Next Article in Journal
Incremental Capacity Curve Health-Indicator Extraction Based on Gaussian Filter and Improved Relevance Vector Machine for Lithium–Ion Battery Remaining Useful Life Estimation
Next Article in Special Issue
Health Risk Assessment of Children Exposed to the Soil Containing Potentially Toxic Elements: A Case Study from Coal Mining Areas
Previous Article in Journal
Recycling Aluminium AA6061 Chips with Reinforced Boron Carbide (B4C) and Zirconia (ZrO2) Particles via Hot Extrusion
Previous Article in Special Issue
Assessment of the Macro- and Microelement Composition of Fly Ash from 50-Year-Old Ash Dumps in the Middle Urals (Russia)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Soil Pollution and Plant Efficiency Indices for Phytoremediation of Heavy Metal(loid)s: Two-Decade Study (2002–2021)

1
Laboratory of Biotechnology, Institute of Natural Sciences and Mathematics, Ural Federal University, 620002 Ekaterinburg, Russia
2
Department of Environmental Science, School of Engineering and Sciences (SEAS), SRM University-Andhra Pradesh, Amaravati 522240, India
3
Department of Environmental Science and Engineering, Centre of Mining Environment, Indian Institute of Technology-Indian School of Mines (IIT-ISM), Dhanbad 826004, India
4
Department of Experimental Biology and Biotechnology, Ural Federal University, 620002 Ekaterinburg, Russia
*
Authors to whom correspondence should be addressed.
Submission received: 31 May 2022 / Revised: 19 July 2022 / Accepted: 2 August 2022 / Published: 8 August 2022
(This article belongs to the Special Issue Sustainable Metal Waste Management: Biological Approaches)

Abstract

:
This paper reviews research on phytoremediation (2002–2021), particularly for the estimation of plant efficiency and soil pollution indices, examining the extraction of metals from soil and plants growing under both artificial (spiked with specific metal) and natural conditions. From the analysis of >200 published experimental results, it was found that contamination factor and geo-accumulation index as well as translocation and bioconcentration factors are the most important soil pollution and plant efficiency metrices, respectively, which are gaining importance to assess the level of metal pollution and its transfer from soil to plant to find a better metal clean-up strategy for phytoremediation. To access the metal concentration, it was found that the most widely accepted extractants to dissolve and extract the metals from the soil and plant were HNO3 and HClO4 (mainly in 5:1; v/v or 4:1; v/v), which are used both in natural and artificial metal contamination studies. Moreover, plants such as Pteris vittata, Monochoria korsakowi, Lolium perenne, Festuca rubra, Poa pratensis, Ricinus communis, and Siegesbeckia orientalis can act as hyperaccumulators under both natural and artificial experiments and can be directly implemented into the fields without checking their further efficiency in phytoremediation.

1. Introduction

An increase in metal concentration due to anthropogenic activities and natural processes results in water, air, and soil pollution. Heavy metals (HMs) are non-biodegradable, which easily mobilize and accumulate in the environment and thus pose risks to human health and its surroundings. In addition, HMs slowly contaminate varieties of land which can be used for commercial purposes such as agriculture, forestry, nursery, horticulture, etc. However, these metals and metalloids slowly enter the food chain and result in oxidative stress, enzyme disruption, chronic anemia, endocrine disruption, autoimmune and carcinogenic diseases, allergic dermatitis, etc. in humans [1,2,3].
For many years, research has been carried out to decontaminate or reduce the metal contamination by means of physical, chemical and biological methods. However, physical and chemical methods are costly and not environmentally safe, resulting in need for a new and safer technology called “phytoremediation” [1,4,5,6,7]. Phytoremediation is a biological method which found popularity in the late 1990s. It is one of the safest, eco-friendly and cost-effective technologies and helps to control the metal problem without creating adverse effects on the ecosystem [5,8].
From time to time, there has been comprehensive reviews reported every year on the progress of research on phytoremediation such as [2,9,10,11,12,13,14]. Most of the reviews were focused on the search of plants-hyperaccumulators and the mechanism involved in metal transfer from soil to plant. Effective research is ongoing in the world for the use of plant varieties which can help in the remediation of metal-contaminated sites. The research was focused on metal-tolerant, hyperaccumulator, accumulator or excluder plant species on naturally contaminated substrates or in artificially metal spiked soil to remove, stabilize or prevent the leaching of toxic metals in the environment [1,15,16]. Some of the plant efficiency factors that help to assess phytoremediation were also studied; however, they vary from one to another. The extractants used to study the bioavailable and total metal content in soil and plant are also quite variable, which creates strong confusion while choosing the right extractant for the recovery of metal from substrate. Moreover, no research reported all the various soil pollution indices and plant efficiency metrics along with various extractants in a single place to exactly understand the metal pollution level in soil and plant to implement the best methods for its remediation. To enhance the efficiency of phytoremediation, there is a great need to understand and integrate both plant metrics and soil factors. This approach will also give clear ideas to early career researchers, volunteers, and industrialists to study the specific parameters
We put forward the hypothesis that our review will form the scientific basis for the unification of methodological approaches to assess the behavior of metals in the soil–plant system and ensure terminological uniformity. In order to understand the various factors, which play a vital role in phytoremediation, a detailed study of the past two decades (2002–2021) of published articles is conducted to gather a depth of knowledge on soil factors and plant metrics, which must be considered while planning and executing a phytoremediation under artificial laboratory-based experiments and/or for naturally contaminated sites. In the present study, >200 high-quality research works on metal contamination (natural and artificial) were reviewed to understand whether the trend and concept of phytoremediation are going in an appropriate direction for the use of plant metrics and soil factors, the use of single or mixed extractants, sequential and total metal extracts in soil, and the total metal content in plants.

2. Phytoremediation

Phytoremediation is novel, sustainable, cost-effective, promising, solar-driven, eco-friendly technology used for the decontamination of metal-contaminated or enriched sites by removing, destroying or sequestering the hazardous metals using varieties of plant species growing naturally (in situ) or under controlled conditions (ex situ) [17,18,19,20,21,22,23,24,25,26,27,28,29]. The HMs are non-degradable and remain for a long time in the environment. The only possible and most effective method is to sequester them into the plant and use harvest to extract metals from plant parts [9,30,31]. However, the ability to accumulate HMs varies significantly between species and cultivars within a species [2,7,32]. This technology can be applied to both organic and inorganic pollutants present in soil (solid substrate), water (liquid substrate) or the air [31] and can be used for the removal of toxic metals from the biosphere [33,34,35]. Phytoremediation is highly applicable for the low to moderately metal contaminated very large fields where other physical and chemical methods are impracticable for a long period of time [6]. Phytoremediation is often referred as botanical bioremediation or green remediation also [9,36]. Out of the various techniques involved in remediation, phytoremediation was found to be least expensive (US$ 5–40/ton) as reported by [37,38].
The main types of phytoremediation involved in the removal of HMs from the contaminated site are:
  • Phytosequestration: plants which accumulate metals mainly in their roots;
  • Phytoextraction: plants which transfer metals from soil to the aerial part;
  • Phytodegradation: plants which help in the biotransformation of pollutants inside them;
  • Phytovolatilization: plants which help in the volatilization of metals from its leaves; and
  • Rhizoremediation: exudates from plants which help the bacteria for the biodegradation of contaminants.

3. Experimental Studies Using Heavy Metal-Contaminated or Spiked Soil

Hundreds of papers in peer-reviewed journals were studied thoroughly to categorize the study pattern in the field of phytoremediation in metal-contaminated soil between 2002 and 2021. Many studies were carried out on the naturally contaminated or non-spiked metallic-ferrous waste; however, there were also numerous reports from artificially metal-spiked soil. Nevertheless, limited studies were performed together by both naturally and artificially contaminated soil to exactly understand the success of phytoremediation under lab and field conditions, which is reported in the present study.

3.1. Naturally/Non-Spiked Metal-Contaminated Substrate Studies

Non-spiked metal-contaminated soils are those substrates that do not involve any direct addition of metals in salt form from outside. These soils were mainly contaminated because of natural weathering lithology or industrial activities including mining. This kinds of soils are generally deficient in carbon (C) and nutrient content (N, P, K, Ca, Mg, Zn, Mn) as well as in biological agents (enzymes and microbes). Additionally, these soils are characterized by unfavorable physical properties (water-holding capacity, porosity, grain size, bulk density, etc.) with a variable range of multiple metals from moderate to very high concentration. The samples were normally collected from the contaminated sites and used for field, plot, greenhouse, and glasshouse-based studies. Some pot studies were also conducted in the field without providing controlled conditions. On the other hand, some were conducted in greenhouses or glasshouses under controlled conditions by providing optimum requirements using the natural substrates (soil) collected from the metal-contaminated field. It was found that vast research lies in the search and identification of naturally growing, colonized plants, herbs, shrubs, and trees, which are effective in metal accumulation, exclusion and hyperaccumulation and testing them on non-spiked metal-contaminated soil. Apart from these properties, researchers are also in search of high biomass plants which can improve the phytoremediation efficiency. A list of important research carried out for contaminated areas during the past two decades is presented in Table 1.
The major benefits for using naturally growing plants in phytoremediation of metal contaminated area are:
  • High metal tolerance;
  • High biomass;
  • Possibility to tolerate climatic variability of that area;
  • Ability to withstand in harsh conditions;
  • Drought or wet resistance;
  • Ability to withstand variable chemical properties;
  • Ability to tolerate nutrient stress;
  • Possibility to grow in poor physical conditions; and
  • Ability to grow in presence of multiple metals.

3.2. Artificially Metal Contaminated Substrate Studies

Artificial metal contamination experiments are those investigations which were conducted by spiking the substrate with a specific amount of metal from an external source: mainly, the salt of metals. Different metals were being used constantly in fixed concentration so as to check the ability of the plant species for its growth and survival and success in phytoremediation. Most of the experiments include only single metal spiking [70,71,72,73,74,75,76,77,78,79]. However, some researchers had also performed multiple metal spiking experiments to find out new varieties of plants, herbs, shrubs and trees which can decontaminate the multi-metal contaminated sites [69,80,81,82,83,84,85,86]. High-quality research articles published in the past two decades were studied and evaluated, and the majority of the researchers had reported many plants which were able to tolerate and accumulate metals in their root part as well as in the aerial parts in high concentration. Some of the research carried out using the spiking of metals in soil during the last two decade is listed in Table 2.

3.3. Both Naturally and Artificially Metal Contaminated Substrate Studies

It is obvious that lab-based experiments sometimes fail when implemented in a natural situation, and thus, both the studies together could help to understand the plant behavior and suitability and survivability for clean-up of the metal-contaminated sites. However, it was found that naturally growing native plants of metal-contaminated sites are more efficient than non-native plants because of the above stated reasons (see Section 3.1). Some of the experiments carried out during the last decade by the researchers to check the efficiency of the same plant under both natural and artificial contamination, which are listed in Table 3. However, much research is needed for those plants that were tested in artificial conditions because these plants were grown under controlled conditions such as by keeping optimum light, moisture, temperature, and humidity. Field studies are further required to check the efficiency of laboratory-tested plants in naturally metal-contaminated fields where there is great competition and variability in biotic and abiotic conditions (temperature, moisture content, water availability, etc.). Keeping all the aspects in mind, it can be concluded that it is better to search native plant species that have high biomass, drought-resistant, and hyperaccumulation properties in roots with high commercial importance.

4. Quantification of Soil Pollution Indices or Metrices

4.1. Enrichment Factor (EFs)

To search for the most likely source of origin of elements in soil, enrichment factors were calculated for individual elements over the average elemental composition of the tailing [103,104,105,106,107,108,109,110,111,112] (Equation (1)).
Enrichment   factor   ( EF ) = [ ( M ) soil × ( Al ) tailing ] [ ( Al ) soil × ( M ) tailing ]  
where (M)soil and (Al)soil are the concentrations of the studied element and Al in the soil, while (M)tailing and (Al)tailing are the mean concentrations of the element and Al, respectively, in the tailing. Here, aluminum is considered as the reference material because of its wider acceptance as a reference element.
Regional geochemical background values are constant and are recommended by Rubio et al. [113] for the assessment of enrichment of metal in contaminated soil. However, the levels of contamination vary with time and place [105,114], and background values are distinctly different among different soil types. For most HMs of environmental interest, concentrations in soil easily vary over two to three orders of magnitude depending on the parent materials [105]. There are different elements which were used as reference elements to study the contamination pattern. However, iron (Fe) and aluminum (Al) are widely used as the reference elements for geochemical normalization because of the following reasons [105,115]:
  • Both are associated with fine solid surfaces;
  • Its geochemistry is similar to that of many trace metals;
  • Its natural concentration tends to be uniform.
Enrichment factors are distributed under five categorizes [111,116]:
  • EF < 2: depletion to minimal enrichment;
  • EF = 2–5: moderate enrichment;
  • EF = 5–20: significant enrichment;
  • EF = 20–40: very high enrichment;
  • EF > 40: extremely high enrichment.
Due to its unitless dimension and relatively simple formula, it is universally accepted for the assessment of the degree of metal enrichment and source of anthropogenic origin caused by metal mining.

4.2. Contamination Factor (CF)

The contamination factor (CF) is the ratio obtained by dividing the concentration of each metal in the soil by the baseline or background value (concentration in unpolluted soil) (Equation (2)):
Contamination   factor ( CF ) =   [ C ] ( heavy   metal ) [ C ] ( background )  
where [C](heavy metal) is the concentration of each metal in contaminated soil and [C](background) is the concentration of each metal in non-contaminated or baseline or unpolluted soil [117].
The contamination levels may be classified based on their intensities on a scale ranging from 1 to 6 [105,106,107,109,114,115,117,118,119]:
  • CF = 0: None;
  • CF = 1: None to medium;
  • CF = 2: Moderate;
  • CF = 3: Moderate to strong;
  • CF = 4: Strongly polluted;
  • CF = 5: Strong to very strong;
  • CF = 6: Very strong.
The highest number indicates that the metal concentration is 100 times greater than what would be expected in the crust.

4.3. Geoaccumulation Index (Igeo)

Geoaccumulation indexes for the metals were determined using Muller’s [114] expression (Equation (3)):
Geoaccumulation   indexe   ( I geo ) =   log 2 [ ( Metal ) s 1.5   ( Metal ) b ]  
where (Metal)s is the concentration of metals examined in soil samples and (Metal)h is the geochemical background concentration of the metal. Factor 1.5 is the background matrix correction factor due to lithospheric effects [120,121].
The geoaccumulation index consists of seven grades or classes [106,107,108,109,111,114,115,119,122,123,124,125]:
  • Igeo ≤ 0: practically uncontaminated;
  • 0 < Igeo < 1: Uncontaminated to moderately contaminated;
  • 0 < Igeo < 2: Moderately contaminated;
  • 2 < Igeo < 3: Moderately to heavily contaminated;
  • 3 < Igeo < 4: Heavily contaminated;
  • 4 < Igeo < 5: Heavily to extremely contaminated;
  • 5 < Igeo: Extremely contaminated and can be a hundredfold greater than the geochemical background value.

4.4. Pollution Load Index (PLI)

It is the integrated index which combines all the HMs to one index and compares the status of pollution of various sites without considering the grain size, natural geochemical variability and changes of heavy metal/reference element ratios which are based on natural processes. This empirical index provides a simple, comparative means for assessing the level of HM pollution. PLI is calculated for the entire sampling site by taking the nth root of the product of the n CF [105,106,107,109,115,117,118,119,120,126,127] (Equation (4)):
Pollution   load   index   ( PLI ) = ( CF 1 ×   CF 2 ×   CF 3 × ×   CF n ) 1 n  
where CFn is the contamination factors of different elements.
  • PLI < 0: Unpolluted;
  • 0 < PLI ≤ 1: Baseline levels of pollutant present;
  • 1 < PLI ≤ 10: Polluted;
  • 10 < PLI ≤ 100: Highly polluted;
  • PLI > 100: Progressive deterioration of environment.

4.5. Risk Assessment Code (RAC)

Risk assessment code (RAC) is a classification system which includes an assessment of available reactive HMs in soil and calculated as the percentage of metals present in exchangeable and carbonate fraction [128,129]. It is the percentage of metal concentration extracted by acetic acid when used in 0.11 M concentration, which is scaled as [109,130]:
  • RAC < 1: No risk;
  • 1< RAC ≤ 10: Low risk;
  • 11 < RAC ≤ 30: Medium risk;
  • 31 < RAC ≤ 50: Very high risk.
This reactive or available metal gives the indication of potential risk to the ecosystem.

4.6. Potential Ecological Risk Index

The potential ecological risk index, proposed by Hakanson [131], was employed to evaluate the potential risk of HMs in the rhizospheric soil [56,125,132].
Based on an overall consideration of the toxicities and the differences in regional background values of HMs, this evaluation method could eliminate the influence of regional differences and embody the toxicities of HMs and their relative contributions. As a result, a comprehensive reflection of the potential of HMs’ impact on the ecological environment was provided, which made it suitable for evaluating the pollution of HMs in a wide range of area. According to the literature [117,133,134], the toxic factors of Hg, Cd, As, Cu, Pb, Ni, Cr and Zn are 40, 30, 10, 5, 5, 5, 2 and 1, respectively.
According to this method, the potential ecological risk index ( E r i ) of individual heavy metal and the comprehensive potential risk index (RI) of several HMs could be calculated by the following Equation (5):
Potential   ecological   risk   index   ( E r i ) = [ T i ] [ C i ]   [ C oi ]
where Ci and C0i are the measured and background concentrations of element i in soil, respectively, while Ti is the toxic factor of element i [121,125]. The potential ecological risk criteria were scaled as:
  • E r i   < 40: Low;
  • 40 < E r i < 80: Moderate;
  • 80 < E r i < 160: Considerable;
  • 160 < E r i < 320: High;
  • E r i   > 320: Very high.

4.7. Potential Risk Index (RI)

It is calculated as the sum of all the four risk factors for HMs in soils. Hakanson [131] had given the standardized heavy metal toxic factor by the order of level of heavy metal present in the soil (Cd > Pb = Cu > Zn) [56,135,136] (Equation (6)):
Potential   risk   index   ( RI ) = i = 1 m ( E i r )
where E i r is the potential ecological risk index.
  • RI < 150: Low;
  • 150 < RI < 300: Moderate;
  • 300 < RI < 600: Considerable;
  • 600 < RI < 1200: High;
  • E r i   ≧ 1200: Very high.

5. Quantification of Plant Phytoremediation Efficiency Metrices

Different efficiency indices were being used in the past two decades in the field of phytoremediation of metal to study the plant–soil interaction, transport mechanism and accumulation pattern in plants. The different efficiency indices of phytoremediation which can be used by researchers to evaluate the actual status of remediation taking place in the implemented area are available. However, few important indices, i.e., translocation factor, bioconcentration and bioaccumulation factors become more popular to evaluate the efficiency of the plant species for the phytoremediation of metal-contaminated soil/sites. Moreover, from time to time, different names were used for the study of the same factor, which created chaos and misunderstanding in the field of phytoremediation, which is being studied here to resolve such problems for researchers. The present review includes different types of efficiency indices which were frequently used to evaluate the phytoremediation potential of plants in both natural and artificial (spiked) condition. However, translocation and bioconcentration/bioaccumulation factors and extracts used to calculate it are most widely studied by the researchers (between 2002 and 2021) which are exhaustively discussed in Table 4 and Table 5.

5.1. Translocation Factor (TF)

The translocation factor, also termed as accumulation factor, uptake factor, and concentration factor, is an important index for evaluating the transfer potential of metals from soil to plant [3,137,138]. It is regularly used for both naturally colonized/growing and artificially grown/cultured plants under controlled conditions. Plants require metals for their proper growth and development and include different specific carriers and mechanisms for the transport of these metals from soil to plant [139]. Metals present in soil become available to plants in a bioavailable form, which easily become absorbed by the roots and transfer into the shoot through suitable carriers. However, the transportation of metals in plants varies from plants to plant and species to species and further depends on many other factors such as the age of the plant, climatic regime, nature of soil, soil chemistry, ecotype, etc. [45,54,140,141,142]. It seems that the transfer factors derived from different types and ranges of soil metal concentrations are not comparable. Efroymson et al. [143] estimated the uptake of inorganic contaminants in soil to the plants by using a single uptake factor, single-variable regression model and multiple regression models with soil. The present work involves the study of two-decade research papers to study the use of a bioavailable portion for the calculation of transfer factors. Although it is considered that the bioavailable portion of toxic metals is the basis of soil risk assessment of soil contaminants, there are still only a few reports that are using the bioavailable metal concentration for this purpose. In most of the cases, total or pseudo-total metal concentrations were used to calculate the translocation factor.
The translocation factor is the efficiency index of the plant species, which indicates the translocation of metals from the root part to the shoot part [41,45,46,49,53,57,61,63,65,91,141,144,145,146] and can be calculated as follows in Equation (7):
Translocation   factor   ( TF ) = [ C ] ( shoot ) [ C ] ( root )    
where C(shoot) indicates the metal concentration accumulated in the shoot part and C(root) indicates the metal concentration accumulated in the root part.
The same factor was calculated by different researchers and denoted as the “shoot” part using different terms such as “aerial part” [39,51,62,98], stem [54,59,100,147], aboveground tissue part [50,56], and leaves [62,100,148]. Similarly, in the case of ferns, the term “frond” [65,97] and “cap” in case of mushroom [149] were also used to denote the shoot part. It was found that the main aim of all the researchers was to calculate the translocation of metals from root to shoot but not on the term used for the plant part, i.e., shoot, aerial part, tissue aboveground part, leaf, and stem.
The translocation factor can also be calculated in percentage (%) by using the following Equation (8) [2,74]:
Translocation   factor   ( TF ) = [ C ] ( aerial   part   ) [ C ] ( root )   × 100
It is evaluated that a translocation factor > 1 for any plant shows its potential to phytoextract the metal from the root into the shoot, whereas TF < 1 indicates its phytostabilizing property. In low and moderately contaminated soils, the TF values were found to be >1. However, it does not imply the same for the highly metal-contaminated sites.
Transfer factor (Tf): It is the efficiency index of the plant species to accumulate metals from its surrounding substrate (soil/sediment) and can be calculated as follows in Equation (9) [42,50,57,62,97,98,150]:
Transfer   factor   ( Tf ) = [ C ] ( plant ) [ C ] ( substrate )    
where C(plant) is the concentration of metal in the whole plant, and C(substrate) is the concentration of metal in the substrate (soil) in which it is growing.
Dynamic factor of metal translocation (TRdyn): Baltrenaite et al. [151] have introduced a dynamic factor that helps to integrate information about metal concentration in different substrates and provide a comparison between control and treated soil. They are related both to internal (physiological) and external (ecological) factors.
To understand the actual transfer and accumulation of metals, Baltrenaite et al. [151] introduced a few formulas (Equation (10)), which include metal transfer assessment compared to control or non-contaminated soil:
  TR ( dyn ) = TR i ,   treated TR i ,   control   = [ C i ,   v ,   treated ] × [ C i , r ,   control ] [ C i , r ,   treated ] × [ C i , v ,   control ]  
where TRi, treated is the translocation factor of metal i in trees on the treated site; TRi, control is the translocation factor of metal i in trees on the control site; Ci,v, treated is the concentration of metal i in tree vegetative organs on the treated site, in mg per kg dry weight (DW); Ci, r, treated is the concentration of metal i in tree roots on the treated site, in mg per kg DW; Ci, r, control is the concentration of metal i in tree roots on the control site, in mg per kg DW; and Ci, v, control is the concentration of metal i in tree vegetative organs on the control site, in mg per kg DW.
Table 4. List of plant and soil extractant, plant efficiency metrics to evaluate the remediation potential of naturally metal-contaminated site/substrate.
Table 4. List of plant and soil extractant, plant efficiency metrics to evaluate the remediation potential of naturally metal-contaminated site/substrate.
Metal(s)Plant
Digest(s)
Soil
Extractant(s)
Plant Efficiency MetricsReferences
A = Aboveground Part/RootB = Aboveground Part/SubstrateC = Root/SubstrateD = Plant/SubstrateE = Tissue/Substrate
Cd, Cr, Cu, Ni, Pb, ZnUS EPA Method 3051 (1994)US EPA Method 3051 (1994)TF = Aerial/PlantBCF = Shoots SoilBCF = Roots/Soil--[39]
As, Cd, Ni, Pb, Zn* HNO3* HNO3TF = Frond/Root biomass conc.--BF = Plant/Soil-[97]
Pb, Cu, ZnC#USEPA Method 3050C#USEPA Method 3050TF = Shoot/Root- BCF= Plant root/Soil-[41]
Cu, Zn, Cd, PbHNO3quaregiaTF = Shoot/Root-BCF = Root/Soil--[152]
Zn, Cu, Pb, NiConc. HNO3 + HClO4 (5:1)DTPA (C# Total; EDTATF = Shoot/RootBioaccumulation Coefficient = Shoot/DTPA in soilBioaccumulation Coefficient = Root/DTPA in soil--[45]
PbAcid digestion (Undefined)nsTF = Aerial/Root--BCF = Plant/Soil-[98]
Cd, Cu, Pb, ZnAquaregia 70% + 305 H2O2DTPA Total-C#TF = Shoot/Root-BCF= Root/Total soil--[91]
Cr, Zn, Cd, Cu, Ni, PbHNO3:H2O2:HCl, 7:1:1HNO3:H2O2:HCl, 7:1:1; DTPA TF = Total shoot/Total rootBF= Total shoot/Total soil---[46]
Cu, Ni, Fe, MnHNO3 + HClO4HF + HClO4 + HNO3TF = Other plant part/Root---BCF = Plant tissues rooted soil[48]
SrHNO3 then HCl + HNO3+ H2O, 1:1:1HCl+HNO3+H20, 1:1:1TLF = Shoot/RootECS = Enrichment coefficient for shoot---[49]
Pb, Mn10 mL 1 M HClSequential Extraction TF = Shoot/RootEFs = Shoot/SoilEFr = Root/Soil
--[99]
MnC#USEPA 3051, 1995Sequential Extraction (C#); USEPA 3052, 1995TF = Aboveground tissue part/roots--BCF = Whole plant DW/Soil-[50]
CdHNO3:HClO4, 5:1HNO3: HF: HClO4, 5:1:1TF leaf = Leaf/Root;
TF stem = Leaf/Root
BCF = Leaf/Soil
BCF = Stem/Soil
BCF = Root/Soil--[100]
Fe, Mn, Zn, Cd, Cu, Pb, Cr, AsHNO3:HClO4, 4:1Aqua regia+HNO3TF = Aerial/Root---BCF = Plant tissue/Background soil conc. in agri. field[51]
As, Fe, Mn, Cu, Co, ZnHNO3 and HCl, 5:1Aqua regia; DTPA and TEA;
Sequential extraction
-
(because it was not possible to separate completely the roots
of the plants)
BF = Bioaccumulation factor; BF = shoots (total DW)/tailings---[52]
Fe, Zn, CuWithout any
chemical treatment
Without any
chemical treatment
TF = Cap/Stripe--BF = Mushroom/Substrate (soil)-[149]
Fe, Zn, Pb and MnAcid digestion (Undefined)Acid digestion (Undefined)TF = Shoot/Root-BCF = Root/Soil--[53]
Fe, Cu, Pb, Mn, Ni, Zn, Cr, CdHNO3:HClO4, 5:1HNO3:HClO4, 5:1TF = Total shoot/Total rootBAF = Shoot/SubstrateBAF= Root/Substrate- [54]
Cu, Cd, Pb, Cr, Mn, NiHNO3:HClO4, 5:1HNO3:HClO4, 5:1TF = Stem/root---BCF = Plant part/Substrate[54]
Hg, Cd, As, Hg, Pb, Cr, Cu, Zn, NiHCl: HNO3: HClO4, HFHCl: HNO3: HClO4, HFTF = Aboveground tissue part/roots---BCF = Tissue/Rhizospheric soil[56]
As, Fe, Mn, Pb, ZnHNO3 (65%) and H2O2 (30%), 5:1HNO3+H2O2+HFconc.+ HCl+H2O, 9:1:3:2:1TF = Shoot/Root--BF = Plant/Soil-[57]
Cu, Fe, Pb, ZnC#C#TF = Shoot/Root-BCF = Root/Soil--[58]
Ni, Cu, Zn, Cd, PbHNO3:HClO4, 5:1HNO3:HClO4, 5:1TF = Stem/root---BCF = Plant part/Substrate[59]
Cd, Zn, Pb, CuHNO3:HClO4, 4:1HNO3:HCl:HClO4, 1:2:2TF = Shoot/RootBCF = Shoot/SoilBCF = Root/Soil--[101]
Cr, Cu, Ni, Pb, CdHNO3:HClO4, 3:1HNO3:HClO4, 3:1TF = Shoot/RootEF = Shoot/Contaminated soilEF = Root/Contaminated soil--[60]
AsH2SO4/HClO4 Natural plants: HNO3: HClO4, 17:3H2SO4/HClO4TF = Shoot/RootBCF = Aerial biomass concentration/Soil---[61]
AsHNO3 then 30% H2O230%H2O2+9.6 M HClTF = Aerial (leaf or stem)/Root--BCF = Plant/Environment (soil)-[62]
Cd, Co, Cu, Cr, Fe, Mn, Ni, Pb, ZnHNO3 (65%)+HClO4 (70%), 3:2HNO3 (65%)+HClO4 (70%), 3:2TF = Shoot/Root-BF = Root/Soil--[63]
Fe, Pb, As, Cu, Mn, Sb, ZnHNO3+HCl, 2:1Aquaregia (1/3 HNO3+2/3 HCl)TF = Shoot/RootBAF = Shoot/Soil---[64]
Fe, Si, As, Cd, PbHNO3: HClO4, 3:1HNO3: HClO4TF = Frond/Root biomass conc.-BAF = Root/Substrate--[65]
HgHNO3: 30% H2O2Aqua regiaTF = CLeaf/CrootBCF = Cleaf/root/stem/Csoil [79]
HgHNO3:H2SO4, 4:1 (v/v)HCl:HNO3, 3:1 (v/v) BCF = Csoot/Csoil [153]
As, B, Fe, Mn, ZnHNO3: HClO4, 3:1HNO3:HClO4, 5:1TF = Cshoot/CrootBCF = Csoot/Csoil [68]
HgHNO3 and 30% H2O2, 5:265% HNO3TF = Cshoot/CrootBCF = Croot/Csoil [29]
CdHNO3: H2O2, 5:2DTPA extractionTF = CLeaf/CrootBCF = Croot/Csoil [154]
A: Translocation/transfer factor; B: Bioconcentration/Bioaccumulation factor in shoot; C: Bioconcentration factor in root; D: Bioconcentration factor in plant; E: Tissue-specific bioconcentration factor; TF: Translocation factor; Tf: Transfer factor; EF = Enrichment factor; ne: not evaluated; c#: cross referenced; *: ratio not specified.

5.2. Bioconcentration Factor (BCF) or Bioaccumulation Factor (BF)

The bioconcentration factor or bioaccumulation factor is the efficiency index of the plant species to accumulate metals in its harvestable tissue part (root or shoot or leaf) from its surrounding substrate (soil/sediment) [39,100,101,132,145,152,155] and can be calculated as follows Equation (11):
Bioconcentration   ( BCF )   or   Bioaccumulation   factor   ( BF ) = [ C ] ( plant   tissue   or   aerial   part ) [ C ] ( substrate )  
where C(plant tissue) indicates the metal concentration accumulated in the plant tissue (shoot or root or leaf) and C(substrate) indicates the metal concentrations accumulated in the substrate (soil/sediment).
Both the factors, bioconcentration factor and bioaccumulation factor, are rigorously used to calculate the ratio in shoot or in root concentration against substrate concentration [39,45,46,63,64,65]. However, few researchers had used “transfer factor” [155,156,157] or “enrichment factor” (EF) [99] in place of BCF or BF.
The bioconcentration factor or bioaccumulation factor (BF) can also be represented in percent according to the following Equation (12) [158,159]:
  Bioconcentration   ( BCF )   or   Bioaccumulation   factor   ( BF )   % = [ C ] ( plant   tissue   or   aerial   part )   [ C ] ( soil )   × 100  
where BF in %; C(plant tissue) or aerial part is the metal concentration in plant tissue and C(soil) is the metal concentration in soil.
It can be stated after reviewing the papers from the last decade that all the factors are different in terms of their name or notations (BCF, BF and EF); however, their purpose was the same, i.e., to calculate the ratio between the concentration of metal in plant tissue (root or shoot or leaves) and that in substrate (soil or sediment).
Dynamic factor of metal bioaccumulation (BAdyn): Similar to the dynamic factor of metal translocation, another factor called the dynamic factor of bioaccumulation (BAdyn) was calculated by comparing the metal concentration in soil and its accumulation in plants of the contaminated area to that of the metal concentration in soil and its accumulation in the plants of control soil using Equation (13) [151]:
Bioaccumulation ,   BA ( dyn ) = [ C i ,   tree ,   treated ] × [ C i ,   soil ,   control ] [ C i ,   soil ,   treated ] × [ C i ,   tree ,   control ]  
where Ci, tree, treated is the concentration of metal i in tree (the whole plant) ash on the treated site; Ci, soil, treated is the concentration of metal i in the treated soil; Ci, soil, control is the concentration of metal i in the control soil; and Ci, tree, control is the concentration of metal i in the control tree (the whole plant) ash. All values are in milligrams per kg DW.

5.3. Enrichment Factor (EF)

It is the ratio of metal concentration in plant of polluted or contaminated soil to that of the metal concentration in control soil plant [160] and calculated as shown in Equation (14):
Enrichment   factor   ( EF ) = [ C ] ( polluted ) [ C ] ( control )  
where Cpolluted and Ccontrol are the metal concentration in the plant parts (roots, shoots) from the contaminated or polluted sampling soil and control or non-polluted soil.

5.4. Tolerance Index (TI)

The tolerance index (TI), also called as the growth ratio (GR) (reported by [84]), is an important factor to evaluate the efficiency of the plant to grow on metal-contaminated sites in respect to control soil and can be calculated on the basis of biomass [71,74,75,77,92,161] (Equation (15)):
Tolerance   Index   ( TI )   or   Growth   ratio   ( GR ) = [ Biomass ] treated   or   contaminated [ Biomass ] control   or   non contaminated
where [Biomass]treated or contaminated is the biomass of the whole plant in treated or metal-contaminated soil; [Biomass]control or non-contaminated is the biomass of the whole plant in control or non-metal-contaminated soil.
However, different research studies have been carried out in the world to evaluate the efficiency of the plant tolerance in compared to non-contaminated soil (control) in respect to plant length, root length, shoot length, and this was calculated as [71,83,84] (Equation (16)):
Tolerance   Index   ( TI ) = [ Growth   parameters ] treated   or   metal   contaminated [ Growth   parameters ] control   or   non metal   contaminated
where [Growth parameter] can be plant length, root length or shoot length in treated or metal-contaminated soil and in control or non-metal-contaminated soil.

5.5. Metal Extraction Ratio (MER)

It is the ratio of metal accumulation in shoot to that of the soil. Metal extraction ratio (MER) is the efficiency assessment of the phytoextraction capability of plants, which is also known as the phytoextraction ratio and phytoextraction efficiency assessment and can be calculated as shown in Equation (17) [44,68,80,82,101,162,163]:
Metal   extraction   ratio   ( MER ) = [ ( C ) plant × ( M ) plant ] [ ( C ) soil × ( M ) rooted   zone ] × 100  
where (C)plant is the metal concentration in the harvested component of the plant biomass, (M)plant is the mass of the harvestable aboveground biomass produced in one harvest, (C)soil is the metal concentration in the soil volume; and (M)rooted zone is the mass of the soil volume rooted by the plant species.

5.6. Plant Effective Number (PEN)

The plant effective number (PEN) helps to evaluate and compare the ability of different plant species to phytoremediate metal-contaminated soil using hyperaccumulator plants. It is defined as the number of plants needed to extract 1 g of metal when the biomass of shoot and of total plant biomass is considered, as shown in Equation (18) [68,80,82,101,164,165]:
Plant   effective   number   ( PEN ) = 1 [ ( B ) Shoot   or   plant × ( M ) Shoot   or   plant ]  
where (B)shoot or plant is the shoot or whole plant biomass; and (M)shoot or plant is the metal concentration in the shoot or in the whole plant.
Table 5. List of plants and soil extractants as well as plant efficiency metrics to evaluate the remediation potential of artificially spiked contaminated site.
Table 5. List of plants and soil extractants as well as plant efficiency metrics to evaluate the remediation potential of artificially spiked contaminated site.
Metal(s)Plant Digest(s)Plant Efficiency MetricsReferences
A = Aboveground Part/RootB = Aerial Part/SubstrateC = Root/SubstrateD = Plant/SubstrateE = Tissue/Substrate
CrHNO3:HClO4, 3:1Ti = Leaves/Root × 100--BCF = Plant
tissue/Soil
-[87]
CdHNO3:HClO4Ti = Leaves/Root × 100---BCF = Plant tissues at harvest/Substrate[70]
Cd, AsHNO3:HClO4, 3:1TF = Shoot/Root--BCF = Plant/Soil-[90]
Cd, PbHNO3TF = Aerial/Root-BCF = Plant/Culture--[81]
Cd, AsHNO3:HClO4, 3:1TF = Stem/Root--BF = Plant/Soil-[82]
As, Cr, ZnHNO3:HClO4TF = Leaf or Stem/Root--BCF = Whole Plant/Soil-[83]
CdHNO3:HClO4, 3:1TF = Shoot/RootBCF = Shoot/SoilBCF = Root Soil--[71]
CdHNO3:HClO4TF = Shoot/Root--BCF = Plant/Soil-[72]
CrHNO3:HClBAF = Shoot/Root-BAF = Root/Soil--[73]
CdHNO3:H2SO4, 6:2.5Tf = Aerial part/Root---BCF = Harvested plant material/Solution[74]
Cd, CrHNO3TF = Shoot/RootBF = Shoot/Culture---[84]
CrHNO3:30% H2O2Ti = Leaves/Root × 100---BCF = Tissue/Soil[75]
AsNSTF = Shoot/Root--BA = Plant/Solution-[76]
CuHNO3:HClO4, 4:1TF= Shoot/Root--BAF = Plant/Soil-[77]
Cu, Zn, Pb, Cr, CdHNO3:HClO4, 4:1TF = Shoot/Root--BCF = Plant tissue/Soil-[85]
CdHNO3:HClO4, 3:1TF = Shoot/RootBCF = Shoot/SoilBCF = Root/Soil--[92]
HgHCl:HNO3TF = Shoot/RootBCF = Shoot/Soil---[93]
Cu, Zn, Cr0.01 M CaCl2TF = Shoot/Root---BCF = Plant tissue/Soil[86]
CrHNO3 (30%)TF = Shoot/RootSCF = Shoot/SoilRCF = Root/Soil--[78]
HgHNO3 :30% H2O2TF = Aerial part/Root---BCF = Root/Soil[79]
PbHNO3:HClO4, 4:1TF = Shoot/Root--- [95]
HgHNO3: 30% H2O2, 5:2TF = Shoot/Root---BCF = Plant/Soil[29]
CdHNO3:HClO4, 3:1TF = Shoot/Root---BCF = Plant/Soil[166]
A: Translocation/transfer factor; B: Bioconcentration/Bioaccumulation factor in shoot; C: Bioconcentration factor in root; D: Bioconcentration factor in plant; E: Tissue-specific bioconcentration factor; TF: translocation factor; Tf: Transfer factor; BA or BAF: Bioaccumulation factor; BCF: Bioconcentration factor; SCF: Shoot concentration factor; RCF: Root concentration factor; EF = Enrichment factor; Ti: Transportation Index.

5.7. Phytoremediation Factors

The phytoextraction efficiency of plants depends on the concentration of HMs accumulated in the dry aboveground biomass of the plants and the plant yields. The remediation factor (RF) [167,168,169] represents the percentage of an element removed by the plant dry aboveground biomass from the total element content in the soil during one cropping season and was calculated as follows (Equation (19)):
Remediation   factor   ( RF ) = [ ( C ) plant × ( B ) plant ] [ ( C ) soil × ( W ) soil ] × 100 ,
where (C)plant is the metal content in plant dry aboveground biomass (mg kg−1); (B)plant is the plant dry aboveground biomass yield (g); (C)soil is the total metal content in soil (mg kg−1) and (W)soil is the amount of soil in the pot (g).
Total metal uptake: Similarly, the effectiveness of the phytoextraction process (total metal uptake) for the phytoremediation of the metal-contaminated site can also be calculated by multiplying the number of plants growing and the remediation factor (Equation (20)) [170]:
Total   metal   uptake   ( % ) = [ ( C ) plant × ( B ) plant ] [ ( C ) soil × ( W ) soil ]   ( N ) plant × 100 ,
where (C)plant is the metal content in plant dry aboveground biomass (mg kg−1); (B)plant is the plant dry aboveground biomass yield (g); (C)soil is the total metal content in soil (mg kg−1); (W)soil is the amount of soil in the pot (g) and (N)plant is the number of plants.
Dynamic factor of phytoremediation: Another factor called the dynamic factor of phytoremediation (FRi) has been introduced by Baltrenaite et al. [151] to evaluate the phytoremediation capacity of the plants growing in contaminated or treated waste compared to control soil and was calculated as follows (Equation (21)):
  FR i = [ C i ,   tree ] ×   B 1000 × [ C i ,   Soil ] ×   ρ   ×   d   ,
where FRi is the annual metal phytoremediation factor, in kg per ha; Ci, tree is the metal concentration in tree, in mg per kg; B is the annual tree increment, in kg per ha; Ci, soil is the metal concentration in a 40 cm soil layer, in mg per kg; ρ is the soil density, in grams per cubic cm; and d is the soil layer (depth), in cm.

5.8. Phytoextraction Potential (PP)

The phytoextraction potential (PP) is the total amount of HMs extracted per ha of soil in a single phytoextraction cycle [71,168,171]. It is calculated as follows (Equation (22)):
Phytoextraction   potential   ( PP ) = [ C ] plant × [ B ] plant   ,
where [C]plant is the metal content in plant dry aboveground biomass (mg kg−1) and [B]plant is the plant dry aboveground matter biomass yield (t ha−1).

5.9. Removal Efficiency (RE)

It is the efficiency index of the plant to remove metal from a contaminated site and can be calculated as shown in Equation (23):
Removal   efficiency   ( RE   % ) = [ C o ] [ C f ]   [ C o ] × 100   ,
where [Co] is the initial metal concentration and [Cf] is the final metal concentration in the soil after plantation.

6. Measurement of Metal Concentration in Soil and Plant

6.1. Measurement of Bioavailable and Total Metal Concentration in Soil/Substrate

Metals which are available to the plants from the soil/substrate are termed as “bioavailable”. These metals can be extracted by using different extraction methods (using different extractants). Some of the widely used different acid or acid mixtures used by various researchers for total metal analysis in plants under natural and artificial contamination are depicted in Table 4 and Table 5.

6.1.1. Single Extraction Methods

It is well documented that the total concentrations of metals in soil do not act as a good indicator of phytoavailability, or a good tool for potential risk assessment, due to the different and complex distribution patterns of metals among various chemical species or solid phases [56,172]. Several authors have used a single extraction method for the evaluation of the availability of metals in soils [173,174,175], while Tessier et al. [176] and Ure et al. [177] used sequential extraction methods for the evaluation of bioavailability of metals [178]. However, the sequential extraction methods were proposed for sediment, which are quite laborious and time consuming. Among single extraction methods, CaCl2, DTPA, EDTA and CH3COOH were the most widely used extractants [179,180,181,182,183,184,185,186,187,188]. DTPA (0.005 M) is suitable for calcareous soils, as it is buffered at a pH 7.3 and therefore prevents CaCO3 from dissolution and releases occluded metals, especially Cd2+ and Zn2+ [173]. EDTA (0.01 M and 0.05 M) is a very good chelating agent, which can solubilize carbonate-occluded metals from soil [180]. The extraction with water is to simulate the metal distribution equilibrium of metals in soil pore water [185]. It has long been recognized that the soluble, exchangeable and loosely adsorbed metals are quite labile and hence more available for plants [189]. Therefore, in order to assess the environmental risk and the phytoavailability of metals, efforts should be concentrated on the measurements of these available fractions. However, different researchers had used various extractants in different concentrations, which restricts the comparison of data. It is now necessary to agree on a uniform method to obtain concrete and comparable results.

6.1.2. Sequential Extraction Method

In the sequential extraction procedure, metals which are exchangeable, carbonates bound, bound to Fe/Mn oxides, bound to organic matter and sulfides and residual are quantified [50,99,176,190,191]. However, several other researchers have also proposed and modified these methods but were not used consistently [192]. It was found that exchangeable and acid-soluble fractions are the main and more reactive form which is available to living organisms when originating from an anthropogenic source [109]. The main reagents used for sequential extractions are depicted in Table 6.

6.1.3. Total Metal (Digestion) Method

Five mineral acids, namely concentrated hydrochloric acid (HCl), nitric acid (HNO3), sulfuric acid (H2SO4), perchloric acid (HClO4) and hydrofluoric acid (HF), have been very widely used for the estimation of total metals or pseudo-total metals [195]. For the simultaneous extraction of the large number of metals, H2SO4 has one of the notable properties of dissolving silica. Thus, it can be used in conjugation with HNO3, HCl or HClO4 for the total decomposition of silicates [196,197]. Sometimes, HF is also used in conjugation with HNO3, HClO4 [129,185,198,199,200] or HCl [48,56,57,100,184,186,201] for the same purpose. The HNO3 is also used separately [202] or either with HCl [126,203] or HClO4 [63,65,182]. Such methods provide a high degree of metal extractability but do not dissolve silicates completely; they destroy organic matter, dissolve all precipitated and adsorbed metals, and leach out a certain amount of the metal from the silicate lattice. HF is used to break the silica matrix. Aqua regia (HNO3: HCl; 1:3; v/v) and HNO3 are weaker extracting agents than HClO4. Aqua regia is a stronger oxidizing and extracting agent than HNO3 as a result of the presence of nascent chlorine. HNO3, aqua regia and HClO4 have their strongest leaching effect when they are boiling. Especially, HClO4 is a strong leaching, dehydrating and oxidizing agent only when it is hot and in concentrated form. The amount of metal extracted by HClO4 depends on the type of mineral and organic matter content.

6.2. Measurement of Total Metal Concentrations in Plants

HNO3 is often used for metal extraction from plant samples [48,180,204]. A binary acid mixture of HNO3 and HClO4, which is the most widely used extractant (4:1 or 5:1 on in 3:1; v/v), has been mainly used by the researchers for many years for the estimation and determination of metal concentrations in plants [51,54,100,182,185,202,205,206,207]. Sometimes, tertiary acid mixtures of HF, HNO3 and HClO4 [46,61,199,201,208] are also used for the same purpose. However, the use of HF is limited because of the lack of silica estimation in plant parts.
The two decades of research articles reviewed (as stated earlier) were categorized into two sections: (a) papers related to plants collected from a natural condition/habitat growing on natural substrate/soil in greenhouse, pot culture, field, plot, etc. without any artificial contamination and (b) papers related to experiments which include the use of artificial contaminants (spiked) for metal enrichment in soil. It was found that out of the 54 experimental papers, 35 belonging to natural contamination had used HNO3 and HClO4 acids as the main metal extractant from the plant. However, some other extractants were also used such as HCl, H2O2 and HF for the same purpose. When artificially contaminated experimental papers (19) were reviewed, it was found that majority (>50%) of the researchers had used HNO3 and HClO4 as the main plant metal extractant. It suggests that in most of the cases, whether belonging to natural or artificial contamination, these two metal extractants are self-sufficient to digest and extract the majority of the metals present in plant parts.
Similarly, in the case of soil (for a similar number of research papers), the metal extractants used were of wide variety. However, the main extractant used for natural metal-contaminated soil remains the same, i.e., HNO3 and HClO4 (in different ratios). Apart from HNO3 and HClO4, the other most suitable widely used extractant is aqua regia mixture. In case of artificial contaminated soil, metals are spiked with known concentration and were estimated only with HNO3 and HClO4 mixture.

7. Conclusions

The analysis and systematization of the large number of research articles published in the past two decades (2002–2021) allows us to identify the most reliable and representative indices, the use of which will provide a more adequate assessment of the accumulative strategy of plants and contribute to the choice of the more effective metal clean-up phytoremediation technologies. To provide new insight, the present review draws the following conclusions:
(a)
Different soil pollution metrics such as contamination factor, geoaccumulation index, enrichment factor, pollution load index and potential risk indexes provide the opportunity to assess the soil metal pollution; however, the usage of the first two metrics is the most important for both artificial and naturally metal-contaminated sites before implementing phytoremediation strategies.
(b)
Different plant efficiency metrics such as translocation factor, bioconcentration factor, phytoremediation factor, dynamic factor, metal extraction ratio, plant effective number, tolerance index, etc. can provide assessment and practical knowledge about the metal uptake, transfer, and its distribution in plants growing on artificial and natural contaminated sites. Among them, the most suitable are translocation and bioconcentration factors.
(c)
Experiments performed under both natural and artificial contamination suggests some of the hyperaccumulators (Pteris vittata, Monochoria korsakowi, Lolium perenne, Festuca rubra, Poa pratensis, Ricinus communis, Siegesbeckia orientalis) identified in the present review provide further strength to the previous studies reported in the literature.
(d)
The available results in this review of the literature indicate that the translocation and bioconcentration factors were the most important factors which can help to select suitable plants for the decontamination of metal and metalloid-contaminated sites.
(e)
From the depth analysis of published results, it can be concluded that most widely accepted extractants to dissolve and extract the metals from the soil and plant are HNO3 and HClO4 (mainly in 5:1, v/v or 4:1, v/v). It is also important to report that for both natural and artificial contamination, we used the same acid to extract the metal from plant and soil.
Researchers are attempting to identify new and potential hyperaccumulators. However, it was found that most of the artificial experiments are conducted by spiking the substrate/soil material with a single metal that does not correspond to the natural conditions. More research is required to analyze the effect of mixture of metals on plants under artificial condition to provide a better strength of its hyperaccumulation property.

Author Contributions

Conceptualization: A.K., T. and S.K.M. software, A.K.; investigation, A.K. and T.; resources, A.K., T. and D.R.; data curation, A.K., T. and D.R.; writing—original draft preparation, A.K., T. and S.K.M.; writing—review and editing, A.K., T., D.R., S.K.M., M.M. and G.B.; project administration, A.K.; funding acquisition, A.K. All authors have read and agreed to the published version of the manuscript.

Funding

The work was funded by the Ministry of Science and Higher Education of the Russian Federation within the framework of the Ural Federal University Development Program in accordance with the Program of Strategic Academic Leadership “Priority-2030”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

A.K. and S.K.M. also acknowledge the online library support by IIT-ISM. D.R. also acknowledges SRM University-AP for the online library support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bhargava, A.; Carmona, F.F.; Bhargava, M.; Srivastava, S. Approaches for enhanced phytoextraction of heavy metals. J. Environ. Manag. 2012, 105, 103–120. [Google Scholar] [CrossRef] [PubMed]
  2. Ali, H.; Khan, E.; Sjad, M.A. Phytoremediation of heavy metals-Concepts and applications. Chemosphere 2013, 91, 869–881. [Google Scholar] [CrossRef] [PubMed]
  3. Raj, D.; Kumar, A.; Maiti, S.K. Evaluation of toxic metal(loid)s concentration in soils around an open-cast coal mine (Eastern India). Environ. Earth Sci. 2019, 78, 645. [Google Scholar] [CrossRef]
  4. Marques, A.P.; Rangel, A.O.; Castro, P.M.L. Remediation of Heavy Metal Contaminated Soils: Phytoremediation as a Potentially Promising Clean-Up Technology. Crit. Rev. Environ. Sci. Technol. 2009, 39, 622–654. [Google Scholar] [CrossRef]
  5. Marques, A.P.G.C.; Rangel, A.O.S.S.; Castro, P.M.L. Remediation of Heavy Metal Contaminated Soils: An Overview of Site Remediation Techniques. Crit. Rev. Environ. Sci. Technol. 2011, 41, 879–914. [Google Scholar] [CrossRef]
  6. Prasad, M.N.V. A State of the Art Report on Bioremediation, Its Applications to Contaminated Sites in India; Ministry of Environment & Forests, Government of India: New Delhi, India, 2011. [Google Scholar]
  7. Wei, Z.; Van Le, Q.; Peng, W.; Yang, Y.; Yang, H.; Gu, H.; Lam, S.S.; Sonne, C. A review on phytoremediation of contaminants in air, water and soil. J. Hazard. Mater. 2021, 403, 123658. [Google Scholar] [CrossRef] [PubMed]
  8. Nahar, K.; Hoque, S. Phytoremediation to improve eutrophic ecosystem by the floating aquatic macrophyte, water lettuce (Pistia stratiotes L.) at lab scale. Egypt. J. Aquat. Res. 2021, 47, 231–237. [Google Scholar] [CrossRef]
  9. Chaney, R.L.; Malik, M.; Li, Y.M.; Brown, S.L.; Brewer, E.P.; Angle, J.S.; Baker, A.J.M. Phytoremediation of soil metals. Curr. Opin. Biotechnol. 1997, 8, 279–284. [Google Scholar] [CrossRef]
  10. Baker, A.J.M.; McGrath, S.P.; Reeves, R.D.; Smith, J.A.C. Metal Hyperaccumulator Plants: A Review of the Ecology and Physiology of a Biological Resource for Phytoremediation of Metal-Polluted Soils. In Phytoremediation of Contaminated Soil and Water; Terry, N., Banuelos, G., Eds.; CRC: Boca Raton, FL, USA, 2000; pp. 85–107. [Google Scholar] [CrossRef]
  11. Pulford, I.D.; Watson, C. Phytoremediation of heavy metal-contaminated land by trees—A review. Environ. Int. 2003, 29, 529–540. [Google Scholar] [CrossRef]
  12. Yasin, N.A.; Khan, W.U.; Ahmad, S.R.; Ali, A.; Ahmed, S.; Ahmad, A. Effect of Bacillus fortis 162 on growth, oxidative stress tolerance and phytoremediation potential of Catharanthus roseus under chromium stress. Int. J. Agric. Biol. 2018, 20, 1513–1522. [Google Scholar]
  13. Raj, D. Bioaccumulation of mercury, arsenic, cadmium, and lead in plants grown on coal mine soil. Hum. Ecol. Risk Assess. Int. J. 2019, 25, 659–671. [Google Scholar] [CrossRef]
  14. Yan, L.; Van Le, Q.; Sonne, C.; Yang, Y.; Yang, H.; Gu, H.; Ma, N.L.; Lam, S.S.; Peng, W. Phytoremediation of radionuclides in soil, sediments and water. J. Hazard. Mater. 2021, 407, 124771. [Google Scholar] [CrossRef] [PubMed]
  15. Quarshie, S.D.-G.; Xiao, X.; Zhang, L. Enhanced Phytoremediation of Soil Heavy Metal Pollution and Commercial Utilization of Harvested Plant Biomass: A Review. Water Air Soil Pollut. 2021, 232, 475. [Google Scholar] [CrossRef]
  16. Manoj, S.R.; Karthik, C.; Kadirvelu, K.; Arulselvi, P.I.; Shanmugasundaram, T.; Bruno, B.; Rajkumar, M. Understanding the molecular mechanisms for the enhanced phytoremediation of heavy metals through plant growth promoting rhizobacteria: A review. J. Environ. Manag. 2020, 254, 109779. [Google Scholar] [CrossRef] [PubMed]
  17. Suresh, B.; Ravishankar, G.A. Phytoremediation—A Novel and Promising Approach for Environmental Clean-up. Crit. Rev. Biotechnol. 2004, 24, 97–124. [Google Scholar] [CrossRef]
  18. LeDuc, D.L.; Terry, N. Phytoremediation of toxic trace elements in soil and water. J. Ind. Microbiol. Biotechnol. 2005, 32, 514–520. [Google Scholar] [CrossRef]
  19. Chehregani, A.; Malayeri, B.E. Removal of heavy metals by native accumulator plants. Int. J. Agric. Biol. 2007, 9, 462–465. [Google Scholar]
  20. Odjegba, V.J.; Fasidi, I.O. Phytoremediation of heavy metals by Eichhornia crassipes. Environmentalist 2007, 27, 349–355. [Google Scholar] [CrossRef]
  21. Lone, M.I.; He, Z.-L.; Stoffella, P.J.; Yang, X.-E. Phytoremediation of heavy metal polluted soils and water: Progresses and perspectives. J. Zhejiang Univ. Sci. B 2008, 9, 210–220. [Google Scholar] [CrossRef] [Green Version]
  22. Kawahigashi, H. Transgenic plants for phytoremediation of herbicides. Curr. Opin. Biotechnol. 2009, 20, 225–230. [Google Scholar] [CrossRef]
  23. Saier, M.H.; Trevors, J.T. Phytoremediation. Water Air Soil Pollut. 2010, 205, 61–63. [Google Scholar] [CrossRef] [Green Version]
  24. Kalve, S.; Sarangi, B.K.; Pandey, R.A.; Chakrabarti, T. Arsenic and chromium hyperaccumulation by an ecotype of Pteris vittata-prospective for phytoextraction from contaminated water and soil. Curr. Sci. 2011, 100, 888–894. [Google Scholar]
  25. Sarma, H. Metal Hyperaccumulation in Plants: A Review Focusing on Phytoremediation Technology. J. Environ. Sci. Technol. 2011, 4, 118–138. [Google Scholar] [CrossRef] [Green Version]
  26. Singh, A.; Prasad, S.M. Reduction of heavy metal load in food chain: Technology assessment. Rev. Environ. Sci. Bio/Technol. 2011, 10, 199–214. [Google Scholar] [CrossRef]
  27. Vithanage, M.; Dabrowska, B.B.; Mukherjee, A.B.; Sandhi, A.; Bhattacharya, P. Arsenic uptake by plants and possible phytoremediation applications: A brief overview. Environ. Chem. Lett. 2012, 10, 217–224. [Google Scholar] [CrossRef]
  28. Marrugo-Negrete, J.; Marrugo-Madrid, S.; Pinedo-Hernández, J.; Durango-Hernández, J.; Díez, S. Screening of native plant species for phytoremediation potential at a Hg-contaminated mining site. Sci. Total Environ. 2016, 542, 809–816. [Google Scholar] [CrossRef]
  29. Raj, D.; Kumar, A.; Maiti, S.K. Brassica juncea (L.) Czern. (Indian mustard): A putative plant species to facilitate the phytoremediation of mercury contaminated soils. Int. J. Phytoremed. 2020, 22, 733–744. [Google Scholar] [CrossRef]
  30. Prasad, M.N.V. Phytoremediation of metals in the environment for sustainable development. Proc. Indian Natl. Sci. Acad. Part B 2004, 70, 71–98. [Google Scholar]
  31. Shah, V.; Daverey, A. Phytoremediation: A multidisciplinary approach to clean up heavy metal contaminated soil. Environ. Technol. Innov. 2020, 18, 100774. [Google Scholar] [CrossRef]
  32. Mukhopadhyay, S.; Maiti, S.K. Techniques for quantitative evaluation of mine site reclamation success: Case study. In Bio-Geotechnologies for Mine Site Rehabilitation; Elsevier: Amsterdam, The Netherlands, 2018; pp. 415–438. [Google Scholar]
  33. Schroeder, P.; Schwitzguebel, J.P. New cost action launched: Phytotechnologies to promote sustainable land use and improve food safety. J. Soils Sediments 2004, 4, 205. [Google Scholar] [CrossRef]
  34. Ma, Y.; Prasad, M.; Rajkumar, M.; Freitas, H. Plant growth promoting rhizobacteria and endophytes accelerate phytoremediation of metalliferous soils. Biotechnol. Adv. 2011, 29, 248–258. [Google Scholar] [CrossRef] [PubMed]
  35. Mahar, A.; Wang, P.; Ali, A.; Awasthi, M.K.; Lahori, A.H.; Wang, Q.; Li, R.; Zhang, Z. Challenges and opportunities in the phytoremediation of heavy metals contaminated soils: A review. Ecotoxicol. Environ. Saf. 2016, 126, 111–121. [Google Scholar] [CrossRef]
  36. Shackira, A.; Jazeel, K.; Puthur, J.T. Phycoremediation and phytoremediation: Promising tools of green remediation. In Sustainable Environmental Clean-Up; Elsevier: Amsterdam, The Netherlands, 2021; pp. 273–293. [Google Scholar] [CrossRef]
  37. Glass, D.J. U.S. and International Markets for Phytoremediation; Glass Associates: Needham, MA, USA, 1999. [Google Scholar]
  38. Ashraf, S.; Ali, Q.; Zahir, Z.A.; Ashraf, S.; Asghar, H.N. Phytoremediation: Environmentally sustainable way for reclamation of heavy metal polluted soils. Ecotoxicol. Environ. Saf. 2019, 174, 714–727. [Google Scholar] [CrossRef] [PubMed]
  39. Marchiol, L.; Sacco, P.; Assolari, S.; Zerbi, G. Reclamation of Polluted Soil: Phytoremediation Potential of Crop-Related BRASSICA Species. Water Air Soil Pollut. 2004, 158, 345–356. [Google Scholar] [CrossRef]
  40. Maiti, S.K.; Nandhini, S. Bioavailability of Metals in Fly Ash and Their Bioaccumulation in Naturally Occurring Vegetation: A Pilot Scale Study. Environ. Monit. Assess. 2006, 116, 263–273. [Google Scholar] [CrossRef] [PubMed]
  41. Yoon, J.; Cao, X.; Zhou, Q.; Ma, L.Q. Accumulation of Pb, Cu, and Zn in native plants growing on a contaminated Florida site. Sci. Total Environ. 2006, 368, 456–464. [Google Scholar] [CrossRef] [PubMed]
  42. Gupta, A.; Sinha, K.S. Phytoextraction capacity of the plants growing on tannery sludge dumping sites. Bioresour. Technol. 2007, 98, 1788–1794. [Google Scholar] [CrossRef]
  43. Wieshammer, G.; Unterbrunner, R.; García, T.B.; Zivkovic, M.F.; Puschenreiter, M.; Wenzel, W.W. Phytoextraction of Cd and Zn from agricultural soils by Salix ssp. and intercropping of Salix caprea and Arabidopsis halleri. Plant Soil 2007, 298, 255–264. [Google Scholar] [CrossRef]
  44. Zhuang, P.; Yang, Q.W.; Wang, H.B.; Shu, W.S. Phytoextraction of Heavy Metals by Eight Plant Species in the Field. Water Air Soil Pollut. 2007, 184, 235–242. [Google Scholar] [CrossRef]
  45. Maiti, S.K.; Jaiswal, S. Bioaccumulation and translocation of metals in the natural vegetation growing on fly ash lagoons: A field study from Santaldih thermal power plant, West Bengal, India. Environ. Monit. Assess. 2008, 136, 355–370. [Google Scholar] [CrossRef]
  46. Brunetti, G.; Soler-Rovira, P.; Farrag, K.; Senesi, N. Tolerance and accumulation of heavy metals by wild plant species grown in contaminated soils in Apulia region, Southern Italy. Plant Soil 2009, 318, 285–298. [Google Scholar] [CrossRef]
  47. Migeon, A.; Richaud, P.; Guinet, F.; Chalot, M.; Blaudez, D. Metal Accumulation by Woody Species on Contaminated Sites in the North of France. Water Air Soil Pollut. 2009, 204, 89–101. [Google Scholar] [CrossRef]
  48. Wang, S.; Nan, Z.; Liu, X.; Li, Y.; Qin, S.; Ding, H. Accumulation and bioavailability of copper and nickel in wheat plants grown in contaminated soils from the oasis, northwest China. Geoderma 2009, 152, 290–295. [Google Scholar] [CrossRef]
  49. Sasmaz, A.; Sasmaz, M. The phytoremediation potential for strontium of indigenous plants growing in a mining area. Environ. Exp. Bot. 2009, 67, 139–144. [Google Scholar] [CrossRef]
  50. Juárez-Santillán, L.F.; Lucho-Constantino, C.A.; Vázquez-Rodríguez, G.A.; Cerón-Ubilla, N.M.; Beltrán-Hernández, R.I. Manganese accumulation in plants of the mining zone of Hidalgo, Mexico. Bioresour. Technol. 2010, 101, 5836–5841. [Google Scholar] [CrossRef]
  51. Tiwari, K.; Singh, N.; Patel, M.; Tiwari, M.; Rai, U. Metal contamination of soil and translocation in vegetables growing under industrial wastewater irrigated agricultural field of Vadodara, Gujarat, India. Ecotoxicol. Environ. Saf. 2011, 74, 1670–1677. [Google Scholar] [CrossRef]
  52. Kříbek, B.; Mihaljevič, M.; Sracek, O.; Knésl, I.; Ettler, V.; Nyambe, I. The Extent of Arsenic and of Metal Uptake by Aboveground Tissues of Pteris vittata and Cyperus involucratus Growing in Copper- and Cobalt-Rich Tailings of the Zambian Copperbelt. Arch. Environ. Contam. Toxicol. 2011, 61, 228–242. [Google Scholar] [CrossRef]
  53. Nouri, J.; Lorestani, B.; Yousefi, N.; Khorasani, N.; Hasani, A.H.; Seif, F.; Cheraghi, M. Phytoremediation potential of native plants grown in the vicinity of Ahangaran lead–zinc mine (Hamedan, Iran). Environ. Earth Sci. 2011, 62, 639–644. [Google Scholar] [CrossRef]
  54. Pandey, V.C. Invasive species based efficient green technology for phytoremediation of fly ash deposits. J. Geochem. Explor. 2012, 123, 13–18. [Google Scholar] [CrossRef]
  55. Favas, P.J.; Pratas, J.; Prasad, M.N.V. Accumulation of arsenic by aquatic plants in large-scale field conditions: Opportunities for phytoremediation and bioindication. Sci. Total Environ. 2012, 433, 390–397. [Google Scholar] [CrossRef]
  56. Yu, R.; Ji, J.; Yuan, X.; Song, Y.; Wang, C. Accumulation and translocation of heavy metals in the canola (Brassica napus L.)—soil system in Yangtze River Delta, China. Plant Soil 2012, 353, 33–45. [Google Scholar] [CrossRef]
  57. Bech, J.; Duran, P.; Roca, N.; Poma, W.; Sánchez, I.; Barceló, J.; Boluda, R.; Roca-Pérez, L.; Poschenrieder, C. Shoot accumulation of several trace elements in native plant species from contaminated soils in the Peruvian Andes. J. Geochem. Explor. 2012, 113, 106–111. [Google Scholar] [CrossRef]
  58. Bini, C.; Wahsha, M.; Fontana, S.; Maleci, L. Effects of heavy metals on morphological characteristics of Taraxacum officinale Web growing on mine soils in NE Italy. J. Geochem. Explor. 2012, 123, 101–108. [Google Scholar] [CrossRef]
  59. Pandey, V.C. Suitability of Ricinus communis L. cultivation for phytoremediation of fly ash disposal sites. Ecol. Eng. 2013, 57, 336–341. [Google Scholar] [CrossRef]
  60. Kumar, N.; Bauddh, K.; Kumar, S.; Dwivedi, N.; Singh, D.; Barman, S. Accumulation of metals in weed species grown on the soil contaminated with industrial waste and their phytoremediation potential. Ecol. Eng. 2013, 61, 491–495. [Google Scholar] [CrossRef]
  61. Yang, B.; Mengoni, A.; Huang, Y.-L.; He, X.-L.; Li, J.-T.; Liao, B.; Zhou, M.; Shu, W.-S. Exploring the pattern of phenotypic and genetic polymorphism in the arsenic hyperaccumulator Pteris vittata L. (Chinese brake fern). Plant Soil 2013, 373, 471–483. [Google Scholar] [CrossRef]
  62. Liu, C.-W.; Chen, Y.-Y.; Kao, Y.-H.; Maji, S.-K. Bioaccumulation and Translocation of Arsenic in the Ecosystem of the Guandu Wetland, Taiwan. Wetlands 2013, 34, 129–140. [Google Scholar] [CrossRef]
  63. Jasion, M.; Samecka-Cymerman, A.; Kolon, K.; Kempers, A.J. Tanacetum vulgare as a Bioindicator of Trace-Metal Contamination: A Study of a Naturally Colonized Open-Pit Lignite Mine. Arch. Environ. Contam. Toxicol. 2013, 65, 442–448. [Google Scholar] [CrossRef] [Green Version]
  64. Affholder, M.-C.; Prudent, P.; Masotti, V.; Coulomb, B.; Rabier, J.; Nguyen-The, B.; Laffont-Schwob, I. Transfer of metals and metalloids from soil to shoots in wild rosemary (Rosmarinus officinalis L.) growing on a former lead smelter site: Human exposure risk. Sci. Total Environ. 2013, 454–455, 219–229. [Google Scholar] [CrossRef]
  65. Kumari, A.; Pandey, V.C.; Rai, U.N. Feasibility of fern Thelypteris dentata for revegetation of coal fly ash landfills. J. Geochem. Explor. 2013, 128, 147–152. [Google Scholar] [CrossRef]
  66. Fuentes, I.I.; Espadas-Gil, F.; Talavera-May, C.; Fuentes, G.; Santamaría, J.M. Capacity of the aquatic fern (Salvinia minima Baker) to accumulate high concentrations of nickel in its tissues, and its effect on plant physiological processes. Aquat. Toxicol. 2014, 155, 142–150. [Google Scholar] [CrossRef] [PubMed]
  67. Wan, X.; Dong, H.; Feng, L.; Lin, Z.; Luo, Q. Comparison of three sequential extraction procedures for arsenic fractionation in highly polluted sites. Chemosphere 2017, 178, 402–410. [Google Scholar] [CrossRef] [PubMed]
  68. Palanivel, T.M.; Pracejus, B.; Victor, R. Phytoremediation potential of castor (Ricinus communis L.) in the soils of the abandoned copper mine in Northern Oman: Implications for arid regions. Environ. Sci. Pollut. Res. 2020, 27, 17359–17369. [Google Scholar] [CrossRef]
  69. Mazumdar, K.; Das, S. Multi-metal effluent removal by Centella asiatica (L.) Urban: Prospects in phytoremediation. Environ. Technol. Innov. 2021, 22, 101511. [Google Scholar] [CrossRef]
  70. Ghosh, M.; Singh, S. A comparative study of cadmium phytoextraction by accumulator and weed species. Environ. Pollut. 2005, 133, 365–371. [Google Scholar] [CrossRef]
  71. Shi, G.; Cai, Q. Cadmium tolerance and accumulation in eight potential energy crops. Biotechnol. Adv. 2009, 27, 555–561. [Google Scholar] [CrossRef] [PubMed]
  72. Sun, Y.; Zhou, Q.; Wang, L.; Liu, W. Cadmium tolerance and accumulation characteristics of Bidens pilosa L. as a potential Cd-hyperaccumulator. J. Hazard. Mater. 2009, 161, 808–814. [Google Scholar] [CrossRef]
  73. López-Luna, J.; González-Chávez, M.; Esparza-García, F.; Rodríguez-Vázquez, R. Toxicity assessment of soil amended with tannery sludge, trivalent chromium and hexavalent chromium, using wheat, oat and sorghum plants. J. Hazard. Mater. 2009, 163, 829–834. [Google Scholar] [CrossRef]
  74. Zacchini, M.; Pietrini, F.; Mugnozza, G.S.; Iori, V.; Pietrosanti, L.; Massacci, A. Metal Tolerance, Accumulation and Translocation in Poplar and Willow Clones Treated with Cadmium in Hydroponics. Water Air Soil Pollut. 2009, 197, 23–34. [Google Scholar] [CrossRef]
  75. Diwan, H.; Ahmad, A.; Iqbal, M. Uptake-related parameters as indices of phytoremediation potential. Biologia 2010, 65, 1004–1011. [Google Scholar] [CrossRef] [Green Version]
  76. Mirza, N.; Pervez, A.; Mahmood, Q.; Shah, M.M.; Shafqat, M.N. Ecological restoration of arsenic contaminated soil by Arundo donax L. Ecol. Eng. 2011, 37, 1949–1956. [Google Scholar] [CrossRef]
  77. Chen, G.-C.; Liu, Z.; Zhang, J.; Owens, G. Phytoaccumulation of copper in willow seedlings under different hydrological regimes. Ecol. Eng. 2012, 44, 285–289. [Google Scholar] [CrossRef]
  78. Chigbo, C.; Batty, L. Phytoremediation for co-contaminated soils of chromium and benzo[a]pyrene using Zea mays L. Environ. Sci. Pollut. Res. 2013, 21, 3051–3059. [Google Scholar] [CrossRef]
  79. Xun, Y.; Feng, L.; Li, Y.; Dong, H. Mercury accumulation plant Cyrtomium macrophyllum and its potential for phytoremediation of mercury polluted sites. Chemosphere 2017, 189, 161–170. [Google Scholar] [CrossRef] [PubMed]
  80. Kumar, A.; Tripti; Maleva, M.; Bruno, L.B.; Rajkumar, M. Synergistic effect of ACC deaminase producing Pseudomonas sp. TR15a and siderophore producing Bacillus aerophilus TR15c for enhanced growth and copper accumulation in Helianthus annuus L. Chemosphere 2021, 276, 130038. [Google Scholar] [CrossRef]
  81. Zhi-Xin, N.; Li-Na, S.; Tie-Heng, S.; Yu-Shuang, L.; Hong, W. Evaluation of phytoextracting cadmium and lead by sunflower, ricinus, alfalfa and mustard in hydroponic culture. J. Environ. Sci. 2007, 19, 961–967. [Google Scholar]
  82. Sun, Y.; Zhou, Q.; Diao, C. Effects of cadmium and arsenic on growth and metal accumulation of Cd-hyperaccumulator Solanum nigrum L. Bioresour. Technol. 2008, 99, 1103–1110. [Google Scholar] [CrossRef]
  83. Yadav, S.K.; Juwarkar, A.A.; Kumar, G.P.; Thawale, P.R.; Singh, S.K.; Chakrabarti, T. Bioaccumulation and phyto-translocation of arsenic, chromium and zinc by Jatropha curcas L.: Impact of dairy sludge and biofertilizer. Bioresour. Technol. 2009, 100, 4616–4622. [Google Scholar] [CrossRef]
  84. Buendía-González, L.; Orozco-Villafuerte, J.; Cruz-Sosa, F.; Barrera-Díaz, C.; Vernon-Carter, E. Prosopis laevigata a potential chromium (VI) and cadmium (II) hyperaccumulator desert plant. Bioresour. Technol. 2010, 101, 5862–5867. [Google Scholar] [CrossRef]
  85. Chinmayee, M.D.; Mahesh, B.; Pradesh, S.; Mini, I.; Swapna, T.S. The Assessment of Phytoremediation Potential of Invasive Weed Amaranthus spinosus L. Appl. Biochem. Biotechnol. 2012, 167, 1550–1559. [Google Scholar] [CrossRef]
  86. Branzini, A.; González, R.S.; Zubillaga, M. Absorption and translocation of copper, zinc and chromium by Sesbania virgata. J. Environ. Manag. 2012, 102, 50–54. [Google Scholar] [CrossRef] [PubMed]
  87. Ghosh, M.; Singh, S.P. Comparative Uptake and Phytoextraction Study of Soil Induced Chromium by Accumulator and High Biomass Weed Species. Appl. Ecol. Environ. Res. 2005, 3, 67–79. [Google Scholar] [CrossRef]
  88. Moreno, F.N.; Anderson, C.W.N.; Stewart, R.B.; Robinson, B.H.; Nomura, R.; Ghomshei, M.; Meech, J.A. Effect of Thioligands on Plant-Hg Accumulation and Volatilisation from Mercury-contaminated Mine Tailings. Plant Soil 2005, 275, 233–246. [Google Scholar] [CrossRef]
  89. Peco, J.D.; Higueras, P.; Campos, J.A.; Esbrí, J.M.; Moreno, M.M.; Battaglia-Brunet, F.; Sandalio, L.M. Abandoned Mine Lands Reclamation by Plant Remediation Technologies. Sustainability 2021, 13, 6555. [Google Scholar] [CrossRef]
  90. Sun, Y.; Zhou, Q.; Wei, S.; Ren, L. Growth responses of the newly-discovered Cd-hyperaccumulator Rorippa globosa and its accumulation characteristics of Cd and As under joint stress of Cd and As. Front. Environ. Sci. Eng. China 2007, 1, 107–113. [Google Scholar] [CrossRef]
  91. Karami, M.; Afyuni, M.; Rezainejad, Y.; Schulin, R. Heavy metal uptake by wheat from a sewage sludge-amended calcareous soil. Nutr. Cycl. Agroecosyst. 2009, 83, 51–61. [Google Scholar] [CrossRef]
  92. Shi, G.; Liu, C.; Cui, M.; Ma, Y.; Cai, Q. Cadmium Tolerance and Bioaccumulation of 18 Hemp Accessions. Appl. Biochem. Biotechnol. 2012, 168, 163–173. [Google Scholar] [CrossRef] [PubMed]
  93. Pérez-Sanz, A.; Millán, R.; Sierra, M.J.; Alarcón, R.; García, P.; Gil-Díaz, M.; Vazquez, S.; Lobo, M.C. Mercury uptake by Silene vulgaris grown on contaminated spiked soils. J. Environ. Manag. 2012, 95, S233–S237. [Google Scholar] [CrossRef]
  94. Liu, Z.; Wang, L.-A.; Xu, J.; Ding, S.; Feng, X.; Xiao, H. Effects of different concentrations of mercury on accumulation of mercury by five plant species. Ecol. Eng. 2017, 106, 273–278. [Google Scholar] [CrossRef]
  95. Manzoor, M.; Gul, I.; Silvestre, J.; Kallerhoff, J.; Arshad, M. Screening of Indigenous Ornamental Species from Different Plant Families for Pb Accumulation Potential Exposed to Metal Gradient in Spiked Soils. Soil Sediment Contam. Int. J. 2018, 27, 439–453. [Google Scholar] [CrossRef]
  96. Din, B.U.; Rafique, M.; Javed, M.T.; Kamran, M.A.; Mehmood, S.; Khan, M.; Sultan, T.; Munis, M.F.H.; Chaudhary, H.J. Assisted phytoremediation of chromium spiked soils by Sesbania Sesban in association with Bacillus xiamenensis PM14: A biochemical analysis. Plant Physiol. Biochem. 2020, 146, 249–258. [Google Scholar] [CrossRef] [PubMed]
  97. Fayiga, A.O.; Ma, L.Q.; Cao, X.; Rathinasabapathi, B. Effects of heavy metals on growth and arsenic accumulation in the arsenic hyperaccumulator Pteris vittata L. Environ. Pollut. 2004, 132, 289–296. [Google Scholar] [CrossRef]
  98. Kim, I.S.; Hong, Y.H.; Kang, K.H.; Lee, E.J. Effects of lead on bioaccumulation patterns and the ecophysiological response in Monochoria korsakowi. J. Plant Biol. 2008, 51, 284–290. [Google Scholar] [CrossRef]
  99. Padmavathiamma, P.K.; Li, L.Y. Phytoavailability and fractionation of lead and manganese in a contaminated soil after application of three amendments. Bioresour. Technol. 2010, 10, 5667–5676. [Google Scholar] [CrossRef]
  100. Huang, H.; Yu, N.; Wang, L.; Gupta, D.; He, Z.; Wang, K.; Zhu, Z.; Yan, X.; Li, T.; 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] [PubMed]
  101. Zhang, S.; Lin, H.; Deng, L.; Gong, G.; Jia, Y.; Xu, X.; Li, T.; Li, Y.; Chen, H. Cadmium tolerance and accumulation characteristics of Siegesbeckia orientalis L. Ecol. Eng. 2013, 51, 133–139. [Google Scholar] [CrossRef]
  102. Álvarez-Mateos, P.; Alés-Álvarez, F.-J.; García-Martín, J.F. Phytoremediation of highly contaminated mining soils by Jatropha curcas L. and production of catalytic carbons from the generated biomass. J. Environ. Manag. 2019, 231, 886–895. [Google Scholar] [CrossRef]
  103. Abrahim, G.M.S.; Parker, R.J. Assessment of heavy metal enrichment factors and the degree of contamination in marine sediments from Tamaki Estuary, Auckland, New Zealand. Environ. Monit. Assess. 2008, 136, 227–238. [Google Scholar] [CrossRef]
  104. Ho, H.H.; Swennen, R.; Van Damme, A. Distribution and contamination status of heavy metals in estuarine sediments near Cua ong harbor, Ha Long bay, Vietnam. Geol. Belg. 2010, 13, 37–47. [Google Scholar]
  105. Rashed, M. Monitoring of contaminated toxic and heavy metals, from mine tailings through age accumulation, in soil and some wild plants at Southeast Egypt. J. Hazard. Mater. 2010, 178, 739–746. [Google Scholar] [CrossRef]
  106. Sekabira, K.; Origa, H.O.; Basamba, T.A.; Mutumba, G.; Kakudidi, E. Assessment of heavy metal pollution in the urban stream sediments and its tributaries. Int. J. Environ. Sci. Technol. 2010, 7, 435–446. [Google Scholar] [CrossRef] [Green Version]
  107. Adokoh, C.K.; Obodai, E.A.; Essumang, D.K.; Serfor-Armah, Y.; Nyarko, B.J.B.; Asabere-Ameyaw, A. Statistical Evaluation of Environmental Contamination, Distribution and Source Assessment of Heavy Metals (Aluminum, Arsenic, Cadmium, and Mercury) in Some Lagoons and an Estuary Along the Coastal Belt of Ghana. Arch. Environ. Contam. Toxicol. 2011, 61, 389–400. [Google Scholar] [CrossRef] [PubMed]
  108. Bing, H.; Wu, Y.; Sun, Z.; Yao, S. Historical trends of heavy metal contamination and their sources in lacustrine sediment from Xijiu Lake, Taihu Lake Catchment, China. J. Environ. Sci. 2011, 23, 1671–1678. [Google Scholar] [CrossRef]
  109. Dung, T.T.T.; Cappuyns, V.; Swennen, R.; Phung, N.K. From geochemical background determination to pollution assessment of heavy metals in sediments and soils. Rev. Environ. Sci. Bio/Technol. 2013, 12, 335–353. [Google Scholar] [CrossRef]
  110. Yan, G.; Mao, L.; Liu, S.; Mao, Y.; Ye, H.; Huang, T.; Li, F.; Chen, L. Enrichment and sources of trace metals in roadside soils in Shanghai, China: A case study of two urban/rural roads. Sci. Total Environ. 2018, 631–632, 942–950. [Google Scholar] [CrossRef]
  111. Adimalla, N.; Chen, J.; Qian, H. Spatial characteristics of heavy metal contamination and potential human health risk assessment of urban soils: A case study from an urban region of South India. Ecotoxicol. Environ. Saf. 2020, 194, 110406. [Google Scholar] [CrossRef]
  112. Yang, Q.; Yang, Z.; Filippelli, G.M.; Ji, J.; Ji, W.; Liu, X.; Wang, L.; Yu, T.; Wu, T.; Zhuo, X.; et al. Distribution and secondary enrichment of heavy metal elements in karstic soils with high geochemical background in Guangxi, China. Chem. Geol. 2021, 567, 120081. [Google Scholar] [CrossRef]
  113. Rubio, B.; Nombela, M.; Vilas, F. Geochemistry of Major and Trace Elements in Sediments of the Ria de Vigo (NW Spain): An Assessment of Metal Pollution. Mar. Pollut. Bull. 2000, 40, 968–980. [Google Scholar] [CrossRef]
  114. Muller, G. Index of geoaccumulation in sediments of the Rhine River. Geol. J. 1969, 2, 109–118. [Google Scholar]
  115. Bhuiyan, M.A.; Parvez, L.; Islam, M.; Dampare, S.B.; Suzuki, S. Heavy metal pollution of coal mine-affected agricultural soils in the northern part of Bangladesh. J. Hazard. Mater. 2010, 173, 384–392. [Google Scholar] [CrossRef]
  116. Banwart, S.A.; Malmstrom, M.E. Hydrochemical modeling for preliminary assessment of mine water pollution. J. Geochem. Explor. 2001, 74, 73–97. [Google Scholar] [CrossRef]
  117. Obiri-Nyarko, F.; Duah, A.A.; Karikari, A.Y.; Agyekum, W.A.; Manu, E.; Tagoe, R. Assessment of heavy metal contamination in soils at the Kpone landfill site, Ghana: Implication for ecological and health risk assessment. Chemosphere 2021, 282, 131007. [Google Scholar] [CrossRef] [PubMed]
  118. Praveena, S.M.; Radojevic, M.; Abdullah, M.H. The Assessment of mangrove sediment quality in Mengkabong Lagoon: An index analysis approach. Int. J. Environ. Sci. Educ. 2007, 2, 60–68. [Google Scholar]
  119. Jumbe, A.; Nandini, N. Heavy metals analysis and sediment quality values in urban Lakes. Am. J. Environ. Sci. 2009, 5, 678–687. [Google Scholar] [CrossRef]
  120. Zhang, H.; Walker, T.R.; Davis, E.; Ma, G. Ecological risk assessment of metals in small craft harbour sediments in Nova Scotia, Canada. Mar. Pollut. Bull. 2019, 146, 466–475. [Google Scholar] [CrossRef]
  121. Rao, K.; Tang, T.; Zhang, X.; Wang, M.; Liu, J.; Wu, B.; Wang, P.; Ma, Y. Spatial-temporal dynamics, ecological risk assessment, source identification and interactions with internal nutrients release of heavy metals in surface sediments from a large Chinese shallow lake. Chemosphere 2021, 282, 131041. [Google Scholar] [CrossRef]
  122. Christophoridis, C.; Dedepsidis, D.; Fytianos, K. Occurrence and distribution of selected heavy metals in the surface sediments of Thermaikos Gulf, N. Greece. Assessment using pollution indicators. J. Hazard. Mater. 2009, 168, 1082–1091. [Google Scholar] [CrossRef] [PubMed]
  123. Antoniadis, V.; Shaheen, S.M.; Boersch, J.; Frohne, T.; Du Laing, G.; Rinklebe, J. Bioavailability and risk assessment of potentially toxic elements in garden edible vegetables and soils around a highly contaminated former mining area in Germany. J. Environ. Manag. 2017, 186, 192–200. [Google Scholar] [CrossRef]
  124. Li, C.; Sanchez, G.M.; Wu, Z.; Cheng, J.; Zhang, S.; Wang, Q.; Li, F.; Sun, G.; Meentemeyer, R.K. Spatiotemporal patterns and drivers of soil contamination with heavy metals during an intensive urbanization period (1989–2018) in southern China. Environ. Pollut. 2020, 260, 114075. [Google Scholar] [CrossRef]
  125. Liu, B.; Xu, M.; Wang, J.; Wang, Z.; Zhao, L. Ecological risk assessment and heavy metal contamination in the surface sediments of Haizhou Bay, China. Mar. Pollut. Bull. 2021, 163, 111954. [Google Scholar] [CrossRef]
  126. Ma, J.; Chen, Y.; Antoniadis, V.; Wang, K.; Huang, Y.; Tian, H. Assessment of heavy metal(loid)s contamination risk and grain nutritional quality in organic waste-amended soil. J. Hazard. Mater. 2020, 399, 123095. [Google Scholar] [CrossRef]
  127. Wang, F.; Guan, Q.; Tian, J.; Lin, J.; Yang, Y.; Yang, L.; Pan, N. Contamination characteristics, source apportionment, and health risk assessment of heavy metals in agricultural soil in the Hexi Corridor. Catena 2020, 191, 104573. [Google Scholar] [CrossRef]
  128. Perin, G.; Craboledda, L.; Lucchese, M.; Cirillo, R.; Dotta, L.; Zanetta, M.L.; Oro, A.A. Heavy metal speciation in the sediments of Northern Adriatic Sea. A new approach for environmental toxicity determination. In Heavy Metals in the Environment; Lakkas, T.D., Ed.; CEP Consultants: Edinburg, UK, 1985; Volume 2. [Google Scholar]
  129. Peng, Y.; Zhang, S.; Zhong, Q.; Wang, G.; Feng, C.; Xu, X.; Pu, Y.; Guo, X. Removal of heavy metals from abandoned smelter contaminated soil with poly-phosphonic acid: Two-objective optimization based on washing efficiency and risk assessment. Chem. Eng. J. 2021, 421, 129882. [Google Scholar] [CrossRef]
  130. Passos, E.A.; Alves, J.P.H.; Garcia, C.A.B.; Costa, A.C. Metal fractionation in sediments of the Sergipe River, northeast, Brazil. J. Braz. Chem. Soc. 2011, 22, 828–835. [Google Scholar] [CrossRef] [Green Version]
  131. Hakanson, L. An ecological risk index for aquatic pollution control—A sediment ecological approach. Water Res. 1980, 14, 975–1000. [Google Scholar] [CrossRef]
  132. Zhang, S.; Xu, Y.; Wu, M.; Mao, X.; Yao, Y.; Shen, Q.; Zhang, M. Geogenic enrichment of potentially toxic metals in agricultural soils derived from black shale in northwest Zhejiang, China: Pathways to and risks from associated crops. Ecotoxicol. Environ. Saf. 2021, 215, 112102. [Google Scholar] [CrossRef] [PubMed]
  133. Men, C.; Liu, R.; Xu, F.; Wang, Q.; Guo, L.; Shen, Z. Pollution characteristics, risk assessment, and source apportionment of heavy metals in road dust in Beijing, China. Sci. Total Environ. 2018, 612, 138–147. [Google Scholar] [CrossRef]
  134. Kumar, S.B.; Padhi, R.K.; Mohanty, A.K.; Satpathy, K.K. Distribution and ecological- and health-risk assessment of heavy metals in the seawater of the southeast coast of India. Mar. Pollut. Bull. 2020, 161, 111712. [Google Scholar] [CrossRef]
  135. Baruah, S.G.; Ahmed, I.; Das, B.; Ingtipi, B.; Boruah, H.; Gupta, S.K.; Nema, A.K.; Chabukdhara, M. Heavy metal(loid)s contamination and health risk assessment of soil-rice system in rural and peri-urban areas of lower brahmaputra valley, northeast India. Chemosphere 2021, 266, 129150. [Google Scholar] [CrossRef]
  136. Xiang, M.; Li, Y.; Yang, J.; Lei, K.; Li, Y.; Li, F.; Zheng, D.; Fang, X.; Cao, Y. Heavy metal contamination risk assessment and correlation analysis of heavy metal contents in soil and crops. Environ. Pollut. 2021, 278, 116911. [Google Scholar] [CrossRef]
  137. Wang, G.; Su, M.-Y.; Chen, Y.-H.; Lin, F.-F.; Luo, D.; Gao, S.-F. Transfer characteristics of cadmium and lead from soil to the edible parts of six vegetable species in southeastern China. Environ. Pollut. 2006, 144, 127–135. [Google Scholar] [CrossRef] [PubMed]
  138. Chamba, I.; Rosado, D.; Kalinhoff, C.; Thangaswamy, S.; Sánchez-Rodríguez, A.; Gazquez, M.J. Erato polymnioides—A novel Hg hyperaccumulator plant in ecuadorian rainforest acid soils with potential of microbe-associated phytoremediation. Chemosphere 2017, 188, 633–641. [Google Scholar] [CrossRef] [PubMed]
  139. Shanker, A.K.; Cervantes, C.; Loza-Tavera, H.; Avudainayagam, S. Chromium toxicity in plants. Environ. Int. 2005, 31, 739–753. [Google Scholar] [CrossRef] [PubMed]
  140. Alloway, B.J.; Jackson, A.P. The behaviour of heavy metals in sewage sludge-amended soils. Sci. Total Environ. 1991, 100, 151–176. [Google Scholar] [CrossRef]
  141. Cui, Y.-J.; Zhu, Y.-G.; Zhai, R.-H.; Chen, D.-Y.; Huang, Y.-Z.; Qiu, Y.; Liang, J.-Z. Transfer of metals from soil to vegetables in an area near a smelter in Nanning, China. Environ. Int. 2004, 30, 785–791. [Google Scholar] [CrossRef]
  142. Twining, J.R.; Payne, T.E.; Itakura, T. soil-water distribution coefficients and plant transfer factors for 134 Cs, 85 Sr and 65 Zn under field conditions in tropical Australia. J. Environ. Radioact. 2004, 71, 71–87. [Google Scholar] [CrossRef]
  143. Efroymson, R.A.; Sample, B.E.; Suter, G.W., II. Uptake of inorganic chemicals from soil by plant leaves: Regressions of field data. Environ. Toxicol. Chem. 2001, 20, 2561–2571. [Google Scholar] [CrossRef]
  144. Jean, L.; Bordas, F.; Gautier-Moussard, C.; Vernay, P.; Hitmi, A.; Bollinger, J.-C. Effect of citric acid and EDTA on chromium and nickel uptake and translocation by Datura innoxia. Environ. Pollut. 2008, 153, 555–563. [Google Scholar] [CrossRef]
  145. Peng, C.; Tong, H.; Shen, C.; Sun, L.; Yuan, P.; He, M.; Shi, J. Bioavailability and translocation of metal oxide nanoparticles in the soil-rice plant system. Sci. Total Environ. 2020, 713, 136662. [Google Scholar] [CrossRef]
  146. Tripathi, S.; Sharma, P.; Singh, K.; Purchase, D.; Chandra, R. Translocation of heavy metals in medicinally important herbal plants growing on complex organometallic sludge of sugarcane molasses-based distillery waste. Environ. Technol. Innov. 2021, 22, 101434. [Google Scholar] [CrossRef]
  147. Aibibu, N.; Liu, Y.; Zeng, G.; Wang, X.; Chen, B.; Song, H.; Xu, L. Cadmium accumulation in vetiveria zizanioides and its effects on growth, physiological and biochemical characters. Bioresour. Technol. 2010, 101, 6297–6303. [Google Scholar] [CrossRef] [PubMed]
  148. Sousa, A.I.; Caçador, I.; Lillebø, A.I.; Pardal, M.A. Heavy metal accumulation in Halimione portulacoides: Intra- and extra-cellular metal binding sites. Chemosphere 2008, 70, 850–857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Busuioc, G.; Elekes, C.C.; Stihi, C.; Iordache, S.; Ciulei, S.C. The bioaccumulation and translocation of Fe, Zn, and Cu in species of mushrooms from Russula genus. Environ. Sci. Pollut. Res. 2011, 18, 890–896. [Google Scholar] [CrossRef] [PubMed]
  150. Kacholi, D.S.; Sahu, M. Levels and Health Risk Assessment of Heavy Metals in Soil, Water, and Vegetables of Dar es Salaam, Tanzania. J. Chem. 2018, 2018, 1402674. [Google Scholar] [CrossRef]
  151. Baltrėnaitė, E.; Lietuvninkas, A.; Baltrėnas, P. Use of Dynamic Factors to Assess Metal Uptake and Transfer in Plants—Example of Trees. Water Air Soil Pollut. 2012, 223, 4297–4306. [Google Scholar] [CrossRef]
  152. Nirola, R.; Megharaj, M.; Palanisami, T.; Aryal, R.; Venkateswarlu, K.; Naidu, R. Evaluation of metal uptake factors of native trees colonizing an abandoned copper mine–a quest for phytostabilization. J. Sustain. Min. 2015, 14, 115–123. [Google Scholar] [CrossRef] [Green Version]
  153. Qian, X.; Wu, Y.; Zhou, H.; Xu, X.; Xu, Z.; Shang, L.; Qiu, G. Total mercury and methylmercury accumulation in wild plants grown at wastelands composed of mine tailings: Insights into potential candidates for phytoremediation. Environ. Pollut. 2018, 239, 757–767. [Google Scholar] [CrossRef]
  154. He, L.; Zhu, Q.; Wang, Y.; Chen, C.; He, M.; Tan, F. Irrigating digestate to improve cadmium phytoremediation potential of Pennisetum hybridum. Chemosphere 2021, 279, 130592. [Google Scholar] [CrossRef]
  155. Raj, D.; Maiti, S.K. Bioaccumulation of potentially toxic elements in tree and vegetable species with associated health and ecological risks: A case study from a thermal power plant, Chandrapura, India. Rend. Lincei Sci. Fis. Nat. 2019, 30, 649–665. [Google Scholar] [CrossRef]
  156. Moreno-Jiménez, E.; Peñalosa, J.M.; Manzano, R.; Carpena-Ruiz, R.O.; Gamarra, R.; Esteban, E. Heavy metals distribution in soils surrounding an abandoned mine in NW Madrid (Spain) and their transference to wild flora. J. Hazard. Mater. 2009, 162, 854–859. [Google Scholar] [CrossRef]
  157. Hossain, M.B.; Rakib, R.J.; Jolly, Y.; Rahman, M. Metals uptake and translocation in salt marsh macrophytes, Porteresia sp. from Bangladesh coastal area. Sci. Total Environ. 2021, 764, 144637. [Google Scholar] [CrossRef] [PubMed]
  158. Wei, C.-Y.; Chen, T.-B. Arsenic accumulation by two brake ferns growing on an arsenic mine and their potential in phytoremediation. Chemosphere 2006, 63, 1048–1053. [Google Scholar] [CrossRef] [PubMed]
  159. Wilson, B.; Pyatt, F.B. Heavy Metal Bioaccumulation by the Important Food Plant, Olea europaea L.; in an Ancient Metalliferous Polluted Area of Cyprus. Bull. Environ. Contam. Toxicol. 2007, 78, 390–394. [Google Scholar] [CrossRef] [PubMed]
  160. Alagić, S.; Šerbula, S.S.; Tošić, S.B.; Pavlović, A.N.; Petrović, J.V. Bioaccumulation of Arsenic and Cadmium in Birch and Lime from the Bor Region. Arch. Environ. Contam. Toxicol. 2013, 65, 671–682. [Google Scholar] [CrossRef] [PubMed]
  161. Zhang, W.; Pan, X.; Zhao, Q.; Zhao, T. Plant growth, antioxidative enzyme, and cadmium tolerance responses to cadmium stress in Canna orchioides. Hortic. Plant J. 2021, 7, 256–266. [Google Scholar] [CrossRef]
  162. Zhao, F.; Lombi, E.; McGrath, S. Assessing the potential for zinc and cadmium phytoremediation with the hyperaccumulator Thlaspi caerulescens. Plant Soil 2003, 249, 37–43. [Google Scholar] [CrossRef]
  163. Mertens, J.; Luyssaert, S.; Verheyen, K. Use and abuse of trace metal concentrations in plant tissue for biomonitoring and phytoextraction. Environ. Pollut. 2005, 138, 1–4. [Google Scholar] [CrossRef]
  164. García, G.; Faz, Á.; Cunha, M. Performance of Piptatherum miliaceum (Smilo grass) in edaphic Pb and Zn phytoremediation over a short growth period. Int. Biodeterior. Biodegrad. 2004, 54, 245–250. [Google Scholar] [CrossRef]
  165. Demarco, C.F.; Afonso, T.F.; Pieniz, S.; Quadro, M.S.; Camargo, F.; Andreazza, R. Phytoremediation of heavy metals and nutrients by the Sagittaria montevidensis into an anthropogenic contaminated site at Southern of Brazil. Int. J. Phytoremed. 2019, 21, 1145–1152. [Google Scholar] [CrossRef]
  166. Jan, A.U.; Hadi, F.; Shah, A.; Ditta, A.; Nawaz, M.A.; Tariq, M. Plant growth regulators and EDTA improve phytoremediation potential and antioxidant response of Dysphania ambrosioides (L.) Mosyakin & Clemants in a Cd-spiked soil. Environ. Sci. Pollut. Res. 2021, 28, 43417–43430. [Google Scholar] [CrossRef]
  167. Vysloužilová, M.; Tlustoš, P.; Száková, J. Cadmium and zinc phytoextraction potential of seven clones of Salix spp. planted on heavy metal contaminated soils. Plant Soil Environ. 2011, 49, 542–547. [Google Scholar] [CrossRef] [Green Version]
  168. Neugschwandtner, R.W.; Tlustoš, P.; Komárek, M.; Száková, J. Phytoextraction of Pb and Cd from a contaminated agricultural soil using different EDTA application regimes: Laboratory versus field scale measures of efficiency. Geoderma 2008, 144, 446–454. [Google Scholar] [CrossRef]
  169. Usman, A.R.A.; Mohamed, H.M. Effect of microbial inoculation and EDTA on the uptake and translocation of heavy metal by corn and sunflower. Chemosphere 2009, 76, 893–899. [Google Scholar] [CrossRef] [PubMed]
  170. Titah, H.S.; Abdullah, S.R.S.; Mushrifah, I.; Anuar, N.; Basri, H.; Mukhlisin, M. Effect of applying rhizobacteria and fertilizer on the growth of Ludwigia octovalvis for arsenic uptake and accumulation in phytoremediation. Ecol. Eng. 2013, 58, 303–313. [Google Scholar] [CrossRef] [Green Version]
  171. Kos, B.; Grčman, H.; Leštan, D. Phytoextraction of lead, zinc and cadmium from soil by selected plants. Plant Soil Environ. 2003, 49, 548–553. [Google Scholar] [CrossRef] [Green Version]
  172. Chen, B.; Shan, X.Q.; Qian, J. Bioavailability index for quantitative evaluation of plant availability of extractable soil trace elements. Plant Soil. 1996, 186, 275–283. [Google Scholar] [CrossRef]
  173. Lindsay, W.L.; Norvell, W.A. Development of DTPA tests for Fe, Mn, Cu and Zn. Soil Sci. Soc. Am. 1978, 42, 421–428. [Google Scholar] [CrossRef]
  174. Leschber, R.; Davis, R.D.; Lhermite, P. Chemical Methods for Assessing Bio-Available Metals in Sludges and Soil; Elsevier Applied Science Publishers: London, UK, 1985. [Google Scholar]
  175. Tan, D.; Long, J.; Li, B.; Ding, D.; Du, H.; Lei, M. Fraction and mobility of antimony and arsenic in three polluted soils: A comparison of single extraction and sequential extraction. Chemosphere 2018, 213, 533–540. [Google Scholar] [CrossRef]
  176. Tessier, A.; Campbell, P.G.C.; Bisson, M. Sequential extraction procedure for the speciation of particulate trace metals. Anal. Chem. 1979, 51, 844–851. [Google Scholar] [CrossRef]
  177. Ure, A.M.; Quevauviller, P.; Muntau, H.; Griepink, B. Speciation of Heavy Metals in Soils and Sediments. An Account of the Improvement and Harmonization of Extraction Techniques Undertaken Under the Auspices of the BCR of the Commission of the European Communities. Int. J. Environ. Anal. Chem. 1993, 51, 135–151. [Google Scholar] [CrossRef]
  178. Nannoni, F.; Protano, G. Chemical and biological methods to evaluate the availability of heavy metals in soils of the Siena urban area (Italy). Sci. Total Environ. 2016, 568, 1–10. [Google Scholar] [CrossRef] [PubMed]
  179. Houba, V.J.G.; Lexmond, T.M.; Novozamsky, I.; van der Lee, J.J. State of the art and future developments in soil analysis for bioavailability assessment. Sci. Total Environ. 1996, 178, 21–28. [Google Scholar] [CrossRef]
  180. McGrath, D. Application of single and sequential extraction procedures to polluted and unpolluted soils. Sci. Total Environ. 1996, 178, 37–44. [Google Scholar] [CrossRef]
  181. Novozamsky, I.; Lexmond, T.M.; Houba, V.J.G. A Single Extraction Procedure of Soil for Evaluation of Uptake of Some Heavy Metals by Plants. Int. J. Environ. Anal. Chem. 1993, 51, 47–58. [Google Scholar] [CrossRef]
  182. Yang, B.; Shu, W.; Ye, Z.; Lan, C.; Wong, M. Growth and metal accumulation in vetiver and two Sesbenia species on lead/zinc mine tailings. Chemosphere 2003, 52, 1593–1600. [Google Scholar] [CrossRef] [Green Version]
  183. Freitas, H.; Prasad, M.N.V.; Pratas, J. Plant community tolerant to trace elements growing on the degraded soils of Sao Domingos mine in the south east of Portugal: Enviromental implications. Environ. Int. 2004, 30, 65–72. [Google Scholar] [CrossRef] [Green Version]
  184. Vega, F.; Covelo, E.; Andrade, M.L.; Marcet, P. Relationships between heavy metals content and soil properties in minesoils. Anal. Chim. Acta 2004, 524, 141–150. [Google Scholar] [CrossRef]
  185. Wang, X.-P.; Shan, X.-Q.; Zhang, S.-Z.; Wen, B. A model for evaluation of the phytoavailability of trace elements to vegetables under the field conditions. Chemosphere 2004, 55, 811–822. [Google Scholar] [CrossRef]
  186. Das, M.; Maiti, S.K. Metal Mine Waste and Phytoremediation: A Review. Asian J. Water Environ. Pollut. 2007, 4, 169–176. [Google Scholar]
  187. Liang, J.; Yang, Z.; Tang, L.; Zeng, G.; Yu, M.; Li, X.; Wu, H.; Qian, Y.; Li, X.; Luo, Y. Changes in heavy metal mobility and availability from contaminated wetland soil remediated with combined biochar-compost. Chemosphere 2017, 181, 281–288. [Google Scholar] [CrossRef]
  188. Yao, Y.; Sun, Q.; Wang, C.; Wang, P.-F.; Ding, S.-M. Evaluation of organic amendment on the effect of cadmium bioavailability in contaminated soils using the DGT technique and traditional methods. Environ. Sci. Pollut. Res. 2017, 24, 7959–7968. [Google Scholar] [CrossRef] [PubMed]
  189. Kabata-Pendias, A. Behavioural properties of trace metals in soils. Appl. Geochem. 1993, 2, 3–9. [Google Scholar] [CrossRef]
  190. Sungur, A.; Soylak, M.; Yılmaz, S.; Ozcan, H. Heavy metal mobility and potential availability in animal manure: Using a sequential extraction procedure. J. Mater. Cycles Waste Manag. 2016, 18, 563–572. [Google Scholar] [CrossRef]
  191. Hasan, M.; Kausar, D.; Akhter, G.; Shah, M.H. Evaluation of the mobility and pollution index of selected essential/toxic metals in paddy soil by sequential extraction method. Ecotoxicol. Environ. Saf. 2018, 147, 283–291. [Google Scholar] [CrossRef] [PubMed]
  192. Rao, C.R.M.; Sahuquillo, A.; Sanchez, J.F.L. A Review of the Different Methods Applied in Environmental Geochemistry for Single and Sequential Extraction of Trace Elements in Soils and Related Materials. Water Air Soil Pollut. 2008, 189, 291–333. [Google Scholar] [CrossRef]
  193. Lin, R.; Stuckman, M.; Howard, B.H.; Bank, T.L.; Roth, E.A.; Macala, M.K.; Lopano, C.; Soong, Y.; Granite, E.J. Application of sequential extraction and hydrothermal treatment for characterization and enrichment of rare earth elements from coal fly ash. Fuel 2018, 232, 124–133. [Google Scholar] [CrossRef]
  194. Oral, E.V. Comparison of Modified Tessier and Revised BCR Sequential Extraction Procedures for the Fractionation of Heavy Metals in Malachite Ore Samples Using ICP-OES. At. Spectrosc. 2019, 40, 122–126. [Google Scholar] [CrossRef]
  195. Alloway, B.J. Heavy Metals in Soils; Blackie: Glasgow, UK; London, UK, 1990; p. 339. [Google Scholar]
  196. Gąsiorek, M.; Kowalska, J.; Mazurek, R.; Pająk, M. Comprehensive assessment of heavy metal pollution in topsoil of historical urban park on an example of the Planty Park in Krakow (Poland). Chemosphere 2017, 179, 148–158. [Google Scholar] [CrossRef]
  197. Marrugo-Negrete, J.; Pinedo-Hernández, J.; Díez, S. Assessment of heavy metal pollution, spatial distribution and origin in agricultural soils along the Sinú River Basin, Colombia. Environ. Res. 2017, 154, 380–388. [Google Scholar] [CrossRef]
  198. Shu, W.; Ye, Z.; Lan, C.; Zhang, Z.; Wong, M. Acidification of lead/zinc mine tailings and its effect on heavy metal mobility. Environ. Int. 2001, 26, 389–394. [Google Scholar] [CrossRef]
  199. Shan, X.-Q.; Wang, Z.; Wang, W.; Zhang, S.; Wen, B. Labile rhizosphere soil solution fraction for prediction of bioavailability of heavy metals and rare earth elements to plants. Anal. Bioanal. Chem. 2003, 375, 400–407. [Google Scholar] [CrossRef] [PubMed]
  200. Cui, X.; Geng, Y.; Sun, R.; Xie, M.; Feng, X.; Li, X.; Cui, Z. Distribution, speciation and ecological risk assessment of heavy metals in Jinan Iron & Steel Group soils from China. J. Clean. Prod. 2021, 295, 126504. [Google Scholar] [CrossRef]
  201. Maiz, I.; Arambarri, I.; Garcia, R.; Millán, E. Evaluation of heavy metal availability in polluted soils by two sequential extraction procedures using factor analysis. Environ. Pollut. 2000, 110, 3–9. [Google Scholar] [CrossRef]
  202. Raj, D.; Kumar, A.; Maiti, S.K. Mercury remediation potential of Brassica juncea (L.) Czern. for clean-up of flyash contaminated sites. Chemosphere 2020, 248, 125857. [Google Scholar] [CrossRef]
  203. Tereshatov, E.E.; Burns, J.D.; Haar, A.L.V.; Schultz, S.J.; McIntosh, L.A.; Tabacaru, G.C.; McCann, L.A.; Avila, G.; Hannaman, A.; Lofton, K.N.; et al. Separation, speciation, and mechanism of astatine and bismuth extraction from nitric acid into 1-octanol and methyl anthranilate. Sep. Purif. Technol. 2022, 282, 120088. [Google Scholar] [CrossRef]
  204. Sharma, P.; Tripathi, S.; Chandra, R. Phytoremediation potential of heavy metal accumulator plants for waste management in the pulp and paper industry. Heliyon 2020, 6, e04559. [Google Scholar] [CrossRef]
  205. Brofas, G.; Michopoulos, P.; Alifragis, D. Sewage Sludge as an Amendment for Calcareous Bauxite Mine Spoils Reclamation. J. Environ. Qual. 2000, 29, 811–816. [Google Scholar] [CrossRef]
  206. Rosselli, W.; Keller, C.; Boschi, K. Phytoextraction capacity of trees growing on a metal contaminated soil. Plant Soil 2003, 256, 265–272. [Google Scholar] [CrossRef]
  207. Das, M.; Maiti, S.K. Growth of Cymbopogon citratus and Vetiveria zizanioides on cu mine tailings amended with chicken manure and manure-soil mixtures: A pot scale study. Int. J. Phytoremed. 2009, 11, 651–663. [Google Scholar] [CrossRef]
  208. Marrugo-Negrete, J.; Durango-Hernández, J.; Pinedo-Hernández, J.; Olivero-Verbel, J.; Díez, S. Phytoremediation of mercury-contaminated soils by Jatropha curcas. Chemosphere 2015, 127, 58–63. [Google Scholar] [CrossRef]
Table 1. List of metal(loid)s accumulating, excluding and hyperaccumulating plant species studied for the substrate remediation under non-spiked soil condition.
Table 1. List of metal(loid)s accumulating, excluding and hyperaccumulating plant species studied for the substrate remediation under non-spiked soil condition.
Experiment(s)Metal(s) StudiedConcentration of Metal(s)Type of
Accumulation
Plant(s)Remediation
Status and Capacity
Type of
Remediation
References
PCd, Cr, Cu, Ni, Pb, ZnNSNSBrassica juncea, B. napus, B. carinata, R. sativaBrassica species demonstrated a similar performance for Cd and Zn, whereas for other elements, the bioconcentration factor was very lowNS[39]
FFe, Mn, Zn, Cu, Pb, Ni, CdNSExcluderBlumea lacera,
Avera aspera,
Borrhevia repens,
Cynodon dactylon
Cynodon dactylon can be used for remediation; all studied plants are useful in in situ biostabilizationBS[40]
FPb, Cu, ZnNSAccumulatorPhyla nodiflora,
Gentiana pennelliana,
Cynodon dactylon,
Bidens alba var. radiata,
Rubus fruticosus, and 29 others
Phyla nodiflora was the most efficient in accumulating Cu and Zn in its shoots, while G. pennelliana is a potential phytostabilizer (Pb, Cu and Zn)PS, PR[41]
FCu, Zn, Mn, Cr, PdNSAccumulatorCalotropis. procera,
Sida acuta,
Ricinus communis,
Cassia fistula
Sida acuta and Cassia fistula are suitable for decontamination of metalsPE[42]
PCd, ZnNSAccumulatorSalix caprea,
S. fragilis,
S. × smithiana,
S. × dasyclados,
Arabidopsis halleri
Salix × smithiana suitable for PEPE[43]
FZn, Pb, CdNSAccumulatorVertiveria zizanioides,
Dianthus chinensis,
Rumex K-1 (Rumex upatientia × R. timschmicus),
R. crispus,
R. acetosa,
Viola baoshanensis,
Sedum alfredii
EDTA applied.
Phytoextraction rates of V. baoshanensis and S. alfredii for Cd and Zn were 0.88% and 1.15%, respectively.
Rumex crispus is best for Cd and Zn phytoextraction
PE[44]
FZn, Cu, Pb, NiNSNSTypha latifolia,
Fimbristylis dichotoma,
Amaranthus defluxes,
Saccharum spontaneum,
Cynodon dactylon
Typha latifolia and S. spontaneum can be used for bioremediation:
rhizofiltration for Zn, Cu, Pb, Ni and
phytoextration for Mn
RF
PE
[45]
FCr, Zn, Cd, Cu, Ni, PbNSExcluderStipa austroitalica,
Dasypyrum villosum
Carduus pycnocephalus,
Silybum marianum,
Sinapis arvensis
Carduus pycnocephalus, S. marianum and S. arvensis act as metal excluderPS[46]
FCd, Zn, PbNSAccumulatorPopulus tremala X P. tremuloides,
Acer campestre,
Acer pseudoplatanus,
Alnus glutinosa,
Betula pendula,
Fraxinus excelsior,
Prunus avium,
Quercus robur,
Salix caprea, and 8 others
Salicaceae family can accumulate 950 mg Zn kg−1 DWPE[47]
FCu, Ni, Fe, MnNSNSTriticum aestivumCitric acid and NH4OAc are the good indicators of Cu availabilityNS[48]
FSrNSNSEuphorbia macrocleda,
Verbascum cheirunthifolium,
Astragalus gummifer
Shoots of these plants are good bioaccumulators.
Astragalus gummifer can be useful either for the cleaning of Sr from contaminated soils or for phytoremediation
PR[49]
PLMnNSExcluder,
Accumulator
Equisetum hyemate,
Telypteris kunthii,
Cnidoscolus multilobus,
Platanus mexicana,
Solanum diversifolium,
Asclepius curassavia,
Pluchea sympitifolia
Re-vegetate and stabilize Mn tailings: E. hyemate and T. kunthii are excluders (E) whereas C. multilobus, P. mexicana, S. diversifolium, A. curassqavia, and P. sympitifolia are accumulators (A)PS[50]
FFe, Mn, Zn, Cd, Cu, Pb, Cr, AsNSAccumulatorSpinacia oleracea,
Raphanus sativus,
Lycopersicon esculentum,
Lepidium sativum,
Peucedanum graveolens,
Coriandrum sativum,
Capsicum annum,
Brassica oleracea var capitata,
Solanum melongena,
Hibiscus esculentus
Spinacia oleracea, L. esculentum, C. annum, B. oleracea var capitata, R. sativus can accumulate As, Cd, Cr, Pb, NiNS[51]
FAs, Fe, Mn, Cu, Co, ZnNSHyperaccumulatorPteris vittataPteris vittata is a hyperaccumulator of As and suitable for phytoremediationPR[52]
FFe, Zn, Pb and MnNSAccumulatorScrophularia scoparia,
Centaurea virgata,
Echinophora platyloba,
Scariola orientalis,
Centaurea virgata,
Cirsium congestum
and 6 other species
Scrophularia scoparia was the most suitable for the phytostabilization of Pb,
C. virgata, E. platyloba and S. orientalis had the potential for the phytostabilization of Zn and C. virgata and C. congestum were the most efficient in the phytostabilization of Mn
PS[53]
FFe, Cu, Pb, Mn, Ni, Zn, Cr, CdNSExcluderSaccharum munjaSaccharum munja is suitable for metal rehabilitation and stabilizationPS[54]
FCu, Cd, Pb, Cr, Mn, NiNSAccumulatorIpomea carneaNSPR[54]
FAsNSAccumulatorRanunculus trichophyllus,
Ranunculus peltatus subsp. saniculifolius, Lemna minor,
Azolla caroliniana,
Juncus effusus,
Callitriche lusitanica,
Callitriche brutia
Callitriche stagnalis,
Fontinalis antipyretica
The highest concentration of arsenic was found in: C. lusitanica (2346 mg kg−1 DW), C. brutia (523 mg kg−1 DW), C. stagnalis (354 mg kg−1 DW), L. minor (430 mg kg−1 DW), A. caroliniana (397 mg kg−1 DW), R. trichophyllus (354 mg kg−1 DW), and F. antipyretica (346 mg kg−1 DW).
Callitriche family plants are accumulator.
PF[55]
PLHg, Cd, As, Hg, Pb, Cr,
Cu, Zn, Ni
NSNSBrassica napusPE capacity is limitedPE[56]
FAs, Fe, Mn, Pb, ZnNSAccumulator, HyperaccumulatorPlantago orbignyana,
Lepidium bipinnatifidum,
Sonchus oleraceus,
Baccharis atifolia
Lepidium bipinnatifidum is a phytoextractor, P. orbignyana is a Pb and Zn hyperaccumulatorPE[57]
FCu, Fe, Pb, ZnNSAccumulatorTaraxacum officinaleAccumulates both in root and shootPE[58]
FNi, Cu, Zn, Cd, PbNSAccumulatorRicinnus communisRicinnus communis is suitable for phytostabilization and revegetationPS[59]
FCr, Cu, Ni, Pb, CdNSAccumulatorCalotropis procera,
Croton bonplandianum,
Cyperus rotundus,
Datura stramonium,
Euphorbia hirta,
Parthenium hysterophorus,
Phyllanthus amarus,
Sida cordifolia,
Solanum nigrum,
Solanum xanthocarpum,
Spinacia oleracea,
Tridax procumbens
EF > 1 for all the weed suggests its use for the phytoremediation and restoration of land contaminated toxic metalsPR[60]
F(GL),
P
AsNSHyperaccumulatorPteris vittataNSPE[61]
FAsNSAccumulatorKandelia obovataNSNS[62]
FCd, Co, Cu, Cr, Fe, Mn, Ni, Pb, ZnNSAccumulatorTanacetum vulgareTanacetum vulgare accumulates Cr and Fe in roots. Bioindicator of Cd, Mn, and Zn [63]
FFe, Pb, As, Cu, Mn, Sb, ZnNSNSRosmarinus officinalisHealth risks related to ingestion of contaminated rosemary may be limited ADI for As, Cu, Pb and SbPS[64]
PFe, Si, As, Cd, PbNSAccumulatorThelypteris dentataPhytoremediation/revegetationPR/RV[65]
HPNi16,600 mg Ni kg−1 DWHyperaccumulatorSalvinia minimaThe plant species can be used to hyperaccumulate Ni in their tissuesPE[66]
FAs, Pb6017 mg As kg−1 DW, 499.5 mg Pb kg−1 DWHyperaccumulator (As),
Accumulator (Pb)
Pteris vittata(As)
Pteris vittata (Pb)
The plant species can tolerate high metals accumulation in the mining areas and can be used for phytoextraction purposes.
Can hyperaccumulate As and accumulate a significant concentration of Pb and Cd
PE[67]
As, Pb, Cd1032 mg As kg−1 DW, 2350 mg Pb kg−1 DW, 1201 mg Cd kg−1 DWHyperaccumulator (As), Accumulator (Pb, Cd)Viola principis
PAs, B, Fe, Mn, Zn Ricinus communisPlant species have better efficiency to remove copper from the soilPS[68]
HPCd, Cr, Cu, Pb, Ni, ZnNSNSCentella asiaticaLower the toxic metals in the effluent in the range of: Cd (14–54%), Cr (2–43%), Cu (18–81%), Pb (35–90%), Ni (13–59%), and Zn (20–81%)NS[69]
F: Field; P: Pot; GH: Greenhouse; CT: Culture tubes; PL: Plot; PC: Pot culture; F(PC): Field (Pot culture); GL: Glasshouse; HP: Hydroponic; F(GL): Field glasshouse; CP: Culture plates; DW: Dry weight; NC: Natural contamination; AC: Artificial contamination; E: Excluder; A: Accumulator; H: Hyperaccumulator; RF: Rhizofiltration; PE: Phytoextraction; PS: Phytostabilization; BS: Biostabilization; PR: Phytoremediation; PF: Phytofiltration; BR: Bioremediation; RV: Revegetation; NS: not specified.
Table 2. List of competent plant species used to study the remediation potential under spiked metal contaminated condition.
Table 2. List of competent plant species used to study the remediation potential under spiked metal contaminated condition.
Experiment(s)Metal(loid)s StudiedSpiked/Metal AdditionConcentration of Metal(loid)sType of
Accumulation
Plant(s) StudiedRemediation Status and CapacityType of RemediationReferences
PCCrK2Cr2O70, 5, 10, 20, 50, 100 and 200 mg kg−1ABrassica juncea
Brassica campestris
Ipomea carnea
Phragmytes karka
Lantana camara
Cassia tora
Cr extraction was I. carnea > D. innoxia > C. tora > P. karka > B. juncea > L. camara > B. campestris.
B. juncea and B. campestris are accumulators. Ipomea carnea and P. karka are useful in phytoextraction
PE[87]
PHgCoarse and fine silicansABrassica junceaThiosulfate induced plant Hg accumulationPE[88]
PCCdCd(NO3)210, 20, 50, 100 and 200 mg kg−1NSIpomea carnea,
Brassica juncea,
Dhatira innoxia,
Phragmytes karka
Brassica juncea accumulated maximum Cd; I. carnea followed by D. innoxia and P. karka were the most suitable species for the phytoextraction of cadmium from soilPE[70,89]
F(P)Cd, Cu, Zn, Pb, and AsCdCl2·2.5H2O;
Na2HAsO4·7H2O
Cd|: 10, 25, 50 mg kg−1;
As: 50, 250 mg kg−1
HRorippa globosaAs a Cd hyperaccumulator,
Large amounts of Cd could accumulate in the shoots, and the TF and BF values are >1.0
PR[90]
HPCd, PbCdCl2·2.5H2O, Pb(NO3)2H2OCd: 0, 5, 10, 20 mg L−1 ; Pb: 50, 100, 200 mg L−1AHelianthus annuus,Brassica juncea,
Medicago sativa,
Ricinus communis
Helianthus annuus showed better ability of accumulation than the othersPE[81]
F(PC)Cd, AsCdCl2·2.5H2O, Na2HAsO4·7H2O10, 25, and 50 mg kg−1HSolanum nigrumNo reduction in plant height and shoot dry biomass was noted when the plants were grown at Cd concentration of 625 mg kg−1PR[82]
PLCd, Cu, Pb, ZnSludge5 mg Cd kg−1, 385 mg Cu kg−1, 180 mg Pb kg−1, 1885 mg Zn kg−1NSTriticum aestivumNSNS[91]
PC (GH)As, Cr, ZnNa2HAsO4·7H2O,
K2Cr2O7 and ZnSO4·7H2O
As, Cr: 0, 25, 50, 100, 250 and 500 mg kg−1; Zn: 0, 500, 1000, 2000, 3000 and 4000 mg kg−1NSJatropha curcasApplication of organic amendment stabilizes the As, Cr and Zn and reduced their uptake in plant tissues.
Jatropha curcas has the potential for the recovery and reclamation of a metalloid and metal-contaminated soil system
PS[83]
P(GH)CdCdCl2·2.5H2O0, 50, 100 or 200 mg kg−1ACannabis sativa,
Linum usitatissimums,
Arachis hypogaea
and 5 others
Good energy crop on contaminated soil, all are good Cd tolerant.
C. sativa, L. usitatissimums,
A. hypogaea are accuulator.
PS[71]
PCdCdCl2·2.5H2O0, 8, 16, 24, 32, 50, and 100 mg kg−1HBidens pillosaHas the potential for the phytoremediation of HMs-contaminated soilsPR[72]
CPCr, Cr3, Cr6+Tannery sludge;
Cr3+ as CrCl3·6H2O,
Cr6+ as K2Cr2O7
Cr: 0–8000 mg kg−1; Cr3+: 0–2000 and Cr6+: 0–500 mg kg−1.NSAvena sativa,
Sorgham bicolor,
S. sadanense,
Triticum aestivum
Cr accumulated mostly in the roots but not in the shootsNS[73]
PCd0 or 50 μM of CdSO4NSAPoplar and Willow ClonesSalicaceae clones are suitable phytoremediation. Willows had a far greater ability to tolerate Cd than poplarsPR[74]
CT
(50 days)
Cd, CrCdCl2·2H2O,
K2Cr2O7
Cd(II): 0, 0.3, 0.65, 1.3, 2.2 mM;
Cr (VI): 0, 0.5, 1, 2, 3.4 mM
HProsopis laevigataBioaccumulation factor greater than 100 for Cd and 24 for Cr.
Hyperaccumulator of Cd(II) and Cr(VI)
PR[84]
PCrK2Cr2O7 HArabidopsis thalianaA significant increase in Cr accumulation (0.64–4.19 mg g−1)
DW, stem; and 0.77–1.1 mg Cr g−1
PE[75]
PAsAs2O30, 50, 100, 300, 600 and 1000 g L−1NSArundo donaxArundo donax can be used for the remediation of arsenic-contaminated soilsPE,
PV
[76]
PCuCuSO4·5H2O100–400 mg kg−1ASalix jiangsuensis,
S. babylonica
Suitable for use in the phytoremediation of Cu-contaminated wetlandsPS[77]
PCCu, Zn, Pb, Cr, CdCuSO4 (60, 200, 300 mM), ZnSO4 (200, 280, 350 mM), Pb(NO3)2(30, 90, 180 mM), CrCl3 (60, 180, 360 mM) and CdSO4 (20, 25 and 30 mM)NSAAmaranthus spinosusAmaranthus spinosus is a potential agent for accumulation and translocationNS[85]
P(GH)CdCdCl2·2.5H2O0, 25 mg kg−1NSCannabis sativaPhytoremediation of Cd-contaminated soilPR[92]
P(GH)HgHgCl20.6, 5.5 mg kg−1NSSilene vulgarisSilene vulgari is good for phytostabilizationPS[93]
GHCu, Zn, Cr Low: 60 mg Cu kg−1, 125 mg Zn kg−1, 50 mg Cr kg−1.
High: 700 mg Cu kg−1, 1050 mg Zn kg−1, 116 mg Cr
kg−1
NSSesbania virgataSesbania virgata tolerated and stabilized high concentrations of Cu, Zn and CrPS[86]
GHCrK2Cr2O70, 50 and 100 mg kg−1NSZea maysGood for Cr and Co-contaminated soilPR[78]
HPHgHgCl22, 10, 50, 100, 200, 500, and 800 g L−1 Opuntia stricta,
Aloe vera, Setcreasea purpurea, Chlorophytum comosum,
Oxalis corniculata
-PE[94]
PPbPb(NO3)2500, 1000, 1500 to 2000 mg Pb kg−1 of soilHPelargonium hortorum, Mesembryanthemum criniflorumOut of 21 plant species, P. hortorum and M. criniflorum, accumulated more than 1000 mg Pb kg−1PE[95]
PCrNS50, 100, 200 mg Cr kg−1 Sesbania sesbanGood for Cd remediationNS[96]
PHgHgCl210, 50, 100, 500, 1000 mg Hg kg−1NSBrassica junceaThe plant species showed good uptake of Hg up to the concentration level at 500 mg kg−1PR[29]
F: Field; P: Pot; GH: Greenhouse; CT: Culture tubes; PL: Plot; PC: Pot culture; F(PC): Field (Pot culture); GL: Glasshouse; HP: Hydroponic; F(GL): Field glasshouse; CP: Culture plates; F(P): Field (Pot); DW: Dry weight; P(GH): Pot(Green house); NS: not specified; E: Excluder; A: Accumulator; H: Hyperaccumulator; PE: Phytoextraction; PS: Phytostabilization; PR: Phytoremediation; PV: Phytovolatilization.
Table 3. List of competent plant species used to study the remediation potential in both naturally and artificial metal-contaminated site/substrate.
Table 3. List of competent plant species used to study the remediation potential in both naturally and artificial metal-contaminated site/substrate.
Experiment(s)Metal(loid)sSpiked/Metal AdditionMetal(loid)s ConcentrationType of
Accumulation
Plant(s) StudiedRemediation Status and CapacityType of
Remediation
References
GHAs, Cd, Ni, PbNitrate salts of Pb, Cd, Ni, Zn50 or 200 mg kg−1 of each metalHPteris vittata-PE[97]
PL, P, HPPbPb(NO3) 20, 24.1, 48.3, 96.5, 241.3, 482.6 μM Pb
(0, 0.5, 1, 2, 5, 10 mg Pb2+)
AMonochoria korsakowiUseful in phytoremediation of co-contaminated metalPE[98]
PPb, Mn, Cu, Zn 80 mg Cu kg−1,
146 mg Pb kg−1,
408 mg Mn kg−1,
148 mg Zn kg−1
-Lolium perenne,
Festuca rubra,
Poa pratensis
Phosphate addition increased exchangeable Mn fraction by 35%, and a combined application of amendments lowered the exchangeable Mn fraction by 50%PS[99]
PCd-2.8 mg P kg−1ARicinus communis (21 varieties)Phytoremediation of DDTs/Cd co-contaminated soilsPR[100]
F, P, PLCd, Zn, PbCdCl2.2.5H2O0, 5, 10, 30, 60, 90, 120 and 150 mg Cd kg−1HSiegesbeckia orientalisSiegesbeckia orientalis is a Cd-accumulator with hyperaccumulating ability.
In shoot: F: 117.48, P: 192.92, PL: 77.10 mg kg−1, P: 0–150 mg kg−1, PL: Cd, Pb, Zn and Cu concentrations were 28.44, 517.53, 1814.15 and 57.04 mg kg−1, resp.
PR,
PE
[101]
GHHg[Hg(NO3)]21, 5, and 10 µg Hg g−1AJatropha curcasJatropha curcas is accumulatorA[102]
PHgHgCl25, 10, 20, 50, 100, 200, 500 and 1000 mg Hg kg−1HCyrtomium macrophyllumPromising plant species to remediate Hg from the soilns[79]
HPHgHgCl22, 10, 50, 100, 200, 500, and 800 μg Hg L−1 Setcreasea purpurea, Chlorophytum comosum and Oxalis corniculata, Aloe vera, Opuntia strictaOxalis corniculata: most suitable for transferring Hg at concentrations of less than 500 μg L−1ns[94]
PHgHgCl210, 50, 100, 500, and 1000 mg Hg kg−1 soilHBrassica junceaPlant showed better efficiency up to the concentration level of 500 mg Hg kg−1 soilPE[29]
F: Field; P: Pot; GH: Greenhouse; PL: Plot; GL/GH: Glasshouse; HP: Hydroponic; A: Accumulator; H: Hyperaccumulator; PE: Phytoextraction; PS: Phytostabilization; PR: Phytoremediation; ns: not specified.
Table 6. Tessier’s scheme for the sequential extraction of metals [67,176,193,194].
Table 6. Tessier’s scheme for the sequential extraction of metals [67,176,193,194].
Metal FractionsReagents Used
ExchangeableMgCl2 1 mol L−1 at pH 7.0
CarbonaticCH3COONa 1 mol L−1/HOAc at pH 5.0
Oxides Fe/MnNH2OH.HCl 0.04 mol L−1 in 25% HOAc
Organic matter and sulfidicH2O2 8.8 mol L−1/HNO3 + NH4OAC 0.8 mol L−1
ResidualHF/HClO4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kumar, A.; Tripti; Raj, D.; Maiti, S.K.; Maleva, M.; Borisova, G. Soil Pollution and Plant Efficiency Indices for Phytoremediation of Heavy Metal(loid)s: Two-Decade Study (2002–2021). Metals 2022, 12, 1330. https://0-doi-org.brum.beds.ac.uk/10.3390/met12081330

AMA Style

Kumar A, Tripti, Raj D, Maiti SK, Maleva M, Borisova G. Soil Pollution and Plant Efficiency Indices for Phytoremediation of Heavy Metal(loid)s: Two-Decade Study (2002–2021). Metals. 2022; 12(8):1330. https://0-doi-org.brum.beds.ac.uk/10.3390/met12081330

Chicago/Turabian Style

Kumar, Adarsh, Tripti, Deep Raj, Subodh Kumar Maiti, Maria Maleva, and Galina Borisova. 2022. "Soil Pollution and Plant Efficiency Indices for Phytoremediation of Heavy Metal(loid)s: Two-Decade Study (2002–2021)" Metals 12, no. 8: 1330. https://0-doi-org.brum.beds.ac.uk/10.3390/met12081330

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop