Next Article in Journal
Structural Shifts in the Soil Prokaryotic Communities Marking the Podzol-Forming Process on Sand Dumps
Next Article in Special Issue
Nickel Effects on Growth and Phytolith Yield of Grasses in Contaminated Soils
Previous Article in Journal
Selenium and Heavy Metals in Soil–Plant System in a Hydrogeochemical Province with High Selenium Content in Groundwater: A Case Study of the Lower Dniester Valley
Previous Article in Special Issue
Complex Speciation and Distribution of Iron, Sulfur, and Trace Metals in Coal Mine Soils Reflect Grain- and Sub-Grain-Scale Heterogeneity during Pyrite Oxidative Dissolution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Soil Phytomining: Recent Developments—A Review

by
Christos Kikis
,
Georgios Thalassinos
and
Vasileios Antoniadis
*
Department of Agriculture, Crop Production and Rural Environment, University of Thessaly, Fytokou Street, 38446 Volos, Greece
*
Author to whom correspondence should be addressed.
Soil Syst. 2024, 8(1), 8; https://doi.org/10.3390/soilsystems8010008
Submission received: 16 November 2023 / Revised: 14 December 2023 / Accepted: 4 January 2024 / Published: 8 January 2024
(This article belongs to the Special Issue Research on Heavy Metals in Soils and Sediments)

Abstract

:
Phytomining (PM) is defined as the process of using plants capable of bio-extracting metals from soil in order to explore them economically. This relatively new, innovative method has been gathering significant attention in both the academic and commercial domains. Conventional mining methods are often economically unviable when applied to lean ores, and they can lead to secondary pollution in soil—a situation that applies to all excavated metals. On the other hand, PM is an environmentally friendly and economically viable solution that addresses the growing demands for metal resources, while simultaneously contributing to energy production by harnessing biomass energy. This comprehensive review presents the current PM techniques, challenges, and the hyperaccumulator plant species that may be used for the extraction of the main targeted elements in the process. Typically, the targeted metals are those of economic value, which can later be deposited or sold to various industries. This review also analyzes the factors influencing the economic viability of PM and proposes potential enhancements. Undeniably, PM offers the opportunity for economically sustainable exploration of metal-rich soils, but its full commercial viability remains constrained under current conditions as scientists are actively searching for the identification and utilization of new hyperaccumulator plant species in different locations worldwide, while creating new relationships and business avenues within the mining industry. Overall, this review highlights the current status of PM technology and the plants used, emphasizing the need for further research to enhance its commercial implementation and its potential to assist the mining industry. We conclude that PM, although a relatively new and unexplored concept, may provide economic and environmental benefits to soil end-users and managers who must cultivate on metal-contaminated soils as PM may turn yield shortages (of specific commercial crops) to benefits if high-yield hyperaccumulators are cultivated for industrial valorization of their high metal-content biomass.

1. Introduction

Phytoremediation is the process of remediating (or cleaning up) a soil elevated with pollutants with the use of plants (“phyto”). Many endemic plants, also known as metallophytes, can be found to be naturally grown in soils with elevated metal concentrations [1]. These unique plant species, also called “hyperaccumulators”, have portrayed natural tolerance to the presence of metals in soil or an enhanced ability for their accumulation. These plants might be genotypes that have evolved such tolerant characteristics due to their establishment in metalliferous areas for centuries or millennia [2]. The accumulation of metals in plants occurs in their root system or in the aerial biomass like stems, leaves, and flowers. According to Soleymanifar et al. [3], a plant capable of concentrating a metal significantly higher than a “regular” plant in its tissues when grown in the same area is characterized as a hyperaccumulator. Hyperaccumulators have the potential of absorbing metals from soil via their root system and translocating them into their aerial parts [4]. In this way, a contaminated soil may be phytoextracted. Thus, phytoextraction is the process of using hyperaccumulator plants to absorb potentially toxic elements from soil with the aim of remediating the soil. A variation in phytoextraction is phytomining (PM). Phytomining is the process under which plants are used to extract (“mine”) valuable inorganic elements from soil in a natural or in an induced manner for the specific purpose of financial gain after valorizing the produced biomass. The difference between PM and phytoextraction is the valorization of the plant biomass; in the former case, it is solely dedicated to being used for recovering the metals for use in the metallurgical industry. In the latter, it can have any uses except from that—pharmaceutical, aromatic, energy, biochar, etc. Conventional mining is only performed in ore deposits where metal contents are highly enriched (i.e., soil concentrations are well above the maximum legal regulation limits). Nevertheless, these kinds of deposits are being depleted because of increased population and emerging needs, as well as rapid industrial and economic advancement [5]. Even though traditional mining methods are successful, they cannot recover all of the valuable metals. After the completion of the mining process, there is still about 3% to 10% of metals left in the soil [5]. These areas are characterized as “degraded” due to the fact that they contain lower concentrations of metals and minerals than what would be economically explorable [6]. Usually, such ore bodies are related to soils formed on ultramafic and serpentine parent materials. During the cool-down of ultramafic magma, minerals such as olivine and pyroxene, rich in magnesium (Mg) and iron (Fe), have a higher melting point and therefore crystallize at higher temperatures, leading to their dominance, along with some elements like nickel (Ni) and chromium (Cr), in ultramafic rocks. These types of soils are rich in various metals such as Mg, Fe, Ni, Cr, manganese (Mn), cobalt (Co), and titanium (Ti) [7,8].
In order to facilitate the phytoaccumulation, metals can be dissolved from the ore by applying strong acids or with the technique of electrorefining, in which metal ions gain electrons and form soil metal deposits at the electrode’s surface [9]. In this way, metals in soil may become more phytoavailable, and thus their uptake to hyperaccumulator plants may be increased and, in turn, PM may occur. Hence, instead of going through the usual procedures of soil excavation for mining, the extractions are performed with the use of hyperaccumulator plants.
The concept of metal accumulation in plants and PM was first introduced by Jaffré et al. [10]. To date, scientists also refer to the process as “agromining” [11], in order to encompass the entire agricultural system of ore–soil–plants. The whole process offers the opportunity to extract metals from sites where conventional mining methods are not economically viable. The environmental effect of the method is minimal due to the balance action of the plants opposing soil erosion caused by excavation [12]. This method offers a sustainable way to remediate or restore soil, while simultaneously recovering valuable metals from soil [13]. Conventional methods used to remediate and decontaminate polluted sites often result in secondary pollution, aggravating the environmental challenges. The presence of plants minimizes the spread of metals and reduces the risk of wind and water soil erosion. Moreover, in large-scale PM applications, there is the possibility to harness the energy stored within the harvested plant biomass [3].
The bioaccumulation factor (BAF) and translocation factor (TF) are used to determine the efficiency of metal uptake from soil to plant and metal translocation within the plant, respectively [14]. The BAF is an index obtained by dividing the concentration of a metal in the aerial parts of a plant by its concentration in the surrounding soil. This factor provides insights into the ability of the plant to absorb and accumulate metals from the soil. On the other hand, TF is the ratio of metal concentration in the aerial plant parts to that in the roots. Higher TF values indicate a greater ability to efficiently translocate metals from the roots to the above-ground parts of a plant [14]. To be classified as a hyperaccumulator, a plant must have both BAF and TF values exceeding 1.0, signifying a great ability to absorb metals from soil and translocate them upwards in the aerial parts [15]. These indicators can be influenced by the effect of metal concentration in the plant. Therefore, biomass production and metal availability in the soil or substrate must be taken into consideration.
The aim of this review was (a) to provide a comprehensive overview on the current PM status, including challenges associated with the method, (b) to analyze the economic viability of PM as an eco-friendly alternative to conventional mining, and (c) to discuss PM technology and the plants used, highlighting the importance of further research and development to enhance large scale implementation. We see this contribution as a highly necessary input so that essential knowledge and technological and scientific advancements may be updated and the community of scientists, interested stakeholders, and other soil end-users may benefit from the elucidation of mechanisms, processes, and practical considerations.

2. Elements Involved in Phytomining

2.1. Phytomining Elements

The elements of interest in the PM process are Ni, Co, cadmium (Cd), zinc (Zn), selenium (Se), Mn, thallium (Tl), and the noble metals (NMs). However, the majority of research efforts have focused on advancing the PM process of Ni [16]. For each metal, there is an established threshold that needs to be satisfied in order for a plant to be acknowledged as a hyperaccumulator. The selection of the limits is not entirely random but instead an indicator of a unique form of plant response, suggesting the presence of an unusual behavior. For instance, Sheoran et al. [5] suggested a limit of 1000 mg kg−1 of dry plant weight (DW) for Ni, after tests conducted on Homalium and Hybanthus plants in different areas. Similarly, Baker and Brooks [17] set the limit to 10,000 mg kg−1 DW for Mn and Zn after reporting high Zn concentrations in Thlaspi calaminare and Viola calaminaria leaves. Generally, over 700 plant species have been identified for their PM capabilities, most of which are Ni hyperaccumulators. The established limits for most of the elements that demonstrate an interest for being phytomined are shown in Table 1. When selecting plants for metal extraction, certain characteristics are important; these include three important features: high biomass yield, high Ni concentration in shoots (>1%), and fast growth [18]. The reason is that the phytoextracted amount of any given metal is equal to the concentration times biomass yield. All discussion is based on results obtained from field experiments, unless otherwise stated per particular case.
The hyperaccumulation threshold was originally established as a metal concentration in plant shoots 10 to 100 times higher when compared to measurements in the shoots of “normal” plants under similar conditions. It was redefined by Van der Ent et al. [18], who suggested that the initial concentrations should be at least 10 times higher than the measurements obtained in plants on metalliferous soils, and 100 to 1000 times higher in plants grown on regular soils (non-metal-enriched). Typically, most of the hyperaccumulating plant species are endemic to metalliferous soils. Since the root system of these plants is not usually collected during harvest, it is also mandatory for hyperaccumulators to exhibit BAF and TF values over 1, indicating that metals can be relocated in the upper part of the plant and recovered through combustion [14].

2.1.1. Nickel

As already mentioned, soils developed on ultramafic materials are expected to contain large amounts of Ni; hence, they are considered prime cases for Ni PM. Even though Ni in soil can be retained by Fe and Mn oxides, Ni hyperaccumulators can still accumulate a sufficient amount of the metal until the depletion of the source. According to Nkrumah et al. [16], 15–50 years are required before all easily accessible Ni is depleted. In addition, the physical properties of the soil are very important when it comes to PM. Well-drained soils are optimum for the cultivation of Ni hyperaccumulators, while soils that do not drain well may lead to decreased amount of plant biomass that is produced and, therefore, to a decreased amount of Ni that is extracted. For successful and economically viable Ni PM, soils should have an optimum content of moisture to maximize the amount of Ni absorbed by hyperaccumulator plants [21]. In the future, plant breeding techniques can make PM operations more profitable. Table 2 shows the main Ni hyperaccumulator plants that have great potential in Ni extraction from soils. Most of the plants that can hyperaccumulate Ni are species from approximately 40 different families, with the most important representatives being the Brassicaceae, Euphorbiaceae, Phyllanthaceae, Salicaceae, Buxaceae, and Rubiaceae family groups. Most of them are endemic plants found in soils derived from ultramafic rocks. The first report for Ni hyperaccumulation was made by Minguzzi and Vergnano [22], but the work was not widely recognized. Jaffré et al. [10] were the first to establish the term “phytomining” after noticing the accumulation of Ni in Sebertia accuminata, an endemic tree in forests of New Caledonia. Chaney et al. [23] were the first to conduct a field experiment for PM in the US Bureau mines in Reno, Nevada, where a Ni hyperaccumulator Streptanthus polygaloides was found to have bioaccumulated 10,000 mg kg−1 DW in serpentine soils containing 3500 mg kg−1 Ni. The second field trial was a two-year research study conducted by Robinson et al. [24], who extracted up to 8000 mg kg−1 DW Ni and had 110,000 mg kg−1 Ni in ash in Alyssum bertolonii. They also planted Berkheya coddii, which accumulated a total of 5500 mg kg−1 DW. It was also reported that B. coddii was a very good candidate for Ni extraction due to its high biomass yield, perennial life cycle, and tolerance to abiotic stress [24].
Numerous research studies have reported elevated concentrations of Ni in plant tissues when grown in these types of soils. All Ni PM research has focused on plants that can accumulate amounts more than 1% (i.e., 10,000 mg kg−1) in their leaves and the harvested biomass, after collecting small plant samples (1 cm2 or less) from serpentine soils in California. Indeed, Reeves et al. [25] found that Streptanthus polygaloides was able to accumulate Ni up to 1.5%, which led Mohsin et al. [26] to mention this plant as a good option for PM. Cole [27] reported the uptake of 5000 mg kg−1 Ni in the dried leaves of Hybanthus floribundus in lateritic soils in Western Australia, where background levels were approximately 500 mg kg−1 in the surface soil. In another study, Li et al. [7] reported a concentration of up to 22,000 mg kg−1 that was absorbed by Alyssum murale and Alussum corsicum plants (Brassicaceae), grown in Ni-contaminated soils in a 120-day greenhouse experiment. In addition, plants were fertilized and the pH was adjusted. Recently, Alyssum hyperaccumulators have received wide interest regarding their phytoextraction potential [16]. Durand et al. [28] conducted a 7-month pot experiment using Odontarrhena chalcidica (syn. Alyssum murale) spiked with Ni sulfate. They reported increased metal availability as well as BAF and TF values.
Table 2. Ni hyperaccumulation plant species. The accumulation of the elements refers to vegetative plant parts which can be harvested and valorized for the contained metal.
Table 2. Ni hyperaccumulation plant species. The accumulation of the elements refers to vegetative plant parts which can be harvested and valorized for the contained metal.
Plant SpeciesFamilyNi Concentration (mg kg−1)/Biomass Yield (t ha−1)Experimental Conditions—CountryReference
Alyssum baldacciiBrassicaceae1430–17,670Ex situ; plant analysis; serpentine soils; Mediterranean climate; Mt. Smolikas, GreeceBani et al. [29]
Alyssum bertoloniiBrassicaceae4653–13,102/7–9Hydroponics experiment; contained 400 mL Hoagland solution; plants spiked with NiSO4; ItalyGalardi et al. [30]
Alyssum bertoloniiBrassicaceae800/92-year field experiment; applied various amounts of NPK fertilizer; Mediterranean climate; Murlo, ItalyRobinson et al. [24]
Alyssum chalcidicumBrassicaceae1700Ex situ; plant analysis; ultramafic rocks and hot springs; pseudo-total Ni concentration 1560 mg kg−1 soil; Mediterranean climate; Euboea Island, GreeceKanellopoulos and Argyraki [31]
Alyssum euboeumBrassicaceae26–4550Ex situ; plant analysis; serpentine soils; Mediterranean climate; TurkeyBani et al. [29]
Alyssum heldreichiiBrassicaceae11,800Ex situ; plant analysis; serpentine soils; pseud-total Ni concentrations 1160 mg kg−1 soil; Mediterranean climate; Katara pass, GreeceBani et al. [29]
Alyssum lesbiacumBrassicaceae2900–22,400/4.2–10Hydroponics experiment; applied nutrient solution; solution spiked with 0–250 μmol L−1 NiSO4; GreeceAdamidis et al. [32]
Alyssum markgrafiiBrassicaceae<3–19,100Ex situ; plant analysis; serpentine soils; pseudo-total Ni concentration in soil 1670 mg kg−1; Mediterranean climate; Gjegjan; AlbaniaBani et al. [29]
Alyssum muraleBrassicaceae7–34,690/10.2–205-year field experiment; pseudo-total Ni concentration in soil 3180 mg kg−1; fertilized with 100 kg ha−1 N; Mediterranean climate; Pojska, AlbaniaBani et al. [29]
Berkeya coddiiAsteraceae40–5800Pot experiment in glasshouses; plants spiked with 4−1000 mg Ni kg−1 potting mix; New ZealandKeeling et al. [33]
Brackenridgea palustrisMelastomataceae1440/6.74Ex situ; plant analysis; ultramafic rocks; pseudo- total Ni concentration—7051–10,521 mg kg−1; Tropical climate; Sulawesi, IndonesiaVan der Ent et al. [34]
Cloezia artensisMyrtaceae35–1729Field experiment; extractable Ni concentrations in soil—35–208 mg kg−1; Semi-Tropical climate; New CaledoniaBecquer et al. [35]
Codia spatulataCunoniaceae27–1420/6.5Field experiment around mine sites; Semi-Tropical climate; New CaledoniaJaffré et al. [36]
Dichapetalum gelonioidesDichapetalaceae3160Ex situ; plant analysis; ultramafic rocks; pseudo-total Ni concentrations—7051–10,251 mg kg−1 soil; Tropical climate; Sulawesi, IndonesiaVan der Ent et al. [34]
Geissois PruinosaCunoniaceae3000–15,106/38.5Field experiment around mine sites; Semi-tropical climate; New CaledoniaJaffré et al. [36]
Knema matanensisMyristicaceae2500–5000Ex situ; plant analysis; ultramafic rocks; pseudo total Ni concentrations—7051–10,251 mg kg−1 soil; Tropical climate; Sulawesi, IndonesiaVan der Ent et al. [34]
Miscanthus floridulusPoaceae667–778Ex situ; plant analysis; uranium tailings with background Ni concentration 11.5 mg kg−1 soil; Tropical climate; South ChinaLi et al. [37]
Myristica laurifoliaMyristicaceae1100Ex situ; plant analysis; ultramafic rocks; pseudo-total Ni concentrations—7051–10,251 mg kg−1; Tropical climate; Sulawesi, IndonesiaVan der Ent et al. [34]
Peripterygia marginataAraliaceae23–1800Field experiment around mine sites; Semi-Tropical climate; New CaledoniaJaffré et al. [36]
Phyllanthus discolorPhyllanthaceae13,670–31,490Ex situ; plant analysis; ultramafic rocks; Tropical climate; CubaBerazain et al. [38]
Phyllanthus orbicularisPhyllanthaceae4140–10,950Ex situ; plant analysis; ultramafic rocks; Tropical climate; CubaBerazain et al. [38]
Phyllanthus xpallidusPhyllanthaceae15,390–60,170Ex situ; plant analysis; ultramafic rocks; Tropical climate; CubaBerazain et al. [38]
Phyllantus insulaePhyllanthaceae34,330–38,720Ex situ analysis; ultramafic rocks; pseudo-total Ni concentrations—7051–10,251 mg kg−1; Tropical climate; Sulawesi, IndonesiaVan der Ent et al. [34]
Planchonella oxyhedraSapotaceae19,600/10Ex situ; plant analysis; ultramafic rocks; pseudo-total Ni concentrations—7051–10,251 mg kg−1 soil; Tropical climate; Sulawesi, IndonesiaVan der Ent et al. [34]
Planchonella oxyhedraSapotaceae24,100–27,500/30–40Ex situ; plant analysis; ultramafic rocks; Tropical climate; Halmahera Island, IndonesiaHamdan et al. [39]
Rinorea bengalensisViolaceae20,000–25,100/30–40Ex situ; plant analysis sampling; ultramafic rocks; Tropical climate; Halmahera Island, IndonesiaHamdan et al. [39]
Sarcotheca celebicaOxalidaceae700–1000Ex situ; plant analysis; ultramafic rocks; pseudo-total Ni concentrations—7051–10,251 mg kg−1 soil; Tropical climate; Sulawesi, IndonesiaVan der Ent et al. [34]
Streptanthus polygaloidesBrassicaceae4000/21.4Ex situ; plant analysis; serpentine soil; California, USADavis and Boyd [40]
Thlaspi apterumBrassicaceae16,600–21,500Ex situ; plant analysis; serpentine soils; pseudo-total Ni concentrations 2858 mg kg−1 soil; Mediterranean climate; Fotinovo, BulgariaBani et al. [29]
Thlaspi ochroleucumBrassicaceae3400Ex situ; plant analysis; serpentine soils; pseudo-total Ni concentrations 2770 mg kg−1 soil; Mediterranean climate; Kardzali, BulgariaBani et al. [29]
Thlaspi tymphaeumBrassicaceae7000Ex situ; plant analysis; serpentine soils; pseudo-total Ni concentrations 2340 mg kg−1 soil; Mediterranean climate; Malakasi, GreeceBani et al. [29]
Trichospermum kjelbergiiMalvaceae3770Ex situ; plant analysis; ultramafic rocks; pseudo-total Ni concentrations—7051–10,251 mg kg−1 soil; Tropical climate; Sulawesi, IndonesiaVan der Ent et al. [34]

2.1.2. Cobalt

The standard for plants to be classified as Co hyperaccumulators was set at 1000 mg kg−1 by Baker and Brooks [17] until Krämer [19] revised it to 300 mg kg−1, reporting that the former accumulation of Co is rare. Normal amounts of Co in plants are well below 0.1 mg kg−1. Even in soils characterized with elevated amounts of Co, such as those deriving from ultramafic rocks, the amount of Co in plants rarely surpasses 20 mg kg−1. It seems that the presence of Ni in such soils interferes with Co accumulation. For instance, as mentioned before, Berkheya coddii is also a Ni hyperaccumulator. The presence of Co can inhibit the absorption of Ni when these metals are both present in the medium, leading to PM limitations. Keeling et al. [33] reported that Co can be absorbed by B. coddii plants with or without the presence of Ni in the growing medium (50% peat and 50% pumice). On the other hand, increased Co concentrations in soil from 125 up to 1000 mg kg−1 significantly decreased the harvested biomass due to plant toxicity, although the plant’s BAF was not affected. In trials where only Ni (at 1000 mg kg−1) or only Co (at 4–1000 mg kg−1) were present in soil, the crop yielded approximately 14.5 kg ha−1 (when for Ni) and 12.6 kg ha−1 (for Co). On the other hand, when Co and Ni were both present in soil (with concentrations of Ni = 500 and Co = 500 mg kg−1), the metal yield significantly dropped to 0.9 kg of Ni ha−1 and 3.9 kg of Co ha−1, and this indicates that PM may be limited due to interference of each other’s absorption by the plant. Similar to this, Rue et al. [41] suggested that B. coddii could be utilized in areas with high concentrations of Co but low Ni after testing the plant in pot experiments. Specifically, they found that the Co concentration reached 1980 mg kg−1 when treated with 10 mg kg−1 Ni (as Ni(NO3)2·6H2O) and 100 mg kg−1 Co (as Co(NO3)2·6H20), while total biomass can reach up to 22 t ha−1, which could yield 16.3 t Co ha−1. Similar findings were reported by Parks [42], who noticed the competitive role between Ni and Co accumulation in Rinorea bengalensis. However, R. bengalensis was found to be tolerant to Ni and managed to accumulate a concentration of 1200 mg Co kg−1 when grown on ultramafic rocks. Both Ni and Co were added as Ni(NO3)2·6H2O and Co(NO3)2·6H20 at 750 mg kg−1, respectively. An exceptional ability for Co PM was shown after collecting leaf samples of Nyssa sylvatica var. biflora and var. sylvatica (Nyssaceae) grown on ultramafic rocks in New Caledonia, which were found to accumulate up to 845 mg Co kg−1 [43]. The same was reported for the well-known Ni hyperaccumulator Alyssum murale, which accumulated 1320 mg Co kg−1 DW when grown in sandy loam soils (<2 mg Co kg−1 dry soil) spiked with 59.8 mg (1 mmol) Co kg−1 dry soil [44].
The addition of elemental zero-valent sulfur in soil can help increase the amount of metal that plants can absorb. According to Robinson et al. [45], sulfur content in soil at a rate of 5000 mg kg−1 increased the amount of Co that was absorbed by Berkheya coddii plants to 299 mg kg−1 DW, 5 times greater than the control treatment (56 mg kg−1 DW). Hence, this is especially beneficial in soils that do contain some amount of metals but not enough for the plants to take up the maximum quantity they need. In addition, sulfur is a cost–benefit additive, compared to EDTA, ethylenediaminetetraacetic acid, a synthetic molecule frequently used to increase metal mobility and soil, for enhancing the growth of metal-rich plants in soils with low metal concentrations. Overall, plant species that have demonstrated Co hyperaccumulation are shown in Table 3.

2.1.3. Cadmium

Higher concentrations of cadmium (Cd), ranging from 10 to 200 mg kg−1, can be found in soils that have been exposed to waste materials due to Zn mining. In addition, such elevated Cd levels may also arise in soils that have received industrial waste or phosphate fertilizers enriched with Cd [7]. Natural levels of Cd in plants are usually less than 0.1 mg kg−1, but in Cd-contaminated soils these levels can exceed 20 mg kg−1. The cadmium established threshold in plants is defined as having a concentration of higher than 100 mg kg−1, as proposed by Van der Ent et al. [18]. However, in certain Zn-Pb mines, Noccaea species, such as N. caerulescens and N. praecox, have been found to contain Cd concentrations exceeding 100 mg kg−1, or even higher than 1000 mg kg−1. Shoots of N. caerulescens can easily contain over 2000 mg Cd kg−1, even when grown on contaminated soils, and over 20,000 mg kg−1 when tested in nutrient solutions [56]. Other plants like Impatiens walleriana, Pteris vittata, Sedum alfredii, and Thlaspi caerulescens are capable of removing from soil 1168, 6434, 922.6, and 7400 mg kg−1 of Cd, respectively [57]. Specifically, Nedelkoska et al. [58] treated the roots of T. caerulescens with a H+-ATPase inhibitor and measured a concentration of 62,800 mg Cd kg−1 DW (6.3%) in the hair roots of the plant, which was grown in nutrient medium amended with a Cd concentration of 3710 ppm. On the other hand, there are significant variations among sites and within these plants [59]. Similar findings have been reported for Arabidopsis halleri by Bert et al. [60] and Viola baoshanensis by Deng et al. [61]. Specifically, A. halleri has been reported for its great ability to grow in different environments and its capability to store high amounts of Cd, especially when grown in polluted soils. According to Claire-Lise and Nathalie [62], A. halleri was able to accumulate up to 157 mg kg−1 leaf dry weight when treated with 5 μM Cd or more. Also, it has been used as a plant model on metal tolerance and accumulation for identifying the genes involved. However, it has not been further tested due to the fact that it produces low biomass. Nevertheless, it was noted that for a plant to be considered as a Cd hyperaccumulator, it needs to be naturally grown in an area and sustain itself without human intervention [18]. The list of Cd hyperaccumulating species are shown in Table 4.

2.1.4. Zinc

As for Zn, it was found that Thlaspi alpestre var. calaminare (nowadays classified as Noccaea caerulescens) contained a minimum of 1% Zn in the dry leaf biomass [89]. Rascio [90] observed another Zn hyperaccumulator, Thlaspi rotundifolium ssp. Cepaeifolium, while Reeves and Brooks [91] proved that the Thlaspi genus consists of many Zn hyperaccumulators, with accumulated concentrations surpassing 1000 mg kg−1 after experiments in mine tailings in Northern Italy. Baker et al. [92] showed that T. caerulescens, when grown in nutrient-rich soils, could accumulate high concentrations of Zn, Cd, Co, Mn, and Ni. There were even instances where Zn accumulation reached the threshold of 10,000 mg kg−1 [17] or surpassed it, which was initially considered as the established limit to characterize a plant as a Zn hyperaccumulator. Furthermore, Brown et al. [93] added 10,000 μΜ Zn and 200 μΜ Cd in a nutrient solution and caused an accumulation of 33,600 mg Zn kg−1 in T. caerulescens. Also A. halleri, after being tested in a hydroponics experiment treated with 1 μΜ and 1000 μΜ Zn, managed to accumulate Zn in the shoots 300 mg kg−1 and 32,000 mg kg−1, respectively [94]. Also, in a hydroponics experiment conducted in China using Potentilla griffithii, it was shown that the plant was able to accumulate up to 11,400 mg Zn kg−1 leaf dry weight after being treated with 160 mg Zn L−1 [95]. However, the threshold of 10,000 mg kg−1 was later revised to 3000 mg kg−1 by Van der Ent et al. [18]. Also, Grimm [96] reported an accumulation of up to 11,700 mg kg−1 in stems of Brassica juncea, when the soil concentration of Zn was only 330 mg kg−1. The list of all Zn hyperaccumulator plant species is shown in Table 5.

2.1.5. Manganese

Manganese is a major trace nutrient with an expected concentration in plants grown in normal soil of ca. 40–50 mg kg−1 and usual soil pseudo-total concentrations ca. 500–600 mg kg−1. However, in Mn-rich soils, Mn hyperaccumulator plants may accumulate 2 or even 3 orders of magnitude higher Mn concentrations. Research conducted by Jaffré [36,106] reported that 98 out of 450 of the plant species growing on soils developed on ultramafic rocks in New Caledonia with background Mn concentrations of 1000–5000 mg kg−1 had average Mn concentrations above 1000 mg kg−1, while six species among them surpassed even 10,000 mg kg−1. Based on these findings, Baker and Brooks [17] decided to establish a threshold of 10,000 mg kg−1 to define Mn hyperaccumulation. To date, there are 22 species that are known as Mn hyperaccumulators and these are listed in Table 6. In fact, due to exceptionally high levels of Mn, in some cases reaching a concentration of 2–5% of plant DW, the ash of these plants can contain as much as 10–25% Mn.

2.1.6. Selenium

Selenium (Se) levels in soil are often lower than 2 mg kg−1, but they can increase to a few hundreds of mg kg−1. The selenium concentration in plants dry biomass is typically lower than 0.1 mg kg−1 and may even be as low as 0.01 mg kg−1. Nonetheless, it was discovered that legumes belonging to the Astragalus genus were able to accumulate Se to elevated concentrations, locally exceeding 1000 mg kg−1 in the USA in Se-rich soils [119]. Table 7 shows the 20 hyperaccumulators for Se. Due to the typically low levels of Se in plant tissues, Reeves [20] argued that a limit of 100 mg kg−1 should be considered for identifying Se hyperaccumulators. To date, utilizing Se hyperaccumulators for the economic extraction of Se has not been suggested. However, there are potential uses of these plant species in phytoremediation for cleaning up contaminated soils in crops harvested for stock feed and Se biofortification [120].

2.1.7. Thallium

Thallium is a precious metal high in demand but comparatively rare in nature. Although it is very toxic and difficult to find in the environment, natural, along with anthropogenic, origins can introduce significant amounts of Tl in soil. Several plant species are capable of extracting Tl from soils, making them useful for reclaiming it. Specifically, according to Leblanc et al. [130], there was an unexpected accumulation of Tl in Iberis intermedia and Biscutella laevigata in Pb/Zn mines in France. Moreover, for I. intermedia, an uptake of 4000 mg of Tl kg−1 was reported and over 14,000 mg kg−1 DW for B. laevigata, where the background concentration of Tl was 10 mg kg−1 in mine tailings. The presence of Tl in soil at levels that make Tl PM feasible is a possibility. However, there are only a few sites where this could be realistically implemented due to the limited availability of suitable soil conditions.

2.1.8. Noble Metals

Recently, NMs are being examined for their PM potential. This category constitutes metals such as gold (Au) and silver (Ag), as well as a sub-group of platinum metals including iridium (Ir), osmium (Os), palladium (Pd), platinum (Pt), rhodium (Rh), and ruthenium (Ru) [15]. These metals find extensive use across different industries and are often referred to as precious metals, obtaining significant attention due to their economic value. Interestingly, while the demand for NMs increases, their primary ore sources are gradually depleted. Hence, alternative methods of extraction, like PM, are being explored. Recent research and development on the extraction and accumulation of these metals in plants has shed light on the PM pathway for recovering these metals, although real-time implementation remains limited [131]. It should be noted that NMs can not be absorbed by plants from the metallic state as plants are capable of absorbing only the soluble species of any inorganic element found in the soil solution.

Gold

Across the world, considerable quantities of Au deposits can be found in natural enriched soils and mine tailings. Mohan [132] mentioned that PM of Au is a cost-effective way to extract Au from mine tailings and low-grade ores. Conventional mining removes the most Au from ores, leaving behind a significant portion of Au that cannot be extracted due to economic disadvantages. Since 1988, the hyperaccumulation limit of Au has been set to be higher than 1 mg kg−1 DW, based on the normal concentrations of Au in plants, which are usually around 0.01 mg kg−1. However, Au must be in the soluble form before absorption can take place. This can be achieved by adding several types of lixiviants (e.g., NaCN, SCN, S2O3−2) depending on the properties of the substrate. In low-pH sulphate containing substrates, thiocyanate can solubilize Au.
It is reported that conifers growing in areas with Au ores in Canada could accumulate up to 0.02 mg of Au kg−1, while the background levels in plants are 100 times lower [133]. Also, in an experiment where Brassica juncea was planted in pots containing 5 mg Au kg−1 in artificial substrate, Anderson et al. [134] concluded that the application of 160 mg of ammonium thyocanate kg−1 of dry substrate weight achieved an accumulation of 57 mg Au kg−1. Wilson et al. [135] also referred to research where B. juncea was able to absorb 63 mg Au kg−1 DW after soil was treated with sodium cyanide. The plants were grown on an oxidized ore pile containing 0.6 mg Au kg−1. They concluded that an economically viable option for Au extraction that could yield 1 kg of Au ha−1 would require a total harvested biomass of 10,000 kg ha−1 containing 100 mg kg−1. Furthermore, Msuya et al. [136] mentioned that carrots (Daucus carota) accumulated 0.779 mg ha−1 of Au when ammonium thiocyanate was added and 1.45 mg ha−1 when thiosulphate was added, while the concentration of Au in the growing substance was 3.8 mg kg−1. Lamb et al. [137] performed a study where they used cyanide and thiocyanate to test the accumulation of Au in Berkheya coddii, Brassica juncea, and Cichorium intybus. They found that the leaves of B. juncea and C. intybus in soils spiked with cyanide had the highest Au accumulation at a rate of 326 mg kg−1 (B. juncea) and 164 mg kg−1 (C. intybus). Another study conducted by Vural and Safari [138] evaluated the potential of Helichrysum arenarium for Au accumulation; they reported TF and BAF values of 2.04 and 0.59, respectively. They concluded that enhancing the availability of Au in soil by adding chelators like NaCN, SCN, and thiosulfates can lead to the usage of Helichrysum arenarium in the PM process.

Silver

To date, there are no reports about the hyperaccumulation of Ag by plants. Even though the Ag concentration in plants can be up to 1 mg kg−1, some plants such as Lupinus angustifolius can reach up to 126 mg kg−1 by induced accumulation [5]. Harris and Bali [139] were the first to propose the uptake of Ag nanoparticles by plants. They conducted research in which they observed an accumulation of significant amounts of Ag in Brassica juncea and Medicago sativa plants. Specifically, B. juncea absorbed up to 12.4% of its weight in Ag when exposed to 1000 mg L−1 of AgNO3 for 72 h, while M. sativa accumulated up to 13.6% of its weight in Ag when exposed to a solution containing 10,000 mg L−1 AgNO3 for 24 h. Silver nanoparticles that were stored inside the plant’s tissues had a size of 50 nm. In addition, the application of EDTA was not helpful in enhancing the absorption of Ag. EDTA had a rather negative effect as it caused toxicity to plants and this decreased the amount of Ag that would otherwise be capable for extraction. They concluded that plants could be used to produce large quantities of metallic nanoparticles.

Palladium

Out of all the platinum group metals, Pd has gathered significant attention due to its high abundance compared to others. The first report for Pd accumulation in plants was made by Fuchs and Rose [140], who measured a concentration of 285 μg kg−1 in ash from Pinus flexilis. In addition, Kothny [141] reported an accumulation of 400 μg kg−1 of Pd in the ash of Quercus chrysolepsis, while the background level was 140 μg kg−1. The hyperaccumulator Berkeya coddii was investigated for its ability to absorb Pd by Nemutandani et al. [142] from a contaminated site with 70 μg kg−1 Pd. It was found that B. coddii was able to extract 180 μg kg−1 in roots and 710 μg kg−1 in leaves. The plant demonstrated a notable ability to accumulate and translocate Pd, with BAF and TF values reaching 10.1 and 3.9, respectively.
There have been no officially defined Pd hyperaccumulators; it is anticipated that the threshold for all NMs would be approximately 1 mg kg−1, assuming their normal low concentration in plants. On the other hand, Pd, like Au, has limited solubility in soil. Hence, soil amendments are often used in order to increase Pd solubility and improve the accumulation in plants. For instance, KCN has been utilized to increase the absorption of Pd by plants. Walton [143] used KCN at a rate of 10 g L−1 to enhance Pd accumulation in B. coddii when cultivated in mine tailings containing 315 μg kg−1. They reported that plants had accumulated Pd to levels as high as 7677 μg Pd kg−1. In addition, Aquan [144] conducted an experiment utilizing gossan rock as a substrate, containing 205.5 μg Pd kg−1. They assessed the Pd accumulation in Cannabis sativa and found that the average Pd concentration was 30,336 μg kg−1 in the aboveground part of the plant, reaching as high as 62,420 μg kg−1 when KCN was applied. Another study conducted by Harumain et al. [145] managed to trigger the hyperaccumulation of Pd in Salix purpurea and Miscanthus through the use of KCN. They employed a synthetic ore with a Pd concentration of 50–100 mg kg−1 as growth medium. As a result, the leaves of S. purpurea and Miscanthus were found to contain 820 and 505 mg kg−1, respectively. Overall, the idea of phytomining Pd is relatively new. The effectiveness of using plants to extract this valuable element has only been minimally demonstrated so far.

Platinum

While the potential of plants to accumulate Pt in its tissues was mentioned by Fuchs and Rose [140], it is noteworthy that numerous studies on Pt PM have emerged recently. Nemutandani et al. [142] revealed that B. coddii grown in Pt-polluted areas with concentrations of 0.04 mg kg−1 soil achieved an uptake of 0.22 mg kg−1 in leaves and 0.18 mg kg−1 in roots. The plant exhibited BAF (5.5) and TF (1.2) values greater than 1.0, highlighting its ability to accumulate and translocate Pt. A study conducted by Kovacs et al. [146] reported the identification of plants from contaminated brownfield land with a Pt concentration of 3.06 mg kg−1. Also, Diehl and Gagnon [147] found concentrations of 14.6 mg kg−1 in Daucus carota collected from an area near a heavy-traffic motorway. White mustard, Sinapis alba, was also reported to show great ability in accumulating Pt by Kińska and Kowalska [148]. They found an impressive accumulation of 5973 mg kg−1 dry root weight, while the concentration of Pt in the nutrient solution was 1.0 mg L−1. Despite numerous reports on Pt in plants, it is worth mentioning that experiments aimed at Pt PM have not been conducted.

Rhodium

Generally, studies regarding the occurrence of Rh in plant tissues are very limited due to its scarce presence. A study conducted by Diehl and Gagnon [147] reported a Rh concentration of 0.7 mg kg−1 in Daucus carota that was collected from four different areas near heavy-traffic roads. Bonanno [149] found that the Rh concentration in Phragmites australis, collected from an extremely urbanized riverside area, ranged from 1.11 to 1.13 mg kg−1. Generally, it is worth noting that Rh has the highest solubility among the platinum group elements [15]. This suggests that Rh has great potential use in PM applications.

2.2. Factors Influencing Metal Availability and Behavior in Soil

2.2.1. pH

Soil activity plays a crucial role in how easily certain nutrients can dissolve in the soil solution and be taken up by the plants. As a result, soil pH affects the ability of plants to absorb essential nutrients and other elements from soil. For instance, Punjari [150] reported that when the pH of soil drops below 6.5 and 5.3, the amount of soluble Cd and Zn increases. Likewise, the solubility of Cu and Pb increases at pH levels below 4.5 and 3.5, respectively. On the other hand, at this low pH level, the mobility of As ions decreases. This is related to the fact that As is present in the soil solution as oxyanion, either in its trivalent (arsenite; AsIIIO33−) or its pentavalent species (arsenate; AsVO43−). The same is the case with any other contaminant anions (such as hexavalent chromium; CrVIO42−)—their availability is proportional to soil pH. The significance of pH in determining metal solubility and availability can be indicated by the fact that the pH in the root zone of certain plants is often up to 2 units lower than the surrounding soil pH due to the release of organic acids [151]. The amount of metal extraction is directly related to the concentration of the metal in soil and pH of the extracting agent.
There are various ways to reduce soil pH. This can be achieved by using different substances, including fertilizers containing ammonia, acids, and zero-valent sulfur, as amendments. However, there are limitations as to how much the pH can be lowered because most plant species can only thrive within a specific pH range. Usually, the lowest pH that many plants can tolerate is c. 4.5 [5]. Hence, while it is possible to adjust the pH utilizing soil amendments, it is important to carefully consider the range that allows plants to grow without adverse effects. On the other hand, acidic pH values can be corrected with the application of “liming” materials, which are mainly phases with abundant inorganic carbonate phases, i.e., CO32− and HCO3, such as marble dust (either calcitic or dolomitic), as well as their derivative, caustic calcium oxide. Robinson et al. [45] studied the effects of different substances, namely MgCO3, CaCO3 (both alkalinity-bearing materials), and sulfur (an acidifying material), on the uptake of Ni and Co by Berkeya coddii. When MgCO3 was added to the soil, it increased the soil pH from 6.9 to 8.7 which led to reduced uptake of Ni and Co by B. coddii. On the other hand, the application of sulfur resulted in a decrease in soil pH from 6.9 to 5.5, which increased the plant’s uptake of both metals. In addition, the application of Ca salts made plants tolerant to Ni [8]. The authors concluded that the addition of acidifying materials, like sulfur, can help extract more metals from soil by making them more available to plants. The shoot Ni concentration increased in Alyssum sp. at pH levels ranging from 5 to 7, with maximum concentrations being obtained at a pH of 6.5 [152]. The same relationship between cationic metal species and soil pH (which is inversely proportional—the former increases when the latter decreases and vice versa) has repeatedly been confirmed in experimental studies [36,110]. This was also found by Zhong et al. [153] in soil samples collected from heavy metal-contaminated soils in China; they reported a significant availability in soil for Ni, Zn, Cu, Cd, and Cr at pH levels lower than 4.5. They also concluded that soil pH is the most important factor concerning metal availability in soil. Specifically, pH was reported to affect all physicochemical and biological processes in soil, which in turn affect the behavior and the way metals interact with other substances. Since acidic pH tends to decrease metal sorption in soil, metal availability increases due to increased metal concentration in soil solution.

2.2.2. Fertilizers

In order to successfully implement the technique of PM, the availability of metals in soil needs to be ensured. Besides optimizing soil pH levels to guarantee metal absorption, adding fertilizers to achieve high-yielding phytoextraction crops is another factor that needs to be mentioned. Fertilizers containing NH4-N can reduce soil pH due to the fact that NH4+ in aerated soils is readily oxidized within weeks to NO3, a process yielding 2 mols of H+ per mol of nitrified N. This enhances further the availability of metals in soil, as well as providing optimal conditions for the growth of these certain plant species [154]. In fact, hyperaccumulators grown in soils derived from ultramafic rocks exhibit a noticeable positive response to the application of fertilizers. For instance, in Vertisols in Albania, there was a 10-fold increase in the yield of A. murale after applying inorganic fertilizers [23].
Sheoran et al. [5] highlighted that the application of fertilizers significantly increased the maximum annual growth of Alyssum bertolonii compared to the non-fertilized treatment, which resulted in higher biomass yields. Specifically, the application of NPK fertilizers achieved an overall increase of 308%. Thus, fertilization can greatly affect the phytoextraction process, to remove metals from soil, making it more effective. The same work reported that Ni levels were increased in the leaves of Berkeya coddii from 2500 mg kg−1 to 4200 mg kg−1. Similar results were reported by Bennett et al. [155], where the application of NPK fertilizers doubled the annual biomass production of Alyssum bertolonii, Streptanthus polygalonoides, and Thlaspi caerulescens, without reducing the concentration of Ni in their shoots. In another study, the application of 100 kg P ha−1 in serpentine soils led to an improved yield in Alyssum murale plants [7]. Overall, high N fertilization promotes higher accumulation of Ni, Cd, and Zn. Sheoran et al. [156] used Salix on contaminated soil containing low levels of Cd (1.9–2.4 mg kg−1) and showed that the application of (NH4)2SO4 at 100 mg N kg−1 significantly increased the Cd and Zn plant concentrations to 4.5 for Cd and 400 mg kg−1 for Zn. Similar results were reported by Schmidt [157], who noticed a significant increase in Cd in Lolium perenne after adding (NH4)2SO4 as fertilizer. In that same work, in soils with low concentrations of Zn and Ni, the application of this fertilizer also significantly increased the crop metal concentrations without any loss of total yield. The same was agreed by Babau et al. [158], who tested Robinia pseudoacacia for its ability to absorb Pb, Cd, and Cu with and without the presence of NPK fertilizer. We conclude based on the above-reviewed literature that the addition of nutrients to soil (mostly N) increases the ability of plants to absorb metals and thus it is likely to increase the phytomining effectiveness of hyperaccumulators; hence, fertilizers are considered a critical factor for the success of commercial PM processes.

2.2.3. Chelates

In hyperaccumulators, the number of metals these plants can remove from soil depends on biomass production and the concentration of the metals within their tissues; thus, often their extraction ability is limited [159]. In order to increase the extraction of metals, chelates can be a possible option for higher growth rates. Chelates are capable of extracting various metals from soil. They have the ability to bind metals and form water-soluble organometallic complexes. Once they are formed, they can introduce metals into the soil solution by removing them from the surfaces they were attached to [160]. Some of the most used chelating agents are EDTA (ethylenediaminetetraacetic acid), CA (citric acid), DTPA (diethylenetriaminepentaacetic acid), NTA (nitrolenetriaceticacid), EDDS (ethylenediaminedisuccinicacid), and thiosolutions like thiocyanates. The former two agents are the most widely utilized [10,161]. Specifically, Robinson et al. [162] used EDTA and CA on B. coddii in order to phytomine Ni. Interestingly, they found that even though soluble Ni increased in the rhizosphere, the Ni concentration in the shoots decreased. They concluded that competition between EDTA, CA, and the plant’s own excretes resulted in trapping Ni in the root system of the plant. However, Meers et al. [163] reported an increase in Ni uptake in sunflowers which was 1.8 to 2.8 times greater by adding 1.6 mmol kg−1 EDTA and EDDS. Another study conducted by Blaylock et al. [164] where EDTA was applied in heavy metal-contaminated soils showed an enhanced accumulation of Pb, Cd, Cu, and Zn in Brassica juncea shoots. Nevertheless, they concluded that chelators like EDTA may pose an environmental risk due to the fact that they can not be easily degradable by soil microbes, resulting in groundwater pollution. Additionally, Wang et al. [165] reported results from a field experiment where soil-applied EDTA and EDTA−ethyl lactate were found to enhance Cd extraction by 20 and 29%, respectively, in Salix. Gold has also been found to accumulate in plants like Brassica juncea, Zea mays, Daucus carota, and Cichorium spp. by adding chelate agents like thiosolutions.

3. Metal Accumulation in Plants and Their Subsequent Recovery

3.1. Mechanisms of Metal Accumulation in Plants

The process of metal hyperaccumulation in plants, and thus phytomining, consists of several different stages that plants go through. Making the soil in the immediate vicinity of the plant’s roots more acidic and releasing organic compounds is one of the plant’s strategies to absorb metals from soil. The acidification of soil can occur due to the secretion of hydrogen ions (H+), which can help in dissolving metals and making them available for absorption. This process is carried out by essential enzymes like the H+ pumps, such as the plasma-membrane H+ ATPases [164]. Additionally, roots excrete organic acids, such as malonic and oxalic, as well as enzymes and metal-chelating molecules, like phytosiderophores. These substances have the ability to enhance the absorption of metals from soil by binding them and thus assist their phytomining [166]. The movement of metals inside the plant cells can have two potential pathways, which are the apoplastic flow and the symplast transport. Metals can enter the root apoplast, which is the space between cells in the root. Even though solutes are capable of moving upwards through the plant using this pathway, the more effective way is through the plant’s vascular system, the xylem [5]. Moreover, to enter the xylem, where long-distance transport occurs, metals have to go through the Casparian strip, which is a waxy layer of cells that surrounds the xylem and is impermeable to solutes. Thus, metals must cross membranes aided by pumps or channels in the endodermal cells. After crossing the Casparian strip or membranes, the movement of metals within the xylem is called “symplast transport”. Symplast transport is more regulated than apoplastic flow because the movement of metals is controlled by the properties of the cell membranes and the presence of specialized transport proteins or channels that facilitate their movement [164]. Generally, plants have the ability to manage heavy metals, avoid toxicity, and store them safely [167]. Different oxidation levels of heavy metals have varying characteristics during absorption, conveyance, and detoxing within the plant. After being translocated to shoot cells, metals can be stored in specific cellular locations such as in the apoplast, epidermis cells, the mesophyll, and cell walls, where they will not disrupt any cellular processes. This is achieved by sequestering metals away from cellular processes that could otherwise be harmed. The trapping of the metals often takes place in plant vacuoles [168].

3.2. Metal Recovery

To recover the metals accumulated in plants, various metal extraction technologies are employed. However, before extraction can be undertaken, the plant biomass needs to be reduced to a manageable quantity so that any plant-borne metals may be concentrated in the remaining material. This is crucial for reduction in the transportation costs, as well as the size of machinery used for metal extraction. Biomass reduction can occur in various ways, including composting, compaction, and thermal conversion.
According to Hu and Gholizadeh [169], the thermal conversion methods of combustion and ashing are considered to be the most practical for concentrating metals. In fact, these processes involve high temperatures, ranging from 350 to 1000 °C, that lead to fast reaction times and effective biomass volume reduction. During combustion, the polluted biomass turns into gas and energy is released as heat. Metals cannot easily escape as gases during the process, resulting in the formation of ash containing high concentrations of valuable metals. On the other hand, unwanted emissions of C and N oxides are issues that need to be considered. However, the overall approach is considered to be a neutral C-releasing activity in terms of greenhouse gas emissions due to the fact that C can be captured and be stored via oxy-fuel combustion [15].
Ashing is the same process as combustion but takes place at lower temperatures (300–550 °C) and is commonly used in laboratory environments for analytical purposes. For instance, Zhang et al. [170] found that the best conditions of biomass combustion for Ni recovery were ashing at 550 °C for a duration of 3 h. By doing this, they were able to obtain high quality ash reaching 20% of the weight of Ni. However, according to Dastyar et al. [171], in order to recover good quality ash, it is important to understand the behavior of metal dynamics during thermal conversion. Specifically, metal behavior is influenced by various factors, including the type of plant biomass, the equipment that is being used, and combustion conditions like temperature, pressure, and duration. For example, Hazotte et al. [172] compared metal concentrations in three Ni hyperaccumulators (Odontarrhena chalcidica, Leptoplax emarginata, and Berkheya coddii), along with biomass volume reduction in two different temperatures (550 and 900 °C) and standard pressure. They reported that the Ni concentration at 900 °C was higher compared to that at 550 °C in all three hyperaccumulator plant tissues. In addition, they reported that at 550 °C, over 95% of the organic material was broken down and 99% of C was eliminated from all plant species. They concluded that combustion is an effective method for the degradation of plant biomass, while achieving high concentrations of metals with no notable loss up to 900 °C. In another study conducted by Kovacs et al. [147], Au could not be detected in the woody biomass. Nevertheless, after the combustion of plant material, Au could be detected in the residual solid, where it exhibited a considerable concentration of 4.10 mg kg−1. Nevertheless, despite findings regarding the importance of biomass combustion and ashing, there is rather limited information in the literature regarding enrichment in the solid residues with valuable phytomined metals.
After biomass combustion or ashing are finished, a metal-rich ash is obtained. However, it may still contain impurities that need to be further processed in order to isolate the target metal. This can be achieved by employing pyrometallurgical, hydrometallurgical, and bio-metallurgical methods. Pyrometallurgy refers mainly to the traditional process of smelting, but it can contain other methods too, such as incineration, chlorination, and volatilization [15]. Pyrometallurgy involves the use of high temperatures to separate the metals from other materials. However, the environmental effects of this method, like the release toxic fumes, and the expensive equipment, are some limitations of this approach. Hydrometallurgy involves the use of leaching agents in order to extract and separate the metals. Metals within the solution are separated and isolated by utilizing various techniques, including ion capture, liquid separation, ion exchange, and electrochemical reduction. Compared to pyrometallurgy, hydrometallurgy is easier to use, offers great metal selection, is efficient in recovering metals, and less expensive in terms of equipment. Interestingly, most of the studies examine the use of both methods for isolating metals. For instance, Lamb et al. [137] managed to isolate Au from woody plant material after ashing the biomass and using 2 M HCl as a leaching agent. The solution was separated by adding methyl isobutyl ketone, followed by the addition of sodium borohydrate. This resulted in the formation of a black deposit which was heated at 800 °C, which led to the formation of pure Au.
During the last decade, a new approach called bio-metallurgy has become popular for recovering metals. It consists of two main parts: “bioleaching” and “biosorprtion”. Bioleaching refers to the use of microorganisms, like bacteria, fungi, and algae, to interact with metals through oxidation-reduction reactions. Bioleaching has been considered to be an eco-friendly and economically viable option. However, it is not efficient enough for metal recovery and it takes a lot of time to be completed, which leads to restricted use and means it is only used for laboratory-scale operations [173]. On the other hand, biosorprtion utilizes microbes, fungi, yeasts, algae, and waste products, which are used to collect metals. These organisms act like “magnets” by entrapping metals through physical and chemical interactions between metals and the surface of these microbes. Generally, biosorption has emerged as a cost-effective and environmentally friendly technique compared to traditional methods, especially for aquatic media [174]. Processes are shown in Figure 1.

4. Financial Considerations of Phytomining

The economic success of PM relies on several factors, such as (i) the metal content in soil and plants, (ii) the amount of biomass production, and (iii) the amount of retrieved (or phytomined) metal that can be recovered during the combustion of the produced biomass which can then be captured and sold. However, the most imperative factor for PM is the revenue generation, which depends on the market demands and the world price of the metal being extracted through plants [10,174]. Metal prices can vary from USD 0.75 kg−1 for Fe ores to USD 61,375 kg−1 for Au [175]. As a result, metals with high prices offer better economic prospects. Precious metals can be very costly and their price keeps rising consistently (Figure 2); thus, there must be a sufficient amount of these metals in the harvested biomass in order for the PM process to be profitable. The economic proposed concentration of precious metals is shown in Table 8.
Manganese is an example of an element where even though its price is low, USD 6.5 t−1, there are certain plants that can accumulate large quantities of it. Macadamia plants have been reported to accumulate high concentrations of Mn in their tissues; therefore, they can be a feasible and profitable option. However, metal prices often undergo cycles of fluctuations—PM may be viable, if technically possible, even while current market prices happen to be low. There is always the capability of combusting the biomass which contains the extracted metal to recover the economic value; metals can be extracted from plants and the remaining material can be burned to generate energy. After combustion, the ash could be stored until the world price of the extracted metals improves [10,174]. Moreover, companies do not only make profit just from the PM process but also from harnessing the biomass for energy production. Additionally, they may be utilizing the resulting ash as a source of C and potash. Furthermore, they may be deriving advantages from the sale of CO2 credits [57].
As mentioned earlier, during the first field experiment for extracting Ni, Streptanthus polygaloides yielded a total dry biomass of 10,000 kg ha−1 containing 10,000 mg kg−1 Ni. It is reported that a net profit of USD 513 ha−1 could be achieved with a price of USD 7.65 kg−1 at that time. In addition, a net return of USD 131 ha−1 could be achieved due to energy yield during the combustion of the biomass, leading to a gross return of USD 765 ha−1. Considering Ni current prices, of approximately USD 20 kg−1, such yields and Ni uptake could achieve a gross profit of USD 1341 ha−1. Anderson et al. [177] utilized the Tl hyperaccumulator I. intermedia and extracted 8 kg ha−1 of Tl in a total of 10,000 kg ha−1 dry biomass, which was worth a total of USD 2400. This means that a crop with 10,000 kg ha−1 of biomass would need to accumulate at least 170 mg kg−1 Tl, which is achievable by I. intermedia. On the other hand, as previously stated, B. laevigata produces lower biomass (4000 kg ha−1), which means that an average of 425 mg Tl kg−1 is required—it was found that 39% of the plants surpassed this threshold in that work. Selling biomass for power generation could also add a profitable amount to the farmer’s income.
In addition, Nkrumah et al. [178] suggested the use of both an intensive and extensive Ni production system. Moreover, both approaches were found to be profitable. Extensive systems are those where the only costs are herbicides, fertilizers, and any management practices, while an intensive production system requires additional labor and equipment costs. They even made a hypothetical scenario where they calculated the economic potential of Ni PM, comparing the two approaches. Specifically, they (a) considered an average market price of USD 18 over a 5-year period from 2010 to 2015, (b) estimated the amount of Ni that plants can produce in the two systems, which was leveled at 200 kg ha−1 for the intensive system and 110 kg ha−1 for the extensive, (c) calculated the cost of the production in 2016, which was approximately USD 1074 ha−1 year−1 and USD 600 ha−1 year−1 for the intensive and the extensive systems, respectively, and (d) included in that about 20% of the value of Ni that would cover expenses and costs associated with metal recovery from plants. Assuming these factors, they calculated the gross values of an annual PM crop, which was found to be USD 3600 ha−1 year−1 for the intensive system and USD 1980 ha−1 year−1 for the extensive system. After accounting the costs that were mentioned before, they measured a net value of USD 1806 ha−1 year−1 for the intensive system and a net value of USD 980 ha−1 year−1 for the extensive system. In addition, other sources for enhancing profit could be harnessing energy from combustion or selling carbon credits for reducing gas emissions.
Lee et al. [176] mentioned the remarkable metal-accumulating abilities of Haumaniastrum robertii and Aeolanthus biformifolius. Specifically, H. robertii is capable of accumulating up to 1% of its dry biomass as Co, while A. biformifolius can accumulate up to 1% of its dry biomass as Cu. As a result, commercial use of H. robertii can lead to a harvest of 200 kg Co per 20 t of dry biomass per year. Assuming 100% Co recovery and a current Co market price per ton of USD 33,618, this could potentially result in USD 6723 of income ha−1 per year. Similarly, A. biformifolius has the potential to yield 200 kg of Cu from 20 t of dry biomass ha−1 per year. Assuming 100% Cu recovery and a current market price of USD 7963 for Cu, this could lead to a gross income of USD 1592 ha−1 per year.
As the process of PM continues over time, the ore deposits eventually get depleted, making it less economically viable to obtain them. Nevertheless, compared to the conventional mining methods, the time when it will no longer be cost effective to mine these metals through PM is much higher. This is because PM is a cost–benefit method. Moreover, plants used in the PM process can have a multifaceted application until the soil where they are planted is ready for food production once again. Ongoing research towards a better understanding of the mechanisms involved in the PM process, along with other improvements, is expected to extend the economic viability of mined metals through PM.

5. Conclusions and Need for Future Research

While PM holds significant potential, it faces certain limitations, such as the selection of specific plants for certain metals, along with any challenges presented by the formation of the metal compounds in soil. However, PM leverages the natural ability of the plants to accumulate large amounts of valuable metals. It can not only create income for growers but it also provides mineral supplementation and biofuel resources. We indicated that generated income is not improbable; e.g., significant earnings are possible if a yield of 10 t ha−1 of the hyperaccumulator is combined with an average Ni concentration of 1000 mg kg−1. Specifically, economic viability depends on factors like metal content in soil, BAF and TF values of the selected plant species, biomass yields, and metal prices. On the other hand, extensive research remains necessary so that the capabilities and limitations of the PM process may be fully understood. While the majority of the studies have been conducted at the laboratory scale and in a controlled environment for short periods, the true potential of PM awaits investigation at the commercial scale under field conditions for longer periods of time. In addition, interdisciplinary research efforts that combine expertise in different fields of science, such as soil science, plant biology, genetics, and soil chemistry, as well as knowledge of farming practices, are essential for unlocking PM’s full potential. Moreover, it is essential to acknowledge that most of the conducted experiments have taken place in developed countries, leaving a knowledge gap for developing nations. We conclude that PM, although a relatively new and unexplored concept, may provide economic and environmental benefits to soil end-users and managers who must cultivate on metal-contaminated soils as PM may turn yield shortages (of specific commercial crops) into benefits if high-yield hyperaccumulators are cultivated for industrial valorization of their high-metal-content biomass. The future of PM should be approached with a balanced perspective, realizing that it may represent a shift in our resource extraction. Phytomining could be the transformative concept whose time has come, provided that we continue to invest in comprehensive research, innovation, and sustainable practices.

Author Contributions

Conceptualization V.A.; writing—original draft preparation, C.K.; review and editing, V.A. and G.T.; visualization, C.K. and G.T.; supervision, V.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Aziz, I.; Mujeeb, A. Halophytes for phytoremediation of hazardous metal (loid) s: A terse review on metal tolerance, bio-indication and hyperaccumulation. J. Hazard. Mat. 2022, 424, 127309. [Google Scholar] [CrossRef] [PubMed]
  2. Reeves, R.D.; van Der Ent, A.; Echevarria, G.; Isnard, S.; Baker, A.J. Global distribution and ecology of hyperaccumulator plants. In Agromining: Farming for Metals, 2nd ed.; van der Ent, A., Baker, A.J., Echevarria, G., Simonnot, M.O., Eds.; Springer: Cham, Switzerland, 2021; pp. 133–154. [Google Scholar] [CrossRef]
  3. Soleymanifar, S.; Ehsanpour, A.A.; Ghasemi, R.; Schat, H. The effect of L-histidine on nickel translocation and the activities of antioxidant enzymes in hyperaccumulator (Odontarrhena inflata) and non-accumulator (Aurinia saxatilis) plants. Plant Soil 2023, 491, 1–18. [Google Scholar] [CrossRef]
  4. Siyar, R.; Doulati Ardejani, F.; Norouzi, P.; Maghsoudy, S.; Yavarzadeh, M.; Taherdangkoo, R.; Butscher, C. Phytoremediation potential of native hyperaccumulator plants growing on heavy metal-contaminated soil of Khatunabad copper smelter and refinery, Iran. Water 2022, 14, 3597. [Google Scholar] [CrossRef]
  5. Sheoran, V.; Sheoran, A.S.; Poonia, P. Phytomining: A review. Miner. Eng. 2009, 22, 1007–1019. [Google Scholar] [CrossRef]
  6. Akinbile, B.J.; Makhubela, B.C.; Ambushe, A.A. Phytomining of valuable metals: Status and prospective-a review. Int. J. Environ. Anal. Chem. 2023, 103, 39133933. [Google Scholar] [CrossRef]
  7. Li, Y.M.; Chaney, R.L.; Brewer, E.P.; Angle, J.S.; Nelkin, J. Phytoextraction of nickel and cobalt by hyperaccumulator Alyssum species grown on nickel-contaminated soils. Environ. Sci. Technol. 2003, 37, 1463–1468. [Google Scholar] [CrossRef]
  8. Chaney, R.L.; Chen, K.Y.; Li, Y.M.; Angle, J.S.; Baker, A.J. Effects of calcium on nickel tolerance and accumulation in Alyssum species and cabbage grown in nutrient solution. Plant Soil 2008, 311, 131–140. [Google Scholar] [CrossRef]
  9. Rai, V.; Liu, D.; Xia, D.; Jayaraman, Y.; Gabriel, J.C.P. Electrochemical approaches for the recovery of metals from electronic waste: A critical review. Recycling 2021, 6, 53. [Google Scholar] [CrossRef]
  10. Jaffré, T.; Brooks, R.R.; Lee, J.; Reeves, R.D. Sebertia acuminata: A hyperaccumulator of nickel from New Caledonia. Science 1976, 193, 579–580. [Google Scholar] [CrossRef]
  11. Xie, C.; Xiao, Y.; He, C.; Liu, W.S.; Tang, Y.T.; Wang, S.; Qiu, R.L. Selective recovery of rare earth elements and value-added chemicals from the Dicranopteris linearis bio-ore produced by agromining using green fractionation. J. Hazard. Mater. 2023, 443, 130253. [Google Scholar] [CrossRef]
  12. Robinson, B.; Fernández, J.E.; Madejón, P.; Marañón, T.; Murillo, J.M.; Green, S.; Clothier, B. Phytoextraction: An assessment of biogeochemical and economic viability. Plant Soil. 2003, 249, 117–125. [Google Scholar] [CrossRef]
  13. Subašić, M.; Šamec, D.; Selović, A.; Karalija, E. Phytoremediation of cadmium polluted soils: Current status and approaches for enhancing. Soil Syst. 2022, 6, 3. [Google Scholar] [CrossRef]
  14. Islam, M.D.; Hasan, M.M.; Rahaman, A.; Haque, P.; Islam, M.S.; Rahman, M.M. Translocation and bioaccumulation of trace metals from industrial effluent to locally grown vegetables and assessment of human health risk in Bangladesh. SN Appl. Sci. 2020, 2, 1315. [Google Scholar] [CrossRef]
  15. Dinh, T.; Dobo, Z.; Kovacs, H. Phytomining of noble metals—A review. Chemosphere 2022, 286, 131805. [Google Scholar] [CrossRef] [PubMed]
  16. Nkrumah, P.N.; Chaney, R.L.; Morel, J.L. Agronomy of ‘metal crops’ used in agromining. In Agromining: Farming for Metals; Metal Resources Reviews; van der Ent, A., Baker, A.J., Echevarria, G., Simonnot, M.O., Eds.; Springer: Cham, Switzerland, 2021; pp. 23–46. [Google Scholar] [CrossRef]
  17. Baker, A.J.; Brooks, R. Terrestrial higher plants which hyperaccumulate metallic elements. A review of their distribution, ecology and phytochemistry. Biorecovery 1989, 1, 81–126. [Google Scholar]
  18. Van der Ent, A.; Baker, A.J.M.; Reeves, R.D.; Pollard, A.J.; Schat, H. Hyperaccumulators of metal and metalloid trace elements: Facts and fiction. Plant Soil 2013, 362, 319–334. [Google Scholar] [CrossRef]
  19. Krämer, U. Metal hyperaccumulation in plants. Annu. Rev. Plant Biol. 2010, 61, 517–534. [Google Scholar] [CrossRef] [PubMed]
  20. Reeves, R.D. Hyperaccumulation of trace elements by plants. In Phytoremediat. Of Metal-Contaminated Soils; Morel, J.L., Echevarria, G., Goncharova, N., Eds.; Springer: New York, NY, USA, 2006; pp. 1–25. [Google Scholar]
  21. Mustafa, A.; Zulfiqar, U.; Mumtaz, M.Z.; Radziemska, M.; Haider, F.U.; Holatko, J.; Brtnicky, M. Nickel (Ni) phytotoxicity and detoxification mechanisms: A review. Chemosphere 2023, 328, 138574. [Google Scholar] [CrossRef]
  22. Minguzzi, C.; Vergnano, O. Il contenuto di nichel nelle ceneri di Alyssum Bertolonii. Atti. Soc. Tosc. Sci. Nat. 1948, 55, 49–74. [Google Scholar]
  23. Chaney, R.L.; Angle, J.S.; Baker, A.J.M.; Li, J.M. Method for Phytomining of Nickel, Cobalt, and Other Metal from Soil. U.S. Patent 5,711,784, 31 August 1999. [Google Scholar]
  24. Robinson, B.H.; Chiarucci, A.; Brooks, R.R.; Petit, D.; Kirkman, J.H.; Gregg, P.E.H.; Dominicis, V. The nickel hyperaccumulator plant Alyssum bertolonii as a potential agent for phytoremediation and phytomining of nickel. J. Geochem. Explor. 1997, 59, 75–86. [Google Scholar] [CrossRef]
  25. Reeves, R.D.; Brooks, R.R.; Macfarlane, R.M. Nickel uptake by Californian Streptanthus and Caulanthus with particular reference to the hyperaccumulator S. polygaloides Gray (Brassicaceae). Am. J. Bot. 1981, 68, 708–712. [Google Scholar] [CrossRef]
  26. Mohsin, M.; Nawrot, N.; Wojciechowska, E.; Kuittinen, S.; Szczepanska, K.; Dembska, G.; Pappinen, A. Cadmium Accumulation by Phragmites australis and Iris pseudacorus from Stormwater in Floating Treatment Wetlands Microcosms: Insights into Plant Tolerance and Utility for Phytoremediation. J. Environ. Manag. 2023, 331, 117339. [Google Scholar] [CrossRef] [PubMed]
  27. Cole, M.M. Geobotanical and biogeochemical investigations in the sclerophyllous woodland and shrub associations of the Eastern Goldfield Area of Western Australia, with particular reference to the role of Hybanthus floribundus (Lindl.) F. Muell. As a nickel indicator and accumulator plant. J. Appl. Ecol. 1973, 10, 269–320. [Google Scholar] [CrossRef]
  28. Durand, A.; Goux, X.; Lopez, S.; Leglize, P.; Benizri, E. Soil nickel contamination levels entail changes in the bacterial communities associated to the rhizosphere and endosphere of Odontarrhena chalcidica. Plant Soil 2023, 493, 17–43. [Google Scholar] [CrossRef]
  29. Bani, A.; Echevarria, G.; Sulçe, S.; Morel, J.L. Improving the agronomy of Alyssum murale for extensive phytomining: A five-year field study. Int. J. Phytoremediat. 2015, 17, 117–127. [Google Scholar] [CrossRef]
  30. Galardi, F.; Corrales, I.; Mengoni, A.; Pucci, S.; Barletti, L.; Barzanti, R.; Gonnelli, C. Intra-specific differences in nickel tolerance and accumulation in the Ni-hyperaccumulator Alyssum bertolonii. Environ. Exp. Bot. 2007, 60, 377–384. [Google Scholar] [CrossRef]
  31. Kanellopoulos, C.; Argyraki, A. Soil baseline geochemistry and plant response in areas of complex geology. Application to NW Euboea, Greece. Geochemistry 2013, 73, 519–532. [Google Scholar] [CrossRef]
  32. Adamidis, G.C.; Aloupi, M.; Kazakou, E.; Dimitrakopoulos, P.G. Intra-specific variation in Ni tolerance, accumulation and translocation patterns in the Ni-hyperaccumulator Alyssum lesbiacum. Chemosphere 2014, 95, 496–502. [Google Scholar] [CrossRef]
  33. Keeling, S.M.; Stewart, R.B.; Anderson, C.W.N.; Robinson, B.H. Nickel and cobalt phytoextraction by the hyperaccumulator Berkheya coddii: Implications for polymetallic phytomining and phytoremediation. Int. J. Phytoremediat. 2003, 5, 235–244. [Google Scholar] [CrossRef]
  34. Van der Ent, A.; Baker, A.J.M.; Van Balgooy, M.M.J.; Tjoa, A. Ultramafic nickel laterites in Indonesia (Sulawesi, Halmahera): Mining, nickel hyperaccumulators and opportunities for phytomining. J. Geochem. Explor. 2013, 128, 72–79. [Google Scholar] [CrossRef]
  35. Becquer, T.; Rigault, F.; Jaffré, T. Nickel bioavailability assessed by ion exchange resin in the field. Commun. Soil Sci. Plant Anal. 2002, 33, 439–450. [Google Scholar] [CrossRef]
  36. Jaffré, T.; Pillon, Y.; Thomine, S.; Merlot, S. The metal hyperaccumulators from New Caledonia can broaden our understanding of nickel accumulation in plants. Front. Plant. Sci. 2013, 4, 279. [Google Scholar] [CrossRef]
  37. Li, G.Y.; Hu, N.; Ding, D.X.; Zheng, J.F.; Liu, Y.L.; Wang, Y.D.; Nie, X.Q. Screening of plant species for phytoremediation of uranium, thorium, barium, nickel, strontium and lead contaminated soils from a uranium mill tailings repository in South China. Bull. Environ. Contam. Toxicol. 2011, 86, 646–652. [Google Scholar] [CrossRef]
  38. Berazaín, R.; de la Fuente, V.; Rufo, L.; Rodríguez, N.; Amils, R.; Díez, G.B.; Asensi, A. Nickel localization in tissues of different hyperaccumulator species of Euphorbiaceae from ultramafic areas of Cuba. Plant Soil 2007, 293, 99–106. [Google Scholar] [CrossRef]
  39. Hamdan, A.M.; Bijaksana, S.; Tjoa, A.; Dahrin, D.; Kirana, K.H. Magnetic characterizations of nickel hyperaccumulating plants (Planchonella oxyhedra and Rinorea bengalensis) from Halmahera, Indonesia. Int. J. Phytoremediat. 2019, 21, 364–371. [Google Scholar] [CrossRef] [PubMed]
  40. Davis, M.A.; Boyd, R.S. Dynamics of Ni-based defence and organic defences in the Ni hyperaccumulator, Streptanthus polygaloides (Brassicaceae). New Phytol. 2000, 146, 211–217. [Google Scholar] [CrossRef]
  41. Rue, M.; Paul, A.L.; Echevarria, G.; van Der Ent, A.; Simonnot, M.O.; Morel, J.L. Uptake, translocation and accumulation of nickel and cobalt in Berkheya coddii, a ‘metal crop’from South Africa. Metallomics 2020, 12, 1278–1289. [Google Scholar] [CrossRef]
  42. Parks, S. Cobalt hyperaccumulation in Rinorea cf. bengalensis (Violaceae) from Sabah: Accumulation potential and tissue and cellular-level distribution of cobalt. Plant Soil 2020, 455, 289–303. [Google Scholar] [CrossRef]
  43. Jaffré, T. Plant mineral nutrition on ultramafic rocks of New Caledonia. Bot. Lett. 2023, 170, 375–411. [Google Scholar] [CrossRef]
  44. Malik, M.; Chaney, R.L.; Brewer, E.P.; Li, Y.M.; Angle, J.S. Phytoextraction of soil cobalt using hyperaccumulator plants. Int. J. Phytoremediat. 2000, 2, 319–329. [Google Scholar] [CrossRef]
  45. Robinson, B.H.; Brooks, R.R.; Clothier, B.E. Soil amendments affecting nickel and cobalt uptake by Berkheya coddii: Potential use for phytomining and phytoremediation. Ann. Bot. 1999, 84, 689–694. [Google Scholar] [CrossRef]
  46. Faucon, M.P.; Shutcha, M.N.; Meerts, P. Revisiting copper and cobalt concentrations in supposed hyperaccumulators from SC Africa: Influence of washing and metal concentrations in soil. Plant Soil 2007, 301, 29–36. [Google Scholar] [CrossRef]
  47. Morrison, R.S.; Brooks, R.R.; Reeves, R.D.; Malaisse, F. Copper and cobalt uptake by metallophytes from Zaïre. Plant Soil 1979, 53, 535–539. [Google Scholar] [CrossRef]
  48. Homer, F.A.; Morrison, R.S.; Brooks, R.R.; Clemens, J.; Reeves, R.D. Comparative studies of nickel, cobalt, and copper uptake by some nickel hyperaccumulators of the genus Alyssum. Plant Soil 1991, 138, 195–205. [Google Scholar] [CrossRef]
  49. Lange, B.; van Der Ent, A.; Baker, A.J.M.; Echevarria, G.; Mahy, G.; Malaisse, F.; Faucon, M.P. Copper and cobalt accumulation in plants: A critical assessment of the current state of knowledge. New Phytol. 2017, 213, 537–551. [Google Scholar] [CrossRef]
  50. Olawuyi, O.J.; Bamigbegbin, B.J.; Bello, O.B. Genetic variation of morphological and yield characters of Celosia argentea L. germplasm. J. Bas. Appl. Res. Int. 2016, 13, 160–169. [Google Scholar]
  51. Hu, D.; Cheng, M.; Hu, K.; Zhang, W.; Yang, Y.; Xu, Q. Evaluation of cobalt hyperaccumulation and tolerance potential of the duckweed (Lemna minor L.). Ecotoxicol. Environ. Saf. 2019, 179, 79–87. [Google Scholar] [CrossRef]
  52. Van Der Ent, A.; Malaisse, F.; Erskine, P.D.; Mesjasz-Przybyłowicz, J.; Przybyłowicz, W.J.; Barnabas, A.D.; Harris, H.H. Abnormal concentrations of Cu–Co in Haumaniastrum katangense, Haumaniastrum robertii and Aeolanthus biformifolius: Contamination or hyperaccumulation? Metallomics 2019, 11, 586–596. [Google Scholar] [CrossRef]
  53. Medwecka-Kornaś, A. Hibiskus rhodanthus (Malvaceae): An interesting plant species in Southeastern Africa. Pol. Bot. J. 2013, 58, 325–335. [Google Scholar] [CrossRef]
  54. Memon, A.R.; Itô, S.; Yatazawa, M. Taxonomic characteristics in accumulating cobalt and nickel in the temperate forest vegetation of central Japan. Soil Sci. Plant Nutr. 1980, 26, 271–280. [Google Scholar] [CrossRef]
  55. Busbee, W.S.; Conner, W.H.; Allen, D.M.; Lanham, J.D. Composition and aboveground productivity of three seasonally flooded depressional forested wetlands in coastal South Carolina. South. Nat. 2003, 2, 335–346. [Google Scholar] [CrossRef]
  56. Küpper, H.; Leitenmaier, B. Cadmium-accumulating plants. Met. Ions Life Sci. 2012, 11, 373–393. [Google Scholar] [CrossRef]
  57. Ali, A.; Guo, D.; Mahar, A.; Ping, W.; Wahid, F.; Shen, F.; Zhang, Z. Phytoextraction and the economic perspective of phytomining of heavy metals. Solid Earth Discuss. 2017, 6, 1–40. [Google Scholar] [CrossRef]
  58. Nedelkoska, T.V.; Doran, P.M. Hyperaccumulation of cadmium by hairy roots of Thlaspi caerulescens. Biotechnol. Bioeng. 2000, 67, 607–615. [Google Scholar] [CrossRef]
  59. Schwartz, C.; Sirguey, C.; Peronny, S.; Reeves, R.D.; Bourgaud, F.; Morel, J.L. Testing of outstanding individuals of Thlaspi caerulescens for cadmium phytoextraction. Int. J. Phytoremediat. 2006, 8, 339–357. [Google Scholar] [CrossRef] [PubMed]
  60. Bert, V.; Bonnin, I.; Saumitou, L.P.; De Laguérie, P.; Petit, D. Do Arabidopsis halleri from nonmetallicolous populations accumulate zinc and cadmium more effectively than those from metallicolous populations? New Phytol. 2002, 155, 47–57. [Google Scholar] [CrossRef] [PubMed]
  61. Deng, D.M.; Deng, J.C.; Li, J.T.; Zhang, J.; Hu, M.; Lin, Z.; Liao, B. Accumulation of zinc, cadmium, and lead in four populations of Sedum alfredii growing on lead/zinc mine spoils. J. Integr. Plant Biol. 2008, 50, 691–698. [Google Scholar] [CrossRef]
  62. Claire-Lise, M.; Nathalie, V. The use of the model species Arabidopsis halleri towards phytoextraction of cadmium polluted soils. New Biotechnol. 2012, 30, 9–14. [Google Scholar] [CrossRef]
  63. Zhao, F.J.; Jiang, R.F.; Dunham, S.J.; McGrath, S.P. Cadmium uptake, translocation and tolerance in the hyperaccumulator Arabidopsis halleri. New Phytol. 2006, 172, 646–654. [Google Scholar] [CrossRef]
  64. Tang, Y.T.; Qiu, R.L.; Zeng, X.W.; Ying, R.R.; Yu, F.M.; Zhou, X.Y. Lead, zinc, cadmium hyperaccumulation and growth stimulation in Arabis ommercial Franch. Environ. Exp. Bot. 2009, 66, 126–134. [Google Scholar] [CrossRef]
  65. Dai, H.; Wei, S.; Skuza, L. Effects of different soil pH and nitrogen fertilizers on Bidens ommer L. Cd accumulation. Environ. Sci. Pollut. Res. 2020, 27, 9403–9409. [Google Scholar] [CrossRef]
  66. Liu, X.; Peng, K.; Wang, A.; Lian, C.; Shen, Z. Cadmium accumulation and distribution in populations of Phytolacca americana L. and the role of transpiration. Chemosphere 2010, 78, 1136–1141. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, W.; Shu, W.; Lan, C. Viola baoshanensis, a plant that hyperaccumulates cadmium. Chin. Sci. Bull. 2004, 49, 29–32. [Google Scholar] [CrossRef]
  68. Nyquist, J.; Greger, M. Response of two wetland plant species to Cd exposure at low and neutral pH. Environ. Exp. Bot. 2009, 65, 417–424. [Google Scholar] [CrossRef]
  69. Lee, K.K.; Cho, H.S.; Moon, Y.C.; Ban, S.J.; Kim, J.Y. Cadmium and lead uptake capacity of energy crops and distribution of metals within the plant structures. KSCE J. Civ. Eng. 2013, 17, 44–50. [Google Scholar] [CrossRef]
  70. Wei, J.L.; Lai, H.Y.; Chen, Z.S. Chelator effects on bioconcentration and translocation of cadmium by hyperaccumulators, Tagetes patula and Impatiens walleriana. Ecotoxicol. Environ. Saf. 2012, 84, 173–178. [Google Scholar] [CrossRef] [PubMed]
  71. Wang, L.; Chen, X.; Guo, W.; Li, Y.; Yan, H.; Xue, X. Yield and Nutritional Quality of Water Spinach (Ipomoea aquatica) as Influenced by Hydroponic Nutrient Solutions with Different pH Adjustments. Int. J. Agric. Biol. 2017, 19, 635. [Google Scholar] [CrossRef]
  72. Liu, Z.; He, X.; Chen, W.; Yuan, F.; Yan, K.; Tao, D. Accumulation and tolerance characteristics of cadmium in a potential hyperaccumulator—Lonicera japonica Thunb. J. Hazard. Mat. 2009, 169, 170–175. [Google Scholar] [CrossRef]
  73. Zhang, S.; Lin, H.; Deng, L.; Gong, G.; Jia, Y.; Xu, X.; Chen, H. Cadmium tolerance and accumulation characteristics of Siegesbeckia orientalis L. Ecol. Eng. 2013, 51, 133–139. [Google Scholar] [CrossRef]
  74. Giannoulis, K.D.; Karyotis, T.; Sakellariou-Makrantonaki, M.; Bastiaans, L.; Struik, P.C.; Danalatos, N.G. Switchgrass biomass partitioning and growth characteristics under different management practices. NJAS Wagening. J. Life Sci. 2016, 78, 61–67. [Google Scholar] [CrossRef]
  75. Marmiroli, M.; Imperiale, D.; Maestri, E.; Marmiroli, N. The response of Populus spp. to cadmium stress: Chemical, morphological and proteomics study. Chemosphere 2013, 93, 1333–1344. [Google Scholar] [CrossRef] [PubMed]
  76. Stolarski, M.J.; Warmiński, K.; Krzyżaniak, M. Energy value of yield and biomass quality of poplar grown in two consecutive 4-year harvest rotations in the north-east of Poland. Energies 2020, 13, 1495. [Google Scholar] [CrossRef]
  77. Tang, Y.; Qiu, R.; Zeng, X.; Fang, X.; Yu, F.; Zhou, X.; Wu, Y. Zn and Cd hyperaccumulating characteristics of Picris divaricata Vant. Int. J. Environ. Pollut. 2009, 38, 26–38. [Google Scholar] [CrossRef]
  78. Kim, T.Y.; Hong, J.H. Nickel Tolerance and the Complexing Role of Histidine in Raphanus sativus. J. Environ. Sci. Int. 2004, 13, 711–719. [Google Scholar] [CrossRef]
  79. Chen, Y.X.; Lin, Q.; Luo, Y.M.; He, Y.F.; Zhen, S.J.; Yu, Y.L.; Wong, M.H. The role of citric acid on the phytoremediation of heavy metal contaminated soil. Chemosphere 2003, 50, 807–811. [Google Scholar] [CrossRef] [PubMed]
  80. Koutroubas, S.D.; Papakosta, D.K.; Doitsinis, A. Adaptation and yielding ability of castor plant (Ricinus communis L.) genotypes in a Mediterranean climate. Eur. J. Agron. 1999, 11, 227–237. [Google Scholar] [CrossRef]
  81. Huang, H.; Yu, N.; Wang, L.; Gupta, D.K.; He, Z.; Wang, K.; Yang, X.E. The phytoremediation potential of bioenergy crop Ricinus communis for DDTs and cadmium co-contaminated soil. Bioresour. Technol. 2011, 102, 11034–11038. [Google Scholar] [CrossRef]
  82. Vollenweider, P.; Cosio, C.; Günthardt-Goerg, M.S.; Keller, C. Localization and effects of cadmium in leaves of a cadmium-tolerant willow (Salix viminalis L.): Part II Microlocalization and cellular effects of cadmium. Environ. Exp. Bot. 2006, 58, 25–40. [Google Scholar] [CrossRef]
  83. Yang, X.E.; Long, X.X.; Ye, H.B.; He, Z.L.; Calvert, D.V.; Stoffella, P.J. Cadmium tolerance and hyperaccumulation in a new Zn-hyperaccumulating plant species (Sedum alfredii Hance). Plant Soil 2004, 259, 181–189. [Google Scholar] [CrossRef]
  84. Khan, A.R.; Ullah, I.; Khan, A.L.; Hong, S.J.; Waqas, M.; Park, G.S.; Shin, J.H. Phytostabilization and physicochemical responses of Korean ecotype Solanum nigrum L. to cadmium contamination. Water Air Soil Pollut. 2014, 225, 1–11. [Google Scholar] [CrossRef]
  85. Marotti, M.; Piccaglia, R.; Biavati, B.; Marotti, I. Characterization and yield evaluation of essential oils from different Tagetes species. J. Essent. Oil Res. 2004, 16, 440–444. [Google Scholar] [CrossRef]
  86. Martínez, M.; Bernal, P.; Almela, C.; Vélez, D.; García-Agustín, P.; Serrano, R.; Navarro-Aviñó, J. An engineered plant that accumulates higher levels of heavy metals than Thlaspi caerulescens, with yields of 100 times more biomass in mine soils. Chemosphere 2006, 64, 478–485. [Google Scholar] [CrossRef] [PubMed]
  87. Perronnet, K.; Schwartz, C.; Morel, J.L. Distribution of cadmium and zinc in the hyperaccumulator Thlaspi caerulescens grown on multicontaminated soil. Plant Soil 2003, 249, 19–25. [Google Scholar] [CrossRef]
  88. Xie, W.Y.; Huang, Q.; Li, G.; Rensing, C.; Zhu, Y.G. Cadmium accumulation in the rootless macrophyte Wolffia ommerc and its potential for phytoremediation. Int. J. Phytoremediat. 2013, 15, 385–397. [Google Scholar] [CrossRef] [PubMed]
  89. Oladoye, P.O.; Olowe, O.M.; Asemoloye, M.D. Phytoremediation technology and food security impacts of heavy metal contaminated soils: A review of literature. Chemosphere 2022, 288, 132555. [Google Scholar] [CrossRef] [PubMed]
  90. Rascio, N. Metal accumulation by some plants growing on zinc-mine deposits. Oikos 1977, 29, 250–253. [Google Scholar] [CrossRef]
  91. Reeves, R.D.; Brooks, R.R. Hyperaccumulation of lead and zinc by two metallophytes from mining areas of Central Europe. Environ. Pollut. Ser. A Ecol. Biol. 1983, 31, 277–285. [Google Scholar] [CrossRef]
  92. Baker, A.J.M.; Reeves, R.D.; Hajar, A.S.M. Heavy metal accumulation and tolerance in British populations of the metallophyte Thlaspi caerulescens J. & C. Presl (Brassicaceae). New Phytol. 1994, 127, 61–68. [Google Scholar] [CrossRef]
  93. Brown, S.L.; Chaney, R.L.; Angle, J.S.; Baker, A.J.M. Zinc and cadmium uptake by hyperaccumulator Thlaspi caerulescens grown in nutrient solution. Soil Sci. Soc. Am. J. 1995, 59, 125–133. [Google Scholar] [CrossRef]
  94. Zhao, F.J.; Lombi, E.; Breedon, T.M.S.P. Zinc hyperaccumulation and cellular distribution in Arabidopsis halleri. Plant Cell Environ. 2000, 23, 507–514. [Google Scholar] [CrossRef]
  95. Hu, P.J.; Qiu, R.L.; Senthilkumar, P.; Jiang, D.; Chen, Z.W.; Tang, Y.T.; Liu, F.J. Tolerance, accumulation and distribution of zinc and cadmium in hyperaccumulator Potentilla griffithii. Environ. Exp. Bot. 2009, 66, 317–325. [Google Scholar] [CrossRef]
  96. Grimm, M. Feasibility of Brassica juncea as a Hyperaccumulator in Phytomining of Cu and Zn. In Proceedings of the EGU General Assembly Conference Abstracts, Vienna, Austria, 23–27 May 2022. [Google Scholar] [CrossRef]
  97. Grison, C.M.; Mazel, M.; Sellini, A.; Escande, V.; Biton, J.; Grison, C. The leguminous species Anthyllis vulneraria as a Zn-hyperaccumulator and eco-Zn catalyst resources. Environ. Sci. Pollut. Res. 2015, 22, 5667–5676. [Google Scholar] [CrossRef]
  98. Peer, W.A.; Mahmoudian, M.; Freeman, J.L.; Lahner, B.; Richards, E.L.; Reeves, R.D.; Salt, D.E. Assessment of plants from the Brassicaceae family as genetic models for the study of nickel and zinc hyperaccumulation. New Phytol. 2006, 172, 248–260. [Google Scholar] [CrossRef]
  99. Reeves, R.D.; Baker, A.J.; Jaffré, T.; Erskine, P.D.; Echevarria, G.; van Der Ent, A. A global database for plants that hyperaccumulate metal and metalloid trace elements. New Phytol. 2018, 218, 407–411. [Google Scholar] [CrossRef]
  100. Balafrej, H.; Bogusz, D.; Triqui, Z.E.A.; Guedira, A.; Bendaou, N.; Smouni, A.; Fahr, M. Zinc hyperaccumulation in plants: A review. Plants 2020, 9, 562. [Google Scholar] [CrossRef] [PubMed]
  101. Lin, W.; Xiao, T.; Wu, Y.; Ao, Z.; Ning, Z. Hyperaccumulation of zinc by Corydalis davidii in Zn-polluted soils. Chemosphere 2012, 86, 837–842. [Google Scholar] [CrossRef] [PubMed]
  102. Shimwell, D.W.; Laurie, A.E. Lead and zinc contamination of vegetation in the Southern Pennines. Environ. Pollut. 1972, 3, 291–301. [Google Scholar] [CrossRef]
  103. Reeves, R.D.; Brooks, R.R. European species of Thlaspi L. (Cruciferae) as indicators of nickel and zinc. J. Geochem. Explor. 1983, 18, 275–283. [Google Scholar] [CrossRef]
  104. Yang, X.E.; Long, X.; Ni, W.; Fu, C. Sedum alfredii H: A new Zn hyperaccumulating plant first found in China. Chin. Sci. Bull. 2002, 47, 1634–1637. [Google Scholar] [CrossRef]
  105. Qiu, R.; Fang, X.; Tang, Y.; Du, S.; Zeng, X.; Brewer, E. Zinc hyperaccumulation and uptake by Potentilla griffithii Hook. Int. J. Phytoremediat. 2006, 8, 299–310. [Google Scholar] [CrossRef]
  106. Wu, C.; Liao, B.; Wang, S.L.; Zhang, J.; Li, J.T. Pb and Zn accumulation in a Cd-hyperaccumulator (Viola baoshanensis). Int. J. Phytoremediat. 2010, 12, 574–585. [Google Scholar] [CrossRef] [PubMed]
  107. Losfeld, G.; L’huillier, L.; Fogliani, B.; Coy, S.M.; Grison, C.; Jaffré, T. Leaf-age and soil-plant relationships: Key factors for reporting trace-elements hyperaccumulation by plants and design applications. Environ. Sci. Pollut. Res. 2015, 22, 5620–5632. [Google Scholar] [CrossRef] [PubMed]
  108. Jaffré, T. Accumulation Du Manganèse Par Les Proteacées de Nouvelle-Calédonie. Acad. Sci. 1979, 289, 425–428. [Google Scholar]
  109. Mizuno, T.; Asahina, R.; Hosono, A.; Tanaka, A.; Senoo, K.; Obata, H. Age-dependent manganese hyperaccumulation in Chengiopanax sciadophylloides (Araliaceae). J. Plant Nutr. 2008, 31, 1811–1819. [Google Scholar] [CrossRef]
  110. Fernando, D.R.; Marshall, A.; Baker, A.J.; Mizuno, T. Microbeam methodologies as powerful tools in manganese hyperaccumulation research: Present status and future directions. Front. Plant Sci. 2013, 4, 319. [Google Scholar] [CrossRef]
  111. Fernando, D.R.; Guymer, G.; Reeves, R.D.; Woodrow, I.E.; Baker, A.J.; Batianoff, G.N. Foliar Mn accumulation in eastern Australian herbarium specimens: Prospecting for ‘new’Mn hyperaccumulators and potential applications in taxonomy. Ann. Bot. 2009, 103, 931–939. [Google Scholar] [CrossRef]
  112. Abubakari, F.; Nkrumah, P.N.; Erskine, P.D.; Brown, G.K.; Fernando, D.R.; Echevarria, G.; van Der Ent, A. Manganese (hyper) accumulation within Australian denhamia (Celastraceae): An assessment of the trait and manganese accumulation under controlled conditions. Plant Soil 2021, 463, 205–223. [Google Scholar] [CrossRef]
  113. Liu, J.; Shang, W.; Zhang, X.; Zhu, Y.; Yu, K. Mn accumulation and tolerance in Celosia argentea Linn: A new Mn-hyperaccumulating plant species. J. Hazard. Mater. 2014, 267, 136–141. [Google Scholar] [CrossRef]
  114. Bidwell, S.D.; Woodrow, I.E.; Batianoff, G.N.; Sommer-Knudsen, J. Hyperaccumulation of manganese in the rainforest tree Austromyrtus bidwillii (Myrtaceae) from Queensland, Australia. Funct. Plant Biol. 2002, 29, 899–905. [Google Scholar] [CrossRef]
  115. Xue, S.G.; Chen, Y.X.; Reeves, R.D.; Baker, A.J.; Lin, Q.; Fernando, D.R. Manganese uptake and accumulation by the hyperaccumulator plant Phytolacca ommerc Roxb. (Phytolaccaceae). Environ. Pollut. 2004, 131, 393–399. [Google Scholar] [CrossRef]
  116. Deng, H.; Li, M.; Chen, Y.; Luo, Y.; Yu, F. A new discovered manganese hyperaccumulator-Polygonum pubescens Blume. Fresenius Environ. Bull. 2010, 19, 94–99. [Google Scholar]
  117. Liu, K.; Yu, F.; Chen, M.; Zhou, Z.; Chen, C.; Li, M.S.; Zhu, J. A newly found manganese hyperaccumulator—Polygonum lapathifolium Linn. Int. J. Phytoremediat. 2016, 18, 348–353. [Google Scholar] [CrossRef]
  118. Yang, S.X.; Deng, H.; Li, M.S. Manganese uptake and accumulation in a woody hyperaccumulator, Schima superba. Plant Soil Environ. 2008, 54, 441–446. [Google Scholar] [CrossRef]
  119. White, P.J. Selenium accumulation by plants. Ann. Bot. 2016, 117, 217–235. [Google Scholar] [CrossRef] [PubMed]
  120. Parker, D.R.; Feist, L.J.; Varvel, T.W.; Thomason, D.N.; Zhang, Y. Selenium phytoremediation potential of Stanleya pinnata. Plant Soil 2003, 249, 157–165. [Google Scholar] [CrossRef]
  121. Reeves, R.D. Metal-Accumulating Plants. In Phytoremediat. Of Toxic Metals: Using Plants to Clean Up the Environment; Ensley, B.D., Ed.; John Wiley & Son: London, UK, 2000; pp. 193–229. [Google Scholar]
  122. Moon, D.M. Yield potential of Asparagus officinalis L. N. Z. J. Exp. Agric. 1976, 4, 435–438. [Google Scholar] [CrossRef]
  123. Nadaf, S.K.; Al-Khamisi, S.A.; AI-Lawati, A.H.; Ibrahim, Y.M.; EI-Hag, M.G. Productivity of saltbush (Atriplex) species under saline soil and water conditions in Oman. J. Agric. Mar. Sci. 1999, 4, 41–46. [Google Scholar] [CrossRef]
  124. Nedjimi, B. Measurement of selenium in two Algerian chenopods (Atriplex canescens (Pursh.) Nutt. And Suaeda ommercia (Linn.) Forssk). Measurement 2018, 129, 256–259. [Google Scholar] [CrossRef]
  125. Ma, Y.; Yin, J.; Wang, J.; Liu, X.; He, J.; Zhang, R.; Wu, M. Selenium speciation and volatile flavor compound profiles in the edible flowers, stems, and leaves of selenium-hyperaccumulating vegetable Cardamine violifolia. Food Chem. 2023, 427, 136710. [Google Scholar] [CrossRef]
  126. Huang, C.; Ying, H.; Yang, X.; Gao, Y.; Li, T.; Wu, B.; Wang, L. The Cardamine enshiensis genome reveals whole genome duplication and insight into selenium hyperaccumulation and tolerance. Cell Discov. 2021, 7, 62. [Google Scholar] [CrossRef]
  127. O’Donohue, B.; Hiti-Bandaralage, J.; Gleeson, M.; O’Brien, C.; Harvey, M.A.; van der Ent, A.; Hayward, A. Tissue culture tools for selenium hyperaccumulator Neptunia amplexicaulis for development in phytoextraction. Nat. Prod. Bioprosp. 2022, 12, 28. [Google Scholar] [CrossRef]
  128. Pinto Irish, K.; Harvey, M.A.; Erskine, P.D.; van der Ent, A. Root foraging and selenium uptake in the Australian hyperaccumulator Neptunia amplexicaulis and non-accumulator Neptunia gracilis. Plant Soil 2021, 462, 219–233. [Google Scholar] [CrossRef]
  129. Harris, J.; Schneberg, K.A.; Pilon-Smits, E.A. Sulfur–selenium–molybdenum interactions distinguish selenium hyperaccumulator Stanleya pinnata from non-hyperaccumulator Brassica juncea (Brassicaceae). Planta 2014, 239, 479–491. [Google Scholar] [CrossRef]
  130. Leblanc, M.; Petit, D.; Deram, A.; Robinson, B.H.; Brooks, R.R. The phytomining and environmental significance of hyperaccumulation of thallium by Iberis intermedia from southern France. Econ. Geol. 1999, 94, 109–113. [Google Scholar] [CrossRef]
  131. Sinha, R.; Singh, A.K.; Bauddh, K.; Sharma, T.R.; Sharma, P. Phytomining: A sustainable approach for recovery and extraction of valuable metals. In Phytorestoration of Abandoned Mining and Oil Drilling Sites; Bauddh, K., Korstad, J., Sharma, P., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 487–506. [Google Scholar] [CrossRef]
  132. Mohan, B.S. Phytomining of gold. Curr. Sci. 2005, 88, 1021–1022. [Google Scholar]
  133. Corzo, R.A.; Chaney, R.L.; Baker, A.J.; Edraki, M.; Erskine, P.D.; Echevarria, G.; van der Ent, A. Phytoextraction of high value elements and contaminants from mining and mineral wastes: Opportunities and limitations. Plant Soil 2020, 449, 11–37. [Google Scholar] [CrossRef]
  134. Anderson, C.W.; Brooks, R.R.; Stewart, R.B.; Simcock, R. Harvesting a crop of gold in plants. Nature 1998, 395, 553–554. [Google Scholar] [CrossRef]
  135. Wilson, C.V.; Anderson, C.W.; Rodriguez, L.M. Gold phytomining. A review of the relevance of this technology to mineral extraction in the 21st century. J. Environ. Manag. 2012, 111, 249–257. [Google Scholar] [CrossRef]
  136. Msuya, F.A.; Brooks, R.R.; Anderson, C.W. Chemically-induced uptake of gold by root crops: Its significance for phytomining. Gold Bull. 2000, 33, 134–137. [Google Scholar] [CrossRef]
  137. Lamb, A.E.; Anderson, C.W.N.; Haverkamp, R.G. The extraction of gold from plants and its application to phytomining. Chem. N. Z. 2001, 65, 31–33. [Google Scholar]
  138. Vural, A.; Safari, S. Phytoremediation ability of Helichrysum arenarium plant for Au and Ag: Case study at Demirören village (Gümüşhane, Turkey). Gold Bull. 2022, 55, 129–136. [Google Scholar] [CrossRef]
  139. Harris, A.T.; Bali, R. On the formation and extent of uptake of silver nanoparticles by live plants. J. Nanopart. Res. 2008, 10, 691–695. [Google Scholar] [CrossRef]
  140. Fuchs, W.A.; Rose, A.W. The geochemical behavior of platinum and palladium in the weathering cycle in the Stillwater Complex, Montana. Econ. Geol. 1974, 69, 332–346. [Google Scholar] [CrossRef]
  141. Kothny, E.L. Palladium in plant ash. Plant Soil 1979, 53, 547–550. [Google Scholar] [CrossRef]
  142. Nemutandani, T.; Dutertre, D.; Chimuka, L.; Cukrowska, E.; Tutu, H. The potential of Berkheya coddii for phytoextraction of nickel, platinum, and palladium contaminated sites. Toxicol. Environ. Chem. 2006, 88, 175–185. [Google Scholar] [CrossRef]
  143. Walton, D. The Phytoextraction of Gold and Palladium from Mine Tailings. Ph.D. Thesis, Massey University, Wellington, New Zealand, 2002. [Google Scholar]
  144. Aquan, H.M. Phytoextraction of Palladium and Gold from Broken Hill Gossan. Ph.D. Thesis, Massey University, Manawatū, New Zealand, 2015. [Google Scholar]
  145. Harumain, Z.A.S.; Parker, H.L.; Austin, M.J.; McElroy, M.C.; Andrea, M.G.; Robert, M.C.; WN, A.C.; Luca, C.; Graedel, T.E.; Rylott, E.L.; et al. Toward Financially Viable Phytoextraction and Production of Plant-Based Palladium Catalysts. Environ. Sci. Technol. 2017, 51, 2992–3000. [Google Scholar] [CrossRef]
  146. Kovacs, H.; Dobo, Z.; Koos, T.; Gyimesi, A.; Nagy, G. Influence of the flue gas temperature on the behavior of metals during biomass combustion. Energy Fuels 2018, 32, 7851–7856. [Google Scholar] [CrossRef]
  147. Diehl, D.B.; Gagnon, Z.E. Interactions between essential nutrients with platinum group metals in submerged aquatic and emergent plants. Water Air Soil Pollut. 2007, 184, 255–267. [Google Scholar] [CrossRef]
  148. Kińska, K.; Kowalska, J. Comparison of platinum, rhodium, and palladium bioaccumulation by Sinapis alba and their influence on phytochelatin synthesis in plant tissues. Pol. J. Environ. Stud. 2019, 28, 1735–1740. [Google Scholar] [CrossRef]
  149. Bonanno, G. Trace element accumulation and distribution in the organs of Phragmites australis (common reed) and biomonitoring applications. Ecotoxicol. Environ. Saf. 2011, 74, 1057–1064. [Google Scholar] [CrossRef]
  150. Pujari, G.N. Phytomining of Low Grade Ore. SGAT Bull. 2018, 19, 15–22. [Google Scholar]
  151. Custos, J.M.; Moyne, C.; Sterckeman, T. How root nutrient uptake affects rhizosphere pH: A modelling study. Geoderma 2020, 369, 114314. [Google Scholar] [CrossRef]
  152. Chaney, R.L.; Angle, J.S.; Broadhurst, C.L.; Peters, C.A.; Tappero, R.V.; Sparks, D.L. Improved understanding of hyperaccumulation yields commercial phytoextraction and phytomining technologies. J. Environ. Qual. 2007, 36, 1429–1443. [Google Scholar] [CrossRef]
  153. Zhong, X.; Chen, Z.; Li, Y.; Ding, K.; Liu, W.; Liu, Y.; Qiu, R. Factors influencing heavy metal availability and risk assessment of soils at typical metal mines in Eastern China. J. Hazard. Mater. 2020, 400, 123289. [Google Scholar] [CrossRef] [PubMed]
  154. Yan, A.; Wang, Y.; Tan, S.N.; Mohd Yusof, M.L.; Ghosh, S.; Chen, Z. Phytoremediation: A promising approach for revegetation of heavy metal-polluted land. Front. Plant Sci. 2020, 11, 359. [Google Scholar] [CrossRef] [PubMed]
  155. Bennett, F.A.; Tyler, E.K.; Brooks, R.R.; Gregg, P.E.H.; Stewart, R.B. Fertilisation of hyperaccumulators to enhance their potential for phytoremediation and phytomining. In Plants That Hyperaccumulate Heavy Metals; Brooks, R.R., Ed.; CAB International: New York, NY, USA, 1998; pp. 249–259. [Google Scholar]
  156. Sheoran, V.; Sheoran, A.S.; Poonia, P. Factors affecting phytoextraction: A review. Pedosphere 2016, 26, 148–166. [Google Scholar] [CrossRef]
  157. Schmidt, U. Enhancing phytoextraction: The effect of chemical soil manipulation on mobility, plant accumulation, and leaching of heavy metals. J. Environ. Qual. 2003, 32, 1939–1954. [Google Scholar] [CrossRef]
  158. Băbău, A.M.C.; Micle, V.; Damian, G.E.; Sur, I.M. Sustainable ecological restoration of sterile dumps using Robinia pseudoacacia. Sustainbility 2021, 13, 14021. [Google Scholar] [CrossRef]
  159. Ebbs, S.D.; Lasat, M.M.; Brady, D.J.; Cornish, J.; Gordon, R.; Kochian, L.V. Phytoextraction of Cadmium and Zinc from a Contaminated Soil. J. Environ. Qual. 1997, 26, 1424–1430. [Google Scholar] [CrossRef]
  160. Barona, A.; Aranguiz, I.; Elıas, A. Metal associations in soils before and after EDTA extractive decontamination: Implications for the effectiveness of further clean-up procedures. Environ. Pollut. 2001, 113, 79–85. [Google Scholar] [CrossRef]
  161. Evangelou, M.W.; Ebel, M.; Schaeffer, A. Chelate assisted phytoextraction of heavy metals from soil. Effect, mechanism, toxicity, and fate of chelating agents. Chemosphere 2007, 68, 989–1003. [Google Scholar] [CrossRef]
  162. Robinson, B.H.; Brooks, R.R.; Howes, A.W.; Kirkman, J.H.; Gregg, P.E.H. The potential of the high-biomass nickel hyperaccumulator Berkheya coddii for phytoremediation and phytomining. J. Geochem. Explor. 1997, 60, 115–126. [Google Scholar] [CrossRef]
  163. Meers, E.; Ruttens, A.; Hopgood, M.J.; Samson, D.; Tack, F.M.G. Comparison of EDTA and EDDS as potential soil amendments for enhanced phytoextraction of heavy metals. Chemosphere 2005, 58, 1011–1022. [Google Scholar] [CrossRef] [PubMed]
  164. Blaylock, M.J.; Salt, D.E.; Dushenkov, S.; Zakharova, O.; Gussman, C.; Kapulnik, Y.; Raskin, I. Enhanced accumulation of Pb in Indian mustard by soil-applied chelating agents. Environ. Sci. Technol. 1997, 31, 860–865. [Google Scholar] [CrossRef]
  165. Wang, L.; Hou, D.; Shen, Z.; Zhu, J.; Jia, X.; Ok, Y.S.; Rinklebe, J. Field trials of phytomining and phytoremediation: A critical review of influencing factors and effects of additives. Crit. Rev. Environ. Sci. Technol. 2020, 50, 2724–2774. [Google Scholar] [CrossRef]
  166. Rascio, N.; Navari-Izzo, F. Heavy metal hyperaccumulating plants: How and why do they do it? And what makes them so interesting? Plant Sci. 2011, 180, 169–181. [Google Scholar] [CrossRef] [PubMed]
  167. Gendre, D.; Czernic, P.; Conéjéro, G.; Pianelli, K.; Briat, J.F.; Lebrun, M.; Mari, S. TcYSL3, a member of the YSL gene family from the hyper-accumulator Thlaspi caerulescens, encodes a nicotianamine-Ni/Fe transporter. Plant J. 2007, 49, 1–15. [Google Scholar] [CrossRef]
  168. Baruah, S.; Bora, M.S.; Sharma, P.; Deb, P.; Sarma, K.P. Understanding of the distribution, translocation, bioaccumulation, and ultrastructural changes of Monochoria hastata plant exposed to cadmium. Water Air Soil Pollut. 2017, 228, 1–21. [Google Scholar] [CrossRef]
  169. Hu, X.; Gholizadeh, M. Biomass pyrolysis: A review of the process development and challenges from initial researches up to the ommercialization stage. J. Energy Chem. 2019, 39, 109–143. [Google Scholar] [CrossRef]
  170. Zhang, X.; Houzelot, V.; Bani, A.; Morel, J.L.; Echevarria, G.; Simonnot, M.O. Selection and combustion of Ni-hyperaccumulators for the phytomining process. Int. J. Phytoremediat. 2014, 16, 1058–1072. [Google Scholar] [CrossRef]
  171. Dastyar, W.; Raheem, A.; He, J.; Zhao, M. Biofuel production using thermochemical conversion of heavy metal-contaminated biomass (HMCB) harvested from phytoextraction process. J. Chem. Eng. 2019, 358, 759–785. [Google Scholar] [CrossRef]
  172. Hazotte, C.; Laubie, B.; Pacault, S.; Dufaud, O.; Simonnot, M.O. Evaluation of the performance of nickel hyperaccumulator plants as combustion fuel. Biomass Bioenergy 2020, 140, 105671. [Google Scholar] [CrossRef]
  173. Soundararajan, R.K.V.; Varjani, S.; Chandrasekaran, R.; Kandasamy, D. Recovery of Gold and Other Precious Metals by Biosorption. In Sustainable Resource Management; Guo, W., Ngo, H.H., Surampalli, R.Y., Zhang, T.C., Eds.; John Wiley & Sons: Weinheim, Germany, 2021; Volume 2, pp. 463–488. [Google Scholar] [CrossRef]
  174. Won, S.W.; Kotte, P.; Wei, W.; Lim, A.; Yun, Y.S. Biosorbents for recovery of precious metals. Bioresour. Technol. 2014, 160, 203–212. [Google Scholar] [CrossRef] [PubMed]
  175. U.S. Department of the Interior. Mineral Commodity Summaries. U.S. Geological Survey 2003; U.S. Department of the Interior: Reston, VA, USA, 2003; p. 210. [CrossRef]
  176. Lee, J.; Kaunda, R.B.; Sinkala, T.; Workman, C.F.; Bazilian, M.D.; Clough, G. Phytoremediation and phytoextraction in Sub-Saharan Africa: Addressing economic and social challenges. Ecotoxicol. Environ. Saf. 2021, 226, 112864. [Google Scholar] [CrossRef]
  177. Anderson, C.W.N.; Brooks, R.R.; Chiarucci, A.; LaCoste, C.J.; Leblanc, M.; Robinson, B.H.; Stewart, R.B. Phytomining for nickel, thallium and gold. J. Geochem. Explor. 1999, 67, 407–415. [Google Scholar] [CrossRef]
  178. Nkrumah, P.N.; Baker, A.J.; Chaney, R.L.; Erskine, P.D.; Echevarria, G.; Morel, J.L.; van der Ent, A. Current status and challenges in developing nickel phytomining: An agronomic perspective. Plant Soil 2016, 406, 55–69. [Google Scholar] [CrossRef]
Figure 1. Main stages in the phytomining process for metal extraction.
Figure 1. Main stages in the phytomining process for metal extraction.
Soilsystems 08 00008 g001
Figure 2. Metal price trends of the last decade.
Figure 2. Metal price trends of the last decade.
Soilsystems 08 00008 g002
Table 1. Minimum threshold concentration of a metal in plants so that the species may be regarded as a hyperaccumulator. The accumulation refers to vegetative parts which can be harvested, most likely aerial biomass.
Table 1. Minimum threshold concentration of a metal in plants so that the species may be regarded as a hyperaccumulator. The accumulation refers to vegetative parts which can be harvested, most likely aerial biomass.
ElementHyperaccumulation Threshold (mg kg−1)Number of Hyperaccumulator Plant SpeciesReference
Ag1-Sheoran et al. [5]
Au1-Sheoran et al. [5]
Cd10024Van der Ent et al. [18]
Co30029Krämer et al. [19]
Mn10,00022Baker and Brooks [17]
Ni1000532Soleymanifar et al. [3]
Se10020Reeves [20]
Tl1001Sheoran et al. [5]
Zn300024Van der Ent et al. [18]
Table 3. List of all Co hyperaccumulators. Concentrations in plants refer to leaves unless otherwise stated.
Table 3. List of all Co hyperaccumulators. Concentrations in plants refer to leaves unless otherwise stated.
Plant SpeciesFamilyCo (mg kg−1 DW)/Biomass Yield (t ha−1)Experimental Conditions—CountryReferences
Acalypha cupricolaEuphorbiaceae207–904Ex situ; plant analysis; samples from Co/Cu mine sites; Equatorial climate; Katanga, CongoFaucon et al. [46]
Aeollanthus subacaulisLamiaceae0–1900Pot experiment (200 g sterilized soil); soil containing 10,000 mg Co kg−1 as nitrate; CongoMorisson et al. [47]
Alectra sessilifloraOrobanchaceae21–866Ex situ; plant analysis; samples collected from Co/Cu mine sites; Equatorial climate; Katanga, CongoFaucon et al. [46]
Alyssum corsicumBrassicaceae1080 (shoot)/4.2–10Field experiment in various soil types; 150 kg N ha−1, 100 kg P ha−1, and 75 kg K ha−1 were applied; Beltsville, USAMalik et al. [44]
Alyssum heldreichiiBrassicaceae1000Pot experiment; plants spiked with different concentrations of metals; New ZealandHomer et al. [48]
Alyssum muraleBrassicaceae2070 (shoot)/10.2–20Field experiment in various soil types; 150 kg N ha−1, 100 kg P ha−1, and 75 kg K ha−1 were applied; Beltsville, USAMalik et al. [44]
Alyssum pintodasilvaeBrassicaceae1000Pot experiment; plants spiked with different concentrations of metals; New ZealandHomer et al. [48]
Alyssum teniumBrassicaceae1000Pot experiment; plants spiked with different concentrations of metals; New ZealandHomer et al. [48]
Alyssum troodiiBrassicaceae10–2325Pot experiment; plants spiked with different concentrations of metals; CyprusLange et al. [49]
Anisopappus chinensisAsteraceae3–1300Ex situ; leaf analysis; Equatorial climate; CongoLange et al. [49]
Aporosa chalarocarpaPhyllanthaceae468Ex situ; leaf analysis; samples collected from ultramafic rocks; Tropical climate; Sabah, MalaysiaLange et al. [49]
Berkeya coddiiAsteraceae40–2116/4.96–9.30Experiment in controlled climate glasshouses; plants spiked with 4–500 mg Co kg−1; New ZealandKeeling et al. [33]
Rue et al. [41]
Buchnera henriquesiiOrobanchaceae404–930Ex situ; plant analysis; plant samples collected from Co/Cu mine sites; Equatorial climate; Katanga, CongoFaucon et al. [46]
Celosia trigyna L.Amaranthaceae75–501/9.5Ex situ; plant analysis; plant samples collected from Co/Cu mine sites; Equatorial climate; Katanga, CongoOlawuyi et al. [50]
Crepidorhopalon perennisLinderniaceae61–1105/0.32Ex situ; plant analysis; plant samples collected from Co/Cu mine sites; Equatorial climate; Katanga, CongoFaucon et al. [46]
Crepidorhopalon tenuisLinderniaceae8–605/0.16–0.32Ex situ; plant analysis; plant samples collected from Co/Cu mine sites; Equatorial climate; Katanga, CongoFaucon et al. [46]
Lemna minor L.Araceae5000Pre-cultivated under laboratory conditions; fertilized with Hoagland solution; plants spiked with 0–200 μΜ CoCl2·6H2O; Nanjing, ChinaHu et al. [51]
Glochidion cf. sericeumPhyllanthaceae442–1310Ex situ; leaf analysis; leaf samples collected from ultramafic rocks; Tropical climate; Sabah, MalaysiaLange et al. [49]
Haumaniastrum katangenseLamiaceae864Ex situ; leaf analysis; leaf samples collected from Cu/Co-Zn/Cu mine sites; background Co levels 1.2–85 mg kg−1; CongoVan der Ent et al. [52]
Haumaniastrum robertiiLamiaceae6160/20Ex situ; leaf analysis; leaf samples collected from Cu/Co-Zn/Cu mine sites; background Co levels 1.2–85 mg kg−1; CongoLange et al. [49]
Hibiscus rhodanthusMalvaceae42–274/0.054–1.6Ex situ; plant analysis; Sub-tropical climate; Nyika Plateu, Zambia and MozambiqueMedwecka [53]
Nyssa sylvatica var. bifloraNyssaceae0–438/6Ex situ; plant analysis; plant samples from granite derivatives; available Co content in soil—2.5%; Inabu-cho, JapanMemon and Yatazawa [54]
Nyssa sylvatica var. sylvaticaNyssaceae800/6Ex situ; plant analysis; South Carolina, USABusbee et al. [55]
Rinorea bengalensisViolaceae0.5–1200/30–40Ex situ; leaf analysis; ultramafic rocks; Tropical climate; Halmahera Island, IndonesiaHamdan et al. [39]
Rinorea javanicaViolaceae3–670/30–40Ex situ; leaf analysis; leaf samples collected from ultramafic rocks; Tropical climate; Halmahera Island, IndonesiaHamdan et al. [39]
Triumfetta welwitschiiMalvaceae162–1971Ex situ; plant analysis; plant samples collected from Co/Cu mine sites; Equatorial climate; Katanga, CongoFaucon et al. [46]
Vernoniastrum latifoliumAsteraceae82–549Ex situ; plant analysis; Equatorial climate; CongoLange et al. [49]
Vigna dolomiticaFabaceae241–540Ex situ; plant analysis; Equatorial climate; CongoLange et al. [49]
Table 4. List of Cd hyperaccumulating plant species.
Table 4. List of Cd hyperaccumulating plant species.
Plant SpeciesFamilyAchieved Concentration (mg kg−1)/Biomass Yield (t ha−1)Experimental Conditions—CountryReferences
Arabidopsis halleriBrassicaceae117.2Pot experiment; nutrient solution applied as fertilizer; plants exposed to 0–100 μΜ CdSO4; Blankerode, GermanyZhao et al. [63]
Arabis paniculataBrassicaceae457Hydroponics experiment; Hoagland solution applied as fertilizer; plants spiked with 0–267 μΜ CdCl2·2.5H2O; Yunnan Province, ChinaTang et al. [64]
Bidens pilosaAsteraceae400.7/5–6.7Field experiment with total Cd concentration in soil 19.63 mg kg-1; (NH4)2SO4 and Ca(NO3)2 were added at 200 mg kg−1 as fertilizers; Shaanxi, ChinaDai et al. [65]
Centela asiaticaApiaceae105Field and laboratory experiment; field pseudo-total Cd concentration—124–1000 mg kg−1, plants were spiked at levels 0–5 mg L−1 in lab; South ChinaLiu et al. [66]
Viola baoshanensisCyperaceae1168Field survey and greenhouse experiment; nutrient solution was applied as fertilizer; plants spiked with 0–30 mg L−1; Humid sub-tropical climate; Hunan Province, ChinaLiu et al. [67]
Elodea canadensisHydrocharitaceae300Hydroponics experiment; plants exposed to 0–0.5 μΜ CdCl2; Stockholm, SwedenAli et al. [57]
Nyquist and Greger [68]
Helianthus annusHelianthoideae230/4.5Pot experiment; background total Cd concentration in soil—30 mg kg−1; pot spiked with 0–500 mg Cd L−1 solution; Tropical climate; KoreaLee et al. [69]
Impatiens wallerianaBalsaminaceae1168Pot experiment; containing total Cd—0.61 mg kg−1; soil spiked with 20, 40, and 80 mg Cd(NO3) kg−1; ChinaWei et al. [70]
Ipomoea aquaticaConvolvulaceae138/7–30Hydroponics experiment; nutrient solution containing 1 M of HNO3, H3PO4, and H2SO4 was added; Beijing, ChinaWang et al. [71]
Lonicera japonicaCaprifoliaceae470.25/2.1–3.8Hydroponics experiment; plant samples were collected from uncontaminated site; Hoagland solution was applied as fertilizer; plants were spiked with 0–50 mg CdCl2·2.5H2O L−1; ChinaLiu et al. [72]
Siegesbeckia orientalis L.Asteraceae192.92–6762Field experiment and pot experiment; total Cd concentration 28.44 and 150 mg kg−1 in soil and pots respectively; Huidong, ChinaZhang et al. [73]
Panicum virgatumPoaceae280/18Field experiment; N-fertilization ranged from 0–240 kg ha−1; Mediterranean climate; Thessaly, GreeceGiannoulis et al. [74]
Phytolacca americanaPhytolaccaceae42–637.7/0.53Field experiment and hydroponic experiment; pseudo-total concentration in soil 1–1000 mg kg−1, plants spiked with 0–100 μΜ CdCl2; Hoagland solution applied as fertilizer; ChinaLiu et al. [66]
Potentilla griffithiRosaceae174Field experiment; contaminated soil contained a total of 331 mg Cd kg−1; Monsoon climate; Lanping, ChinaHu et al. [51]
Populus nigraSalicaceae96.8/82Hydroponics experiment; plant cuttings were exposed to 0–20 μΜ CdSO4; Rome, ItalyMarmiroli et al. [75]
Stolarski et al. [76]
Picris divaricataAsteraceae1343Pot experiment with contaminated soil; total Cd amount in soil—15 mg kg−1; Monsoon climate; Lanping, ChinaTang et al. [77]
Raphanus sativusBrassicaceae326.75/33–42Hydroponics experiment; 400 μmol L−1 CaCl2 was added; Korea, and Jiangxi, ChinaKim and Hong [78]
Chen et al. [79]
Ricinus communisEuphorbiaceae288/2–2.6Ex situ; plant analysis; plant sampling from Cd contaminated sites; Cd concentration on soil—2.8 mg kg−1; Subtropical climate; Zhejiang, ChinaKoutroubas et al. [80]
Huang et al. [81]
Salix viminalisSalicaceae200/12Hydroponics experiment; plant material spiked at 0, 10, 50, 200 μΜ CdCl2; Birmensdorf, SwitzerlandStolarski et al. [76] Vollenweider et al. [82]
Sedum alfrediiCrassulaceae6500–9000Greenhouse experiment in pots; nutrient solution was applied; plants were treated with 12.5–800 μmol Cd L−1; Quzhou, ChinaYang et al. [83]
Solanum nigrum L.Solanaceae307–3163Pot experiment; soil in pots spiked with various Cd concentrations (as CdCl2) up to 80 mg kg−1; Daegu, KoreaKhan et al. [84]
Tagetes patulaAsteraceae324/26.8Pot experiment; containing total Cd—0.61 mg kg−1; soil spiked with 20, 40, and 80 mg Cd(NO3) kg−1; ChinaMarotti et al. [85]
Wei et al. [70]
Thlaspi caerulescensBrassicaceae380–7400/1.6Pot experiment containing contaminated soil—total Cd 20.3 mg kg−1; fertilized with 13 mg N, 33 mg P, and 41 mg K kg−1 soil; Northern FranceMartínez et al. [86]; Perronnet et al. [87]
Wolffia globosaAraceae500/0.08Hydroponics experiment in greenhouse; Hoagland solution applied as fertilizer; solution spiked at 0–100 μΜ Cd(NO3); ChinaXie et al. [88]
Table 5. Zn hyperaccumulating plant species.
Table 5. Zn hyperaccumulating plant species.
Plant SpeciesFamilyAchieved Concentrations (%) (Aerial Parts Are Specified)/Biomass Yield (t ha−1)Experimental Conditions—CountryReferences
Anthyllis vulnerariaFabaceae>1% (leaf)/0.1–0.4Ex situ; plant analysis; plant samples were assisted by symbiotic bacteria; spiked with 0–35 mM ZnSO4; Saint-Laurent-Le-Minier, FranceGrison et al. [97]
Arabidopsis helleriBrassicaceaa>1% (leaf)Pot experiment; nutrient solution applied as fertilizer; plants exposed to 0–100 μΜ ZnSO4; Blankerode, GermanyZhao et al. [94]
Peer et al. [98]
Arabis paniculataBrassicaceae>2% (leaf)Hydroponics experiment; Hoagland solution applied as fertilizer; plants spiked with 0–2447 μΜ ZnSO4·7H2O; Yunnan Provinence, ChinaReeves et al. [99]; Tang et al. [64]
Corydalis davidiiPapaveraceae>1% (leaf)Field study and hydroponics experiment; total Zn in soil—9847 mg kg−1; plants spiked with 50–400 mg Zn(NO3)2 L−1 in hydroponic experiment; Sub-tropical climate; Magu, ChinaBalafrej et al. [100]; Lin et al. [101]
Haumaniastrum katangenseLamiaceae1.98% (shoot)/20Field study, in situ leaf samples; ZairBalafrej et al. [100]
Noccaea alpestreBrassicaceaa>1% (leaf)Ex situ; plant analysis; total Zn in soil 1900–35,000 mg kg−1; Derbyshire, EnglandShimwell and Laurie [102]
Noccaea brachypetalumBrassicaceaa2% (shoot)Field study, in situ leaf samples; Paris, FranceReeves and Brooks [103]
Noccaea bulbosum SprunerBrassicaceaa1.05% (shoot)Field study, in situ leaf samples; Mediterranean climate; Athens, GreeceReeves and Brooks [103]
Noccaea caerulescensBrassicaceaa2.73% (leaf)/1.6Field study, in situ leaf samples; Temperate climate; BelgiumMartínez et al. [85] Reeves and Brooks [103]
Noccaea calaminareBrassicaceaa3.96% (leaf)Field study, in situ leaf samples; Temperate climate; München, GermanyReeves et al. [99] Reeves and Brooks [103]
Noccaea cepaeifoliumBrassicaceaa2.10% (shoot)Field study; total Zn concentration in soil—17,300 mg kg−1; Mediterranean climate; Northern ItalyReeves and Brooks [103]
Noccaea eburneosaBrassicaceaa1.05% (shoot)Field study, in situ leaf samples; Temperate climate; Zurich, SwitzerlandBalafrej et al. [100] Reeves and Brooks [103]
Noccaea limosellifoliumBrassicaceaa1.10% (shoot)Field study, in situ leaf samples; Paris, FranceReeves and Brooks [103]
Noccaea praecoxBrassicaceaa2.10% (leaf)Field study, in situ leaf samples; Mediterranean climate; BulgariaReeves and Books [103]
Noccaea stenopterumBrassicaceaa1.60% (leaf)Field study, in situ leaf samples; Temperate climate; SpainReeves and Brooks [103]
Sedum alfrediiCrassulaceae0.45–1.9% (shoot)Field survey and greenhouse experiment; plants exposed to 5–320 mg Zn L−1 (as ZnSO4·7H2O); Sub-tropical humid climate; Zhejiang Province, ChinaYang et al. [104]
Pieris divaricataEricaceae>6% (shoot)Pot experiment with contaminated soil; total Zn amount in soil—352 mg kg−1; Sub-tropical climate; Lanping, ChinaBalafrej et al. [100]
Potentilla griffithiiRosaceae>2.2% (leaf)Field survey and hydroponics experiment; Hoagland solution was added as fertilizer; plants spiked with ZnSO4·7H2O at 20–320 mg L−1; Yunnan Province; ChinaReeves et al. [99] Qiu et al. [105]
Sedum plumbizincicolaCrassulaceae0.45–1.9% (shoot)Field survey and greenhouse experiment; plants exposed to 5–320 mg Zn L−1 (as ZnSO4·7H2O); Sub-tropical humid climate; Zhejiang Province, ChinaReeves et al. [99]
Yang et al. [104]
Viola baoshanensisViolaceae>0.9%Field survey and greenhouse experiment; modified Hoagland solution was applied as fertilizer; plants spiked with 0–130 mg ZnSO4·7H2O L−1; Humid sub-tropical climate; Hunan Province, ChinaBalafrej et al. [100];
Wu et al. [106]
Table 6. List of Mn hyperaccumulator species.
Table 6. List of Mn hyperaccumulator species.
Plant SpeciesFamilyAchieved Concentration (mg kg−1)/Biomass Yield (t ha−1)Experimental Conditions—CountryReference
Alyxia poyaensisApocynaceae1.4%Ex situ; plant analysis; Semi-tropical climate; New CaledoniaLosfeld et al. [107]
Beaupreopsis paniculataUrticaceae1.2%Ex situ; plant analysis; Semi-tropical climate; New CaledoniaJaffré [108]
Chengiopanax sciadophylloidesAraliaceae2.4%7-month field trial; exchangeable Mn in soil—931 mg kg−1; Mie, JapanMizuno et al. [109]
Denhamia cunninghamiiSapindaceae2.5%Ex situ; plant analysis; Queensland, AustraliaFernando et al. [110]; Fernando et al. [111]
Denhamia silvestrisSapindaceae1.2%Pot experiment; medium spiked with soluble Mn2+ at 0, 250, 500, and 1000 mg kg−1; Queensland, AustraliaAbubakari et al. [112]
Denhamia fournieriSapindaceae2.0%Ex situ; plant analysis; Semi-tropical climate; New CaledoniaFernando et al. [110];
Celosia argenteaSapindaceae3.3%Field survey and hydroponics experiment; Hoagland solution applied as fertilizer; plants exposed to 2.5–600 mg Mn L−1; Guangxi Province, ChinaJaffré [36]
Liu et al. [113]
Garcinia amplexicaulisClusiaceae1.2%/280Ex situ; plant analysis; Semi-tropical climate; New CaledoniaJaffré [36]
Gossia bamagensisMyrtaceae4.0%Ex situ; leaf analysis; Queensland, AustraliaFernando et al. [110]; Fernando et al. [111]
Gossia bidwilliiMyrtaceae1.9% (leaf), 2.6% (bark)Ex situ; plant analysis; Sub-tropical climate; Brisbane, AustraliaLosfeld et al. [107]; Bidwell et al. [114]
Gossia clusioidesMyrtaceae1.0%Ex situ; plant analysis; Semi-tropical climate; New CaledoniaJaffré [36]
Gossia diversifoliaMyrtaceae1.8%Ex situ; plant analysis; Semi-tropical climate; New CaledoniaLosfeld et al. [107]
Gossia fragrantissimaMyrtaceae3.5%Ex situ; leaf analysis; Queensland, AustraliaFernando et al. [110] Fernando et al. [111]
Gossia gonocladaMyrtaceae1.5%Ex situ; leaf analysis; Queensland, AustraliaFernando et al. [110]; Fernando et al. [111]
Gossia lucidaMyrtaceae1.5%Ex situ; leaf analysis; Queensland, AustraliaFernando et al. [110] Fernando et al. [111]
Gossia sankowskiorumMyrtaceae3.0%Ex situ; leaf analysis; Queensland, AustraliaFernando et al. [110] Fernado et al. [111]
Gossia shepherdiiMyrtaceae1.5%Ex situ; leaf analysis; Queensland, AustraliaFernando et al. [110] Fernado et al. [111]
Phytolacca americamaProteaceae1.1%Ex situ; analysis; Semi-tropical climate; New CaledoniaJaffré [108]
Phytolacca acinosaPhytolaccaceae1.9%Field survey and hydroponics experiment; Hoagland’s nutrient solution was applied; plants were exposed to 0–15,000 μmol Mn L−1 (as MnCl2·4H2O); Hunan Province, ChinaXue et al. [115]
Polygonum pubescensPolygonaceae1.6%Field survey and hydroponics experiment; Hoagland solution applied; Plants exposed to 0–20 mmol Mn L−1 (as MnCl2·4H2O); ChinaDeng et al. [116]
Polygonum lapathifoliumAraliaceae1.4%Field survey and hydroponics experiment; Hoagland solution was applied; plants exposed to 0–16 mmol Mn L−1 (as MnCl2·4H2O); Guangxi Province, ChinaLiu et al. [117]
Schima superbaTyphaceae1.2%Pot experiment; Hoagland solution was added as a fertilizer; S. superba saplings were exposed to 0–200 mmol L−1 MnCl2·4H2O; ChinaYang et al. [118]
Virotia neurophyllaTyphaceae5.5%Ex situ; plant analysis; Semi-tropical climate; New CaledoniaJaffré [108]
Table 7. List of Se hyperaccumulator species.
Table 7. List of Se hyperaccumulator species.
Plant SpeciesFamilyAchieved Concentration (mg kg−1)/Biomass Yield (t ha−1)Experimental Conditions—CountryReference
Asparagus bisculateusAsparagaceae0.88%/12.3Ex situ; plant analysis; CanadaReeves [121]
Moon [122]
Asparagus patersoniiAsparagaceae0.25%/12.3Ex situ; plant analysis; USAReeves [121]
Moon [122]
Asparagus pectinatusAsparagaceae0.51%Ex situ; plant analysis; USAReeves [121]
Asparagus racemosusAsparagaceae1.5%Ex situ; plant analysis; USAReeves [121]
Atriplex canescensAmaranthaceae0.17%/34.75Ex situ; plant and soil analysis; Arid climate; AlgeriaNadaf et al. [123]; Nedjimi [124]
Castilleja chromosaOrobanchaceae0.18%Ex situ; plant analysis; USAReeves [121];
Cardamine violifoliaBrassicaceae0.48%Ex situ; plant analysis; Korea and Yutangba, ChinaMa et al. [125]
Huang et al. [126]
Lecythis ollariaLecythidaceae1.8%Ex situ leaf analysis; VenezuelaReeves [121]
Machaeranthera gabriulsculaAsteraceae0.18%Ex situ leaf analysis; USAReeves [121]
Machaeranthera parryiAsteraceae0.53%Ex situ; leaf analysis; USAReeves [121]
Machaeranthera ramosaAsteraceae0.13%Ex situ; leaf analysis; USAReeves [121]
Machaeranthera venustaAsteraceae0.34%Ex situ; leaf analysis; Colorado, USAReeves [121]
Neptunia amplexicaulisFabaceae0.11%/43.3Pot experiment; soil spiked with 0, 5, and 30 mg Se kg−1 (as Na2SeO4 or CaSeO3); Queensland, AustraliaO’Donohue et al. [127]; Irish et al. [128]
Stanleya bipinnataBrassicaceae0.23%Ex situ; leaf analysis; USAReeves [121]
Stanleya pinnataBrassicaceae0.11%Climate controlled experiment; Hoagland solution applied as fertilizer; plants exposed to 0–80 μΜ N2SeO4; Colorado, USAHarris et al. [129]
Table 8. Economically viable concentrations of precious metals in the harvested biomass that are necessary to make the phytomining process profitable.
Table 8. Economically viable concentrations of precious metals in the harvested biomass that are necessary to make the phytomining process profitable.
ElementNatural Occurrence in the Earth’s Crust (mg kg−1)Economically Viable Plant Metal Concentration (mg kg−1)Economic Proposed Concentration in Biomass (mg kg−1)Reference
Au0.004060.06–0.6Lee et al. [176]
Ag0.0750100010–100Lee et al. [176]
Pd0.01505–150.05–1.5Lee et al. [176]
Pt0.005040.04–0.4Dinh et al. [15]
Rh0.00105–150.05–1.5Dinh et al. [15]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kikis, C.; Thalassinos, G.; Antoniadis, V. Soil Phytomining: Recent Developments—A Review. Soil Syst. 2024, 8, 8. https://doi.org/10.3390/soilsystems8010008

AMA Style

Kikis C, Thalassinos G, Antoniadis V. Soil Phytomining: Recent Developments—A Review. Soil Systems. 2024; 8(1):8. https://doi.org/10.3390/soilsystems8010008

Chicago/Turabian Style

Kikis, Christos, Georgios Thalassinos, and Vasileios Antoniadis. 2024. "Soil Phytomining: Recent Developments—A Review" Soil Systems 8, no. 1: 8. https://doi.org/10.3390/soilsystems8010008

Article Metrics

Back to TopTop