerance. Characterization of a novel heavy metal transporting ATPase, Plant Physiol. 136 (2004) 3814–3823. [71] R.F. Mils, G.C. Krjiger, P.J. Baccarini, J.L. Hall, L.E. Williams, Functional expression of AtHMA4, a P-1B-type ATPase of the Zn/Co/Cd/Pb subclass, Plant J. 35 (2003) 164–176. [72] M. Hanikenne, et al., Evolution of metal hyperaccumulation required cisregulatory changes and triplication of HMA4, Nature 453 (2008) 391–395. [73] M. Courbot, et al., A major quantitative trait locus for cadmium tolerance in Arabidopsis halleri colocalizes with HMA4, a gene encoding a heavy metal ATPase, Plant Physiol. 144 (2007) 1052–1065. [74] G. Willems, et al., The genetic basis of zinc tolerance in the metallophyte Arabidopsis halleri ssp. halleri (Brassicaceae): an analysis of quantitative trait loci, Genetics 176 (2007) 659–674. [75] G. Willems, et al., Quantitative trait loci analysis of mineral element concentrations in the Arabidopsis halleri × Arabidopsis lyrata petraea F2 progeny on cadmium-contaminated soil, New Phytol. 187 (2010) 368–379. [76] J.E. van de Mortel, et al., Large expression differences in genes for iron and zinc homeostasis, stress response and lignin biosynthesis distinguish roots of Arabidopsis thaliana and the related metal hyperaccumulator Thlaspi caerulescens, Plant Physiol. 142 (2006) 1127–1147. [77] T.P. Durrett, W. Gassmann, E.E. Rogers, The FRD3-mediated efflux of citrate into the root vasculature is necessary for efficient iron translocation, Plant Physiol. 144 (2007) 197–205. [78] U. Krämer, I.N. Talke, M. Hannikenne, Transition metal transport, FEBS Lett. 581 (2007) 2263–2272. [79] E.P. Colangelo, M.L. Guerinot, Put the metal to the petal: metal uptake and transport throughout plants, Curr. Opin. Plant Biol. 9 (2006) 322–330. [80] D. Gendre, et al., TcYSL3, a member of the YSL gene family from the hyperaccumulator Thlaspi caerulescens, encodes a nicotinamine–Ni/Fe transporter, Plant J. 49 (2006) 1–15. [81] U Krämer, R.D. Smith, W.W. Wenzel, I. Raskin, D.E. Salt, The role of metal transport and tolerance in nickel hyperaccumulation by Thlaspi goesingense Hálácsy, Plant Physiol. 115 (1997) 1641–1650. [82] Y.H. Su, S.P. McGrath, Y.G. Zhu, F.J. Zhao, Highly efficient xylem transport of arsenite in the arsenic hyperaccumulator Pteris vittata, New Phytol. 180 (2008) 434–441. [83] G.L. Duan, Y.G. Zhu, Y.P. Tong, C. Cai, R. Kneer, Characterization of arsenate reduction in the extract of roots and fronds of Chinese brake fern, an arsenic hyperaccumulator, Plant Physiol. 138 (2005) 461–469. [84] Z. Liu, et al., Arsenite transport by mammalian aquagliceroporins AQP7 and AQP9, Proc. Natl. Acad. Sci. U.S.A. 99 (2002) 6053–6058. [85] J.F. Ma, et al., Transporters of arsenite in rice and their role in arsenic accumulation in rice grain, Proc. Natl. Acad. Sci. U.S.A. 105 (2008) 9931–9935. [86] T. Kamiya, et al., NIP1;1 an aquaporin homolog, determines the arsenite sensitivity of Arabidopsis thaliana, J. Biol. Chem. 284 (2009) 2114–2120. [87] F.J. Zhao, J.F. Ma, A.A. Meharg, S.P. McGrath, Arsenic uptake and metabolism in plants, New Phytol. 181 (2009) 777–794. [88] T.G. Sors, D.R. Ellis, D.E. Salt, Selenium uptake, translocation, assimilation and metabolic fate in plants, Photosinth. Res. 86 (2005) 373–389. [89] D.E. Salt, R.C. Prince, A.J.M. Baker, I. Raskin, I.J. Pickering, Zinc ligands in the metal accumulator Thlaspi caerulescens as determined using X-ray absorption spectroscopy, Environ. Sci. Technol. 33 (1999) 713–717. [90] D. Ueno, T. Iwashita, F.J. Zhao, J.F. Ma, Characterization of Cd translocation and identification of the Cd form in xylem sap of the Cd-hyperaccumulator Arabidopsis halleri, Plant Cell Physiol. 49 (2008) 540–548. [91] S.D. Bidwell, Hyperaccumulation of Metals in Australian Native Plants. PhD thesis, The University of Melbourne, Australia, 2001. [92] D.L. Callahan, et al., LC–MS and GC–MS metabolite profiling of nickel (II) complexes in the latex of the nickel hyperaccumulating tree Sebertia acumi-

[93]

[94]

[95]

[96]

[97]

[98]

[99]

[100]

[101] [102] [103]

[104]

[105]

[106]

[107]

[108]

[109]

[110] [111]

[112]

[113] [114]

[115]

[116]

[117]

[118]

[119] [120]

179

nata and identification of methylated aldaric acid as a new nickel (II) ligand, Phytochemistry 69 (2008) 240–251. S.D. Bidwell, S.A. Crawford, I.E. Woodrow, J. Sommer-Knudsen, A.T. Marshall, Sub-cellular localization of Ni in the hyperaccumulator, Hybanthus floribundus (Lindley) F. Muell, Plant Cell Environ. 27 (2004) 705–716. J.F. Ma, D. Ueno, F.J. Zhao, S.P. McGrath, Subcellular localisation of Cd and Zn in the leaves of a Cd-hyperaccumulating ecotype of Thlaspi caerulescens, Planta 220 (2005) 731–736. T. Asemaneh, S.M. Ghaderian, S.A. Crawford, A.T. Marshall, A.J.M. Baker, Cellular and subcellular compartmentation of Ni in the Eurasian serpentine plants Alyssum bracteatum, Alyssum murale (Brassicaceae) and Cleome heratensis (Capparaceae), Planta 225 (2006) 193–290. J.L. Freeman, et al., Spatial imaging, speciation and quantification of Se in the hyperaccumulator plants Astragalus bisulcatus and Stanleya pinnata, Plant Physiol. 142 (2006) 124–134. H. Küpper, E. Lombi, F.J. Zhao, S.P. McGrath, Cellular compartmentation of cadmium and zinc in relation to other elements in the hyperaccumulator Arabidopsis halleri, Planta 212 (2000) 75–84. B.H. Robinson, E. Lombi, F.J. Zhao, S.P. McGrath, Uptake and distribution of nickel and other metals in the hyperaccumulator Berkheya coddii, New Phytol. 158 (2003) 279–285. B. Frey, C. Keller, K. Zierold, R. Schulin, Distribution of Zn in functionally different leaf epidermal cells in the hyperaccumulator Thlaspi caerulescens, Plant Cell Environ. 23 (2000) 675–687. C.L. Broadhurst, R.L. Chaney, J.S. Angle, E.F. Erbe, T.K. Maugel, Nickel localization and response to increasing Ni soil levels in leaves of the Ni hyperaccumulator Alyssum murale, Plant Soil 265 (2004) 225–242. C. Cosio, L. De Santis, B. Frey, S. Diallo, C. Keller, Distribution of cadmium in leaves of Thlaspi caerulescens, J. Exp. Bot. 56 (2005) 565–575. J.P. Hammond, et al., A comparison of the Thlaspi caerulescens and Thlaspi arvense shoot transcriptomes, New Phytol. 170 (2006) 239–260. D. Kim, et al., The plant CDF family member TgMTP1 from the Ni/Zn hyperaccumulator Thlaspi goesingense acts to enhance efflux of Zn at the plasma membrane when expressed in Saccharomyces cerevisiae, Plant J. 39 (2004) 237–251. B.D. Dräger, et al., Two genes encoding Arabidopsis halleri MTP1 metal transport proteins co-segregate with zinc tolerance and account for high MTP1 transcript levels, Plant J. 39 (2004) 425–439. J.L. Gustin, et al., MTP1-dependent Zn sequestration into shoot vacuoles suggests dual roles in Zn tolerance and accumulation in Zn hyperaccumulating plants, Plant J. 57 (2009) 1116–1127. M.W. Persant, K. Nieman, D.E. Salt, Functional activity and role of cation-efflux family members in Ni hyperaccumulation in Thlaspi goesingense, Plant Biol. 98 (2001) 9995–10000. A.R. Cracium, et al., Comparative cDNA-AFLP analysis of Cd-tolerant and sensitive genotypes derived from crosses between the Cd hyperaccumulator Arabidopsis halleri and Arabidopsis lyrata spp. petraea, J. Exp. Bot. 57 (2006) 2967–2983. J.E. van de Mortel, et al., Expression differences for genes involved in lignin, glutathione and sulphate metabolism in response to cadmium in Arabidopsis thaliana and the related Zn/Cd-hyperaccumulator Thlaspi caerulescens, Plant Cell Environ. 31 (2008) 301–324. U. Krämer, I.J. Pickering, R.C. Prince, I. Raskin, D.E. Salt, Subcellular localization and speciation of nickel in hyperaccumulator and non-accumulator Thlaspi species, Plant Physiol. 122 (2000) 1343–1354. G. Sarret, et al., Forms of zinc accumulated in the hyperaccumulator Arabidopsis halleri, Plant Physiol. 130 (2002) 1815–1826. H. Schat, M. Llugany, R. Vooijs, J. Hartley-Whitaker, P.M. Bleeker, The role of phytochelatins in constitutive and adaptive heavy metal tolerances in hyperaccumulator and non-hyperaccumulator metallophytes, J. Exp. Bot. 53 (2002) 2381–2392. A. Raab, J. Feldman, A.A. Meharg, The nature of arsenic–phytochelatin complexes in Holcus lanatus and Pteris cretica, Plant Physiol. 134 (2004) 1113–1122. F.J. Zhao, S.J. Dunham, S.P. McGrath, Arsenic hyperaccumulation by different fern species, New Phytol. 156 (2002) 27–31. H.C. Chiang, J.-C. Lo, K.-C. Yeh, Genes associated with heavy metal tolerance and accumulation in Zn/Cd hyperaccumulator Arabidopsis halleri: a genomic survey with cDNA microarray, Environ. Sci. Technol. 40 (2006) 6792–6798. T.G. Sors, C.P. Martin, D.E. Salt, Characterization of selenocysteine methyltransferases from Astragalus species with contrasting selenium accumulation capacity, Plant J. 59 (2009) 110–122. R.S. Boyd, S.N. Martens, The raison d’être for metal hyperaccumulation by plants, in: A.J.M. Baker, J. Proctor, R.D. Reeves (Eds.), The Vegetation of Ultramafic (Serpentine) Soils, Intercept Limited, Andover, 1992, pp. 279–289. R.S. Boyd, T. Jaffré, Phytoenrichment of soil Ni content by Sebertia acuminata New Caledonia and the concept of elemental allelopathy, South Afr. J. Sci. 97 (2001) 535–538. L. Zhang, J.S. Angle, R.L. Chaney, Do high-nickel leaves shed by the nickel hyperaccumulator Alyssum murale inhibit seed germination of competing plants? New Phytol. 173 (2007) 509–516. C. Morris, P.R. Grossi, C.A. Call, Elemental allelopathy: processes, progress, and pitfalls, Plant Ecol. 202 (2009) 1–11. E.M. Jhee, R.S. Boyd, M.D. Eubanks, M.A. Davis, Nickel hyperaccumulation by Streptanthus polygaloides protects against the folivore Plutella xylostella (Lepidoptera: Plutellidae), Plant Ecol. 183 (2006) 91–104.

180

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181

[121] R.F. Jiang, D.Y. Ma, F.J. Zhao, S.P. McGrath, Cadmium hyperaccumulation protects Thlaspi caerulescens from leaf feeding damage by thrips (Frankliniella occidentalis), New Phytol. 167 (2005) 805–814. [122] S.T. Behmer, et al., Metal hyperaccumulation in plants: mechanisms of defence against herbivores, Funct. Ecol. 19 (2005) 55–66. [123] B. Rathinasabapathi, et al., Arsenic hyperaccumulator in the Chinese brake fern (Pteris vittata) deters grasshopper (Schistocerca americana) herbivory, New Phytol. 175 (2007) 363–369. [124] M.L. Galeas, et al., Selenium hyperaccumulation reduces plant arthropod loads in the field, New Phytol. 177 (2008) 715–724. [125] S.B. Huitson, M.R. Macnair, Does zinc protect the zinc hyperaccumulator Arabidopsis halleri from herbivory by snails? New Phytol. 159 (2003) 453– 459. [126] N. Noret, P. Meerts, M. Vanhaelen, A. Dos Santos, J. Escarré, Do metal-rich deter herbivores? A field test of the defence hypothesis, Oecologia 152 (2007) 92–100. [127] R.S. Boyd, The defence hypothesis of elemental hyperaccumulation: status, challenges and new directions, Plant Soil 293 (2007) 153–176. [128] R.S. Boyd, Hyperaccumulation as a plant defensive strategy, in: R.R. Brooks (Ed.), Plant that Hyperaccumulate Heavy Metals, CAB International, Oxford, UK, 1998, pp. 181–201. [129] R.S. Boyd, High-nickel insects and nickel hyperaccumulator plants: a review, Insect Sci. 16 (2009) 19–31. [130] M.A. Wall, R.S. Boyd, Melanotrichus boydi (Hemiptera: Miridae) is a specialist on the nickel hyperaccumulator Streptanthus polygaloides (Brassicaceae), Southwest Nat. 31 (2006) 481–489. [131] J.L. Freeman, C.F. Quinn, M.A. Marcus, S. Fakra, E.A.H. Pilon-Smits, Seleniumtolerant diamondback moth disarms hyperaccumulator plant defense, Curr. Biol. 16 (2006) 2181–2192. [132] A.J. Pollard, A.J.M. Baker, Deterrence of herbivory by zinc hyperaccumulation in Thlaspi caerulescens (Brassicaceae), New Phytol. 135 (1997) 655– 658. [133] R.S. Boyd, M.A. Davis, M.A. Wall, K. Balkwill, Nickel defends the South African hyperaccumulator Senecio coronatus (Asteraceae) against Helix aspersa (Mullusca: Pulmonidae), Chemoecology 12 (2002) 91–97. [134] B. Hanson, S.D. Lindblom, M.L. Loeffler, E.A.H. Pilon-Smits, Selenium protects plants from phloem-feeding aphids due to both deterrence and toxicity, New Phytol. 162 (2004) 655–662. [135] M.A. Davis, R.S. Boyd, Dynamics of Ni-based defence and organic defences in the Ni hyperaccumulator Streptanthus polygaloides Gray (Brassicaceae), New Phytol. 146 (2000) 211–217. [136] R.O. Tolrà, C. Poschenrieder, R. Alonso, D. Barceló, J. Barceló, Influence of zinc hyperaccumulation on glucosinolates in Thlaspi caerulescens, New Phytol. 151 (2001) 621–626. [137] E.M. Jhee, R.S. Boyd, M.D. Eubanks, Effectiveness of metal–metal and metal–organic compounds combinations against Plutella xylostella (Lepidoptera: Plutellidae): implication for plant elemental defence, J. Chem. Ecol. 32 (2006) 239–259. [138] F. Navari-Izzo, M.F. Quartacci, Phytoremediation of metals: tolerance mechanisms against oxidative stress, Minerva Biotecnol. 13 (2001) 23– 83. [139] S.P. McGrath, F.J. Zhao, Phytoextraction of metals and metalloids from contaminated soils, Curr. Opin. Biotechnol. 14 (2003) 277–282. [140] E. Pilon-Smits, Phytoremediation, Annu. Rev. Plant Biol. 56 (2005) 15–39. [141] P. Kidd, et al., Trace element behaviour at the root–soil interface: implications in phytoremediation, Environ. Exp. Bot. 67 (2009) 243–259. [142] J.P. Schwitzuébel, J. Kumpiene, E. Comino, T. Vanek, From green to clean: a promising approach towards environmental remediation and human health for the 21st century, Agrochimica, LIII-N 4 (2009) 209–237. [143] R.L. Chaney, Plant uptake of inorganic waste constitutes, in: J.F. Parr, P.B. Marsh, J.M. Kla (Eds.), Land Treatment of Hazardous Wastes, Noyes Data Corp., Park Ridge, 1983, pp. 50–76. [144] S.D. Cunningham, W.R. Berti, J.W. Huang, Phytoremediation of contaminated soils, Trends Biotechnol. 13 (1995) 393–397. [145] S.D. Ebbs, et al., Heavy metals in the environment: phytoextraction of cadmium and zinc from a contaminated soil, J. Environ. Qual. 26 (1997) 1424–1430. [146] S.L. Brown, R.L. Chaney, J.S. Angle, A.J.M. Baker, Zinc and cadmium uptake of Thlaspi caerulescens grown in nutrient solution, Soil Sci. Soc. Am. J. 59 (1995) 125–133. [147] B.H. Robinson, M. Leblanc, D. Petit, The potential of Thlaspi caerulescens for phytoremediation of contaminated soils, Plant Soil 203 (1998) 47– 56. [148] U. Krämer, J.D. Cotter-Howells, J.N. Charnock, A.J.M. Baker, J.A.C. Smith, Free histidine as a metal chelator in plants that accumulate nickel, Nature 379 (1996) 635–638. [149] H. Tamura, M. Honda, T. Sato, H. Kamachi, Pb hyperaccumulation and tolerance in common buckwheat (Fagopyrum esculentum Moench), J. Plant Res. 118 (2005) 355–359. [150] S.G. Xue, Manganese uptake and accumulation by the hyperaccumulator plant Phytolacca acinosa Roxb (Phytolaccaceae), Environ. Pollut. 131 (2004) 393–399. [151] P. Kidd, C. Monterroso, Metal extraction by Alyssum serpyllifolium spp. lusitanicum on mine-spoil soils from Spain, Sci. Total Environ. 336 (2005) 1–11. [152] S.P. McGrath, C.M.D. Sidoli, A.J.M. Baker, R.D. Reeves, The potential for the use of metal-accumulating plants for the in situ decontamination of metal-

[153]

[154]

[155]

[156]

[157]

[158]

[159]

[160]

[161]

[162]

[163]

[164]

[165] [166]

[167]

[168]

[169]

[170]

[171]

[172]

[173]

[174]

[175]

[176] [177] [178]

polluted soils, in: H.J.P. Eijsackers, T. Hamers (Eds.), Integrated Soil and Sediment Research: A Basis for Proper Protection, Kluwer Academic Publ., Dordrecht, The Netherlands, 1993, pp. 673–676. S.L. Brown, R.L. Chaney, J.S. Angle, A.J.M. Baker, Phytoremediation potential of Thlaspi caerulescens and bladder campion for zincand cadmium-contaminated soil, J. Environ. Qual. 23 (1994) 1151– 1157. P. McGrath, et al., Field evaluation of Cd and Zn phytoextraction potential by the hyperaccumulators Thlaspi caerulescens and Arabidopsis halleri, Environ. Pollut. 141 (2006) 115–125. J. Yanai, F.J. Zhao, S.P. McGrath, T. Kosaki, Effect of soil characteristics on Cd uptake by the hyperaccumulator Thlaspi caerulescens, Environ. Pollut. 139 (2006) 167–175. F.J. Zhao, E. Lombi, S.P. McGrath, Assessing the potential for Zn and cadmium phytoremediation with hyperaccumulator Thlaspi caerulescens, Plant Soil 249 (2003) 37–43. S.P. McGrath, S.J. Dunham, R.L. Correll, Potential for phytoextraction of zinc and cadmium from soils using hyperaccumulator plants, in: N. Terry, G. ˜ Banuelos (Eds.), Phytoremediation of Contaminated Soil and Water, Lewis Publishers, Boca Raton, FL, 2000, pp. 109–128. B.H. Robinson, et al., The nickel hyperaccumulator plant Alyssum bertolonii as a potential agent for phytoremediation and phytomining of nickel, J. Geochem. Explor. 59 (1997) 75–86. B.H. Robinson, et al., The potential of the high-biomass nickel hyperaccumulator Berkheya coddii for phytoremediation and phytomining, J. Geochem. Explor. 60 (1997) 115–128. G.M. Kertulis-Tartar, L.Q. Ma, C. Tu, T. Chirenje, Phytoremediation of an arsenic-contaminated site using Pteris vittata L.: a two-year study, Int. J. Phytoremed. 8 (2006) 311–322. A. Salido, K.L. Hasty, J.M. Lin, D.J. Butcher, Phytoremediation of arsenic and lead in contaminated soil using Chinese brake fern (Pteris vittata) and Indian mustard (Brassica juncea), Int. J. Phytoremed. 5 (2003) 89– 103. R.R. Brooks, B.H. Robinson, The potential use of hyperaccumulators and other plants in phytomining, in: R.R. Brooks (Ed.), Plants that Hyperaccumulate Heavy Metals: Their Role in Phytoremediation, Microbiology, Archaeology, Mineral Exploration and Phytomining, CAB International, Wallingford, UK, 1998, pp. 327–356. A.C. Singer, T. Bell, C.A. Heywood, J.A.C. Smith, I.P. Thompson, Phytoremediation of mixed-contaminated soil using the hyperaccumulator plant Alyssum lesbiacum: evidence of histidine as a measure of phytoextractable nickel, Environ. Pollut. 147 (2007) 74–82. E.P. Brewer, J.A. Saunders, J.S. Angle, R.L. Chaney, M.S. McIntosh, Somatic hybridization between the zinc accumulator Thlaspi caerulescens and Brassica napus, Theor. Appl. Genet. 99 (1999) 761–771. D. Gleba, et al., Use of plant roots for phytoremediation and molecular farming, Proc. Natl. Acad. Sci. U.S.A. 96 (1999) 5973–5977. E. Lombi, F.J. Zhao, S.J. Dunham, S.P. McGrath, Phytoremediation of heavy metal contaminated soils: natural hyperaccumulation versus chemically enhanced phytoextraction, J. Environ. Qual. 30 (2001) 1919– 1926. C.G. Kawashima, et al., Heavy metal tolerance of transgenic plants over-expressing cysteine synthase, Biotechnol. Lett. 26 (2004) 153– 157. L.E. Bennet, et al., Analysis of transgenic Indian mustard plants for phytoremediation of metal-contaminated mine tailings, J. Environ. Qual. 32 (2003) 432–440. J. Lee, et al., Functional expression of a bacterial heavy metal transporter in Arabidopsis enhances resistance to and decreases uptake of heavy metals, Plant Physiol. 133 (2003) 589–596. A. Berken, M.M. Mulholland, D.L. LeDuc, N. Terry, Genetic engineering of plants to enhance Selenium phytoremediation, Crit. Rev. Plant Sci. 21 (2002) 567–582. J. Lee, et al., Over-expression of Arabidopsis phytochelatin synthase paradoxically leads to hypersensitivity to cadmium stress, Plant Physiol. 131 (2003) 656–663. W.Y. Song, et al., A novel family of cys-rich membrane proteins mediates cadmium resistance in Arabidopsis, Plant Physiol. 135 (2004) 1027– 1039. F. Verret, et al., Over-expression of AtHMA4 enhances root-to-shoot translocation of zinc and cadmium and plant metal tolerance, FEBS Lett. 576 (2004) 306–312. E. Pilon-Smits, et al., Over-expression of ATP sulfurylase in Brassica juncea leads to increased selenate uptake, reduction and tolerance, Plant Physiol. 119 (1999) 123–132. T. Van Huysen, N. Terry, E.A.H. Pilon-Smits, Exploring the selenium phytoremediation potential of transgenic Indian mustard over-expressing ATP sulfurylase or cystathionine-␥-synthase, Int. J. Phytoremed. 6 (2004) 111–118. F.J. Zhao, S.P. McGrath, Biofortification and phytoremediation, Curr. Opin. Plant Biol. 12 (2009) 373–380. R.R. Brooks, M.F. Chambers, L.J. Nicks, B.H. Robinson, Phytomining, Trends Plant Sci. 1 (1998) 359–362. L.J. Nicks, M.F. Chambers, A pioneering study of the potential of phytomining for nickel, in: R.R. Brooks (Ed.), Plants that Hyperaccumulate Heavy Metals, CAB International, 1998, pp. 313–326.

N. Rascio, F. Navari-Izzo / Plant Science 180 (2011) 169–181 [179] C. Minguzzi, O. Vergnano, Il contenuto in nickel nelle ceneri di Alyssum bertolonii Desv, Mem. Soc. Tosc. Sci. Nat. A55 (1948) 49–77. [180] C.W.N. Anderson, et al., Phytomining for nickel, thallium and gold, J. Geochem. Explor. 67 (1999) 407–415. [181] L.J. Nicks, M.F. Chambers, Farming for metals, Mining Environ. Manage. (1995) 15–18.

181

˜ [182] B.H. Robinson, G. Banuelos, H.M. Conesa, M.W.H. Evangelou, R. Schulin, The phytomanagement of trace elements in soil, Crit. Rev. Plant Sci. 28 (2009) 240–266. [183] D.R. Ellis, D.E. Salt, Plant, selenium and human health, Curr. Opin. Plant Biol. 6 (2003) 273–279. [184] M.G. Palmgren, et al., Zinc biofortification of cereals: problem and solutions, Trends Plant Sci. 13 (2008) 464–473.