Next Article in Journal
Study on Prevention and Control Measures of Water–Sand Inrushes Based on Adaptive Adjustment of Mined Thickness
Next Article in Special Issue
Research Progress and Trend of Agricultural Non-Point Source Pollution from Non-Irrigated Farming Based on Bibliometrics
Previous Article in Journal
Assessment of ERA5-Land Data in Medium-Term Drinking Water Demand Modelling with Deep Learning
Previous Article in Special Issue
Abundance of Microplastics in Two Venus Clams (Meretrix lyrata and Paratapes undulatus) from Estuaries in Central Vietnam
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phytoremediation Prospects for Restoration of Contamination in the Natural Ecosystems

1
Faculty of Forestry, Sher e Kashmir University of Agricultural Sciences and Technology of Kashmir, Benhama Ganderbal, Srinagar 191201, Jammu and Kashmir, India
2
College of Agricultural Engineering and Technology, Sher e Kashmir University of Agricultural Sciences and Technology of Kashmir, Shalimar, Srinagar 190025, Jammu and Kashmir, India
3
Jindal Global Business School (JGBS), O P Jindal Global University, Sonipat 131001, Haryana, India
4
School of Hydrology and Water Resources, Nanjing University of Information Science and Technology, Nanjing 210044, China
*
Authors to whom correspondence should be addressed.
Water 2023, 15(8), 1498; https://doi.org/10.3390/w15081498
Submission received: 6 March 2023 / Revised: 2 April 2023 / Accepted: 8 April 2023 / Published: 11 April 2023

Abstract

:
Toxic substances have a deleterious effect on biological systems if accrued in ecosystems beyond their acceptable limit. A natural ecosystem can become contaminated due to the excessive release of toxic substances by various anthropogenic and natural activities, which necessitates rehabilitation of the environmental contamination. Phytoremediation is an eco-friendly and cost-efficient method of biotechnological mitigation for the remediation of polluted ecosystems and revegetation of contaminated sites. The information provided in this review was collected by utilizing various sources of research information, such as ResearchGate, Google Scholar, the Scopus database and other relevant resources. In this review paper, we discuss (i) various organic and inorganic contaminants; (ii) sources of contamination and their adverse effects on terrestrial and aquatic life; (iii) approaches to the phytoremediation process, including phytoextraction, rhizoremediation, phytostabilization, phytovolatilization, rhizofiltration, phytodegradation, phytodesalination and phytohydraulics, and their underlying mechanisms; (iv) the functions of various microbes and plant enzymes in the biodegradation process and their potential applications; and (v) advantages and limitations of the phytoremediation technique. The reported research aimed to adequately appraise the efficacy of the phytoremediation treatment and facilitate a thorough understanding of specific contaminants and their underlying biodegradation pathways. Detailed procedures and information regarding characteristics of ideal plants, sources of heavy metal contamination, rhizodegradation techniques, suitable species and removal of these contaminants are put forward for further application. Scientists, planners and policymakers should focus on evaluating possible risk-free alternative techniques to restore polluted soil, air and water bodies by involving local inhabitants and concerned stakeholders.

1. Introduction

Various ecosystems are becoming polluted progressively due to environmental contamination and the release of industrial effluents. The nature of contamination is either organic—for instance, pesticides, polycyclic aromatic hydrocarbons (PAHs), herbicides and polychlorinated biphenyls (PCBs)—or inorganic (toxic metals). Potentially toxic elements are recognized as being deleterious to the health of living beings if present in adequately high concentration levels in the biosphere. The majority of them are heavy metals (Co, Cr, Cd, Fe, Hg, Cu, Mo, Pb, Ni, Mn, V, Sn and Zn), though some are metalloids (Sb, As) and non-metals (Se) [1]. These organic and inorganic toxic substances either originate naturally or result from human influence activities, including from physical weathering of bedrock and minerals [2], application of phosphate fertilizers in agriculture [3], use of insecticides [4], wastewater sludge [5], azo dye manufacture [2], waste incinerators [6], timber treatment [7], metal quarrying and smelting operations [8] and fossil fuel combustion and electroplating operations [9], which indicates an urgent need for treatment of contaminated ecosystems. Physicochemical treatments used to reduce the toxicity of contaminants in polluted areas [10,11], such as adsorption, chemical precipitation, membrane filtration, oxidation with hydrogen peroxide, coagulation–flocculation [10], the photocatalytic degradation process, ion exchange, oxidation with ozone and electrochemical and flotation approaches, are very expensive and cannot be utilized as the sole methods to attain environmental quality standards [12]. Therefore, these conventional physicochemical processes for contaminated ecosystems have led to the emergence of a new bioremediation technology called phytoremediation, a plant-based approach [13].
A meta-analysis of the literature discloses that diverse studies have been carried out in recent years on rehabilitating contaminated ecosystems by employing phytoremediation approaches, but comprehension of and deep insights into the mechanisms of different phyto-techniques have not been properly documented yet. Therefore, this review study assayed the sources and impacts of contamination along with the different methods of phyto-technology and their mechanisms and respective suitable plant species. The review also shows the potential application of the phytoremediation technique, its advantages and its limitations. Overall, this study investigates the possibility of achieving ecosystem remediation objectives by using fast-growing plants that also provide recreational benefits, as well as wildlife habitat.

2. Phytoremediation

The term “phytoremediation” is composed of two roots: (i) the word “phyto” originates from Greek and connotes plant species; and (ii) the word “remedium” is Latin and means ability to treat or re-establish [14]. The technique implies using plant species to extract and eradicate elemental contaminants or reduce their biological availability in soil and water bodies and, hence, improve degraded environments [15]. In other words, it is an approach to lower the amount of contamination in polluted sites to environmentally tolerable levels by using green plants [16]. Phytoremediation consists of two systems: the root system hosting the microbial population and the plant itself, which transforms the toxic compounds to further non-toxic substances. It is an eco-friendly method for the rehabilitation of polluted ecosystems using plants and is relatively less expensive, as it is carried out in situ and exploits solar energy [17]. Plant species selection is a vital factor for successful phytoremediation and plants must possess the characteristics shown [18] (Figure 1). Fast-growing woody plant species, such as Populus and Salix, are easy to regenerate with extensive roots and have high evapotranspiration rates that stabilize pollutants; hence, they can flourish on soils with contaminated conditions. Such plants also have higher rates of biomass production and can resist a broad range of potentially harmful elements in their root systems [19]. The most important plant species may commonly be situated in colonizing contaminated places [20], such as plants emerging in mining sites. However, the most appropriate selection of plants for phyto-technology predominantly relies on the contaminant characteristics and their concentration levels [21], depth and volatilization rates [22,23], the age of the location [24] and biological degradability [25]. Furthermore, a clear knowledge of the water-use relations for a selected plant is also essential for the success of the phytoremediation approach.

3. Mechanism of the Phytoremediation Technique

On the basis of the type of contaminant, the media in which the treatment take place and the scope of application, the phyto-technique can be categorized into specific mechanisms (Table 1): phytoaccumulation/phytoextraction, rhizodegradation, phytostabilization, rhizofiltration, phytoevaporation/phytovolatilization, phytodesalination, phytodegradation/phytotransformation and phytohydraulics [26]. Once a plant takes up contaminants, it implements different detoxification mechanisms, such as accumulation, translocation, degradation, excretion or volatilization. The process of degradation involves transformation, fragmentation and/or deposition of the pollutant inside plant tissues [27].

3.1. Phytoextraction

Metal contamination impairs biological systems through direct consumption of metal-infected food, drinking contaminated water or the food chain as it can predispose living beings to serious health problems [29]. This type of contaminant does not undergo a process of biological degradation and, hence, necessitates physically eradication or transformation into non-toxic substances [30]. Phytoextraction/phytoaccumulation is the technique through which plants take up pollutants through their roots with subsequent translocation and accretion in their aboveground parts, which is generally followed by harvesting and final disposal of the biomass of the plant (Figure 2). Phytoextraction is also called phytosequestration and phytoabsorption. The process applies to metals (Cd, Ag, Cu, Cr, Co, Mo, Hg, Pb, Mn, Ni and Zn), radionuclides (137Cs, 90Sr, 238U and 234U), metalloids (Sb and As), non-metals (Se) and some organic substances (mainly hydrocarbons) since these do not usually have their structures further transformed or changed inside the plant system [31]. The occurrence of these contaminants in soil and water is the consequence of either natural forces, such as rock disaggregation, volcanoes and soil erosion, or human actions, such as incomplete fossil combustion, refining of metals, mineral mining, solid waste dumping, electronic product manufacturing, agricultural chemicals, pigments, vehicular gas emissions, military functions, etc. Some heavy metals and their respective adverse impacts are outlined in Table 2 [7,28]. It is critical to conduct remediation operations to stop these heavy metals from penetrating terrestrial, aquatic and atmospheric environments, in addition to restoring the polluted ecosystem [32]. Therefore, phytoextraction is the technology through which plants extract heavy elements from water and soil media by creating complexes via chelation with elements/metals and their metabolites [13], thereby reducing their toxicity.
There are several factors responsible for the absorption mechanism of heavy metals, as discussed in the sections below.

3.1.1. Plant Species

The uptake of heavy metals is influenced by the nature of plant species. The special plants used for phytoextraction, named metallophytes, are divided into three categories, as follows [31]:
Metal excluders: These are the plants that prohibit metals from penetrating aerial shoots or retain low and stable levels of metal concentration across a wide variety of metal concentration levels in the soil. Some examples of excluder species include Oenothera biennis, Commelina communis, Salix spp., Silene maritime and Populus spp. [58];
Metal indicators: An indicator is defined as a plant species containing the same levels of heavy metals in its tissues as in the surrounding soil environment [59]. They can tolerate the present concentration level of metals by forming chelators (intracellular element-binding substances) or by sequestering metals within the plant biomass [23]. Such plants are of bio-ecological significance because they are indicators of pollution and also resemble accumulators in the way they take up pollutants. Examples are Pluchea dioscoridis, Agave sisalana and Cyperus articulates [60];
Hyperaccumulators: Hyperaccumulators are plants that contain higher levels of metal pollutants concentrated in their root systems, shoots or foliage [61]. The variation in the absorbing potential of different hyperaccumulators relies on the occurrence, regulation and expression of responsible genes and the adjoining environment [62]. Metal-accumulating plants can concentrate most heavy metals, such as Co, Cd, Zn, Ni, Pb and Mn, at levels equal to 100 or 1000 times those in excluder plant species [30]. To date, over 500 plants from above 40 genera have been recognized as natural hyperaccumulating species, comprising <0.2% of the total angiosperms [63]. Woody plant species, such as poplars and willows, are utilized for phytoextraction because of several advantages, as they have high biomass in comparison to shrubs and herbs, which promotes the absorption of high concentrations of metals in their aboveground shoots. Furthermore, they possess a deep root system, which lessens soil erosion and helps to avoid the movement of polluted soil to adjoining locations [32].

3.1.2. The Properties of the Medium

The remediation treatment for a contaminated ecosystem can be boosted by employing agronomical practices, such as fertilizers, chelator addition and adjustment of pH.

3.1.3. Root Zone

Plant roots can accumulate pollutants and sequester or metabolize them within plant parts. Roots release root exudates, which cause degradation of pollutants in the growing soil media.

3.1.4. Environmental Conditions

Environmental conditions affect the uptake mechanism. For instance, temperatures affect growth materials and, subsequently, the root length of the plant.

3.1.5. Chemical Properties of Contaminants

Phytoextraction relies on contaminant-specific hyperaccumulators and the consequential metabolic products of the contaminants within plants.

3.1.6. Bioavailability of Metals

The biological availability of metals in the soil is defined as the portion of the total metals in the interstitial space of the water and soil particles accessible to the receptor organisms [64]. Metal accumulation by plants is reliant on the bioavailability of metals, which in turn depends on the upholding capacity time of the metals, as well as their interaction with other compounds in the growing media. Plants affect the soil by enhancing the biological availability of metals by adding physicochemical and biodegradable factors.

3.1.7. Chelating Agent Addition

The uptake of heavy metals can be improved by adding biodegradable substances, such as micronutrients and chelating agents. Moreover, in the process of chelate-assisted remediation, chemical chelating agents, such as ethylenediaminetetraacetic acid (EDTA) and nitrilotriacetic acid (NTA), are supplied to improve the phytoextraction of soil-contaminating metals. Table 3 shows the suitable plant species for phytoextraction based on their underlying degradation mechanisms.

3.2. Rhizodegradation

Rhizoremediation is the degradation of different organic contaminants in growing soil through microbial activity, and it can be enhanced by the existing root zone. It is also called phytostimulation, plant-assisted remediation and rhizosphere biodegradation [72] (Figure 3a). Rhizodegradation is recognized as the most efficient method for polycyclic aromatic hydrocarbon (PAH) treatment in soil. The combined action of the microbial population and plant root system in the rhizosphere, such as the discharge of root exudates (amino acids, sugars, organic acids, etc.), hydrogen cyanide (HCN), siderophores and phosphatases and phytohormone production by plant growth-promoting rhizobacteria (PGPR), helps in the breakdown of contaminants [6]. PAHs originate from either natural or anthropogenic activities; among them, the combustion of organic matter is the major source of contamination (Figure 3b). The mechanism of rhizodegradation involves direct phytohormonal activity, enhancing nutrient availability and regulating microbial population, which makes it possible to degrade contaminants by using a growing root system [73]. The pollutants taken up by plants are either accumulated in harvestable parts or volatilized through leaves or stems (Figure 3c). Plant–microbe interactions, soil parameters and the conditions that promote degradation rates are the factors affecting the rate of rhizodegradation (Figure 3d). Soil characteristics limit the bioavailability of contaminants, and root–microbe associations are the key process in PAH phytoremediation. Several microbial enzymes that take part in the PAH breakdown process [6] and their respective biodegradation pathways are given in Table 4.

3.3. Phytostabilization

Phytostabilization, also known as phytoimmobilization, is the process of making a contaminant immobile (immovable) in soil media via (a) absorption by the plant roots, (b) adsorption on the roots or (c) precipitation in between the rhizospheres of growing media (Figure 4a). The technique reduces the mobility of contaminants by stabilizing them via the development of a vegetative cover at the root zone of the plant species, helping to avoid migration to the groundwater. Microbial growth correlated with the plant root system may enhance the breakdown of organic contaminants, such as petroleum hydrocarbons and pesticides, to likely non-toxic compounds [64]. Plant, soil, contaminant and environmental factors determine the success of phytostabilization technology in contaminated sites in terms of both revegetation and restoration (Figure 4b). Phytostabilization can be enhanced by increasing biological inoculants, such as plant growth-promoting bacteria (PGPB), organic amendments (biosolids, manures) and inorganic amendments (liming materials, clay materials and phosphate compounds), and through geotextile capping. Some of the suitable plants for phytostabilization and their respective biodegradation mechanisms are listed in Table 5.

3.4. Phytovolatilization

Phytovolatilization is the absorption, translocation and evaporation of a pollutant by a plant species, with a modified type of pollutant being discharged into the adjoining atmosphere. Phytoevaporation or phytovolatilization occurs when growing trees and other plant species consume water and pollutants simultaneously, resulting in volatilization through foliage or stems (direct phytovolatilization) or the soil surface because of the root activities of the plant (indirect phytovolatilization process) (Figure 5a). The phytoevaporation process can be employed for detoxification of both hazardous inorganic and organic contaminants [32]. Volatile types of hazardous inorganic substances, such as As, Se and Hg, can be evaporated from plant parts [27]. Examples of some plant species used for phytovolatilization are presented in Table 6.

3.5. Rhizofiltration

Rhizofiltration is a technique for the assimilation of pollutants present in a solution adjoining the root zone, concentrating and precipitating them on or into plant roots through biotic or abiotic methods (Figure 5b). This approach is mostly employed to rehabilitate polluted groundwater sources. Plant roots release root exudates (secondary metabolites) that undergo various biogeochemical processes, resulting in the precipitation of pollutants onto roots or into water bodies. Later on, adsorption of contaminants onto or into the plant root system and translocation to the aboveground portions occur based on the type of plant species and the pollutant type and concentration [10]. Plant species, the nutrient status and chemical characteristics of contaminants and the conditions of the groundwater are the prime factors that are used to evaluate the effectiveness of the rhizofiltration method in treating inorganic (metals) and organic pollutants present in groundwater resources. Some of the suitable plants for rhizofiltration and their pollutants are listed in Table 6.

3.6. Phytodegradation

The phytodegradation process involves the transformation of contaminants absorbed by plants through metabolic pathways inside the plant system (Figure 5c). It is also known as phytotransformation. The complex organic contaminants are converted into simpler products and integrated into the plant tissues to boost plant growth [94]. Plants release specific enzymes that catalyze and enhance the rate of transformation. The phytodegradation technique can be used for the treatment of chemical pollutants; for instance, chlorinated solvents, herbicides and explosives in groundwater sources, aromatic and petroleum hydrocarbons in growing soils and volatile substances from the atmosphere [95]. In environmental applications, this process relies upon the direct absorption of contaminants from soil water and the resultant metabolites accumulated within the plant tissues, provided that the accumulated metabolites are either non-toxic or less toxic than the original contaminant. Plants reported as suitable for phytodegradation are listed in Table 7.
Table 6. Suitable plants for phytoevaporation and rhizofiltration techniques.
Table 6. Suitable plants for phytoevaporation and rhizofiltration techniques.
Plant Species Contaminant References
Phytoevaporation
Arundo donaxAS[96]
Astragalus racemosusSe [43]
Brassica napusSe[97]
Medicago sativaChlorinated solvents[98]
Nicotiana tabacumHg[99]
Salix spp.Trichloroethylene (TCE), tetrachloroethylene (PCE)[27]
Populus spp.Trichloroethylene (TCE), tetrachloroethylene (PCE)[27]
Taxodium distichumTrichloroethylene (TCE)[27]
Rhizofiltration
Arundo donax L.Synthetic dye[10]
Echinodorus cordifoliusCd[100]
Eichhornia crassipes (Mart.) SolmsNi [10]
Heliconia psittacorumZn [100]
Iris pseuda-corusCr and Zn[43]
Lepironia articulatePb [43]
Pteris vittataAs [43]
Raphanus sativusU[101]

3.7. Phytodesalination

Phytodesalination is an eco-friendly technology that is employed to remediate water resources and sodic soils by using Na-hyperaccumulating halophytes. It is the process through which plant species are employed as desalination agents [107], which means the use of the halophytes’ ability to absorb huge quantities of sodium (Na+) ions from affected ecosystems and their elimination by means of accumulation and translocation to different harvestable plant parts (Figure 6a) [108]. Saline-sodic soils cover almost 6% (over 800 million ha) of the land surface worldwide and, to overcome this problem, several authors have emphasized the phytodesalination technique by proving the effectiveness of Na+-hyperaccumulating plants in desalinizing the soil, especially in arid and semi-arid locations [109]. This is because halophytes decrease the ion concentration level of the solution in the plant xylem and cause Na+ and Cl excretion at elevated levels of salt concentration [105]. The potential of phytodesalination is species-specific (Figure 6b) and dependent on soil and climatic conditions (primarily rainfall). Some of the reports utilized for phytodesalination purposes are shown in Table 7.

3.8. Phytohydraulics

Phytohydraulics is the exploitation of deep-rooted plant species to uptake, sequester and breakdown groundwater pollutants that emerge in contact with their root systems (Figure 7). A special class of plant species called phreatophytes are widely used for this purpose. Phreatophytes are deep-rooted, water-loving plants that have high transpiration rates and penetrate their roots into zones of high moisture, and they can also continue to exist under temporary saturation conditions [110]. Prosopis, Eucalyptus, Populus and Salix are typical phreatophytes.

4. Methods Used for Evaluation of Phytoremediation Potential

In attempts to evaluate the phytoremediation competence of different plants, three factors are mainly employed, as described below.

4.1. Biological Concentration Factor (BCF)

The bioconcentration value indicates the accumulation ability of plant roots in relation to the soil. It is estimated as the ratio of the metal/element concentration present in the plant roots to the metal/element concentration levels in soils [111] as determined with the following formula:
BCF = Metal   concentration   in   plant   root Metal   concentration   in   soil

4.2. Bioaccumulation Coefficient (BAC)

The BAC refers to the concentration levels of metals present in plant shoots divided by the metal concentration in the rhizosphere region of soil [112] and is estimated as follows:
BAC = Metal   concentration   level   in   plant   shoot Metal   concentration   level   in   soil

4.3. Translocation Factor (TF)

The TF is defined as the metal ratio of the transfer capacity from a plant’s roots to its shoots and is determined by the following formula [111]:
TF = Metal   concentration   present   in   plant   shoot Metal   concentration   in   root

5. Applications of Phytoremediation Technology

Phytoremediation technology can be employed for the purposes shown below (Figure 8).

5.1. Mitigation of Heavy Metal Contamination

Phytoremediation can assist in the mitigation of heavy metal contamination by using metal excluders [58], metal indicators [60] and hyperaccumulator plants [61].

5.2. Bioremediation of Air Pollutants

Phytofiltration is the use of vegetation to shield from dust. A green vegetation belt 8 m wide may cut down dust fall by two to three times. Various morphological characteristics of plant species, such as leaf orientation on the principal axis, surface nature, size and shape, the existence or lack of trichomes and wax accumulation, are responsible for the entrapment of pollutants from the ambient atmosphere [113]. Inorganic compounds, such as sulfur dioxide (SO2), nitrogen oxides (NOX), etc., undergo degradation inside plants. SO2 released due to fossil fuel burning generally enters plants through the stoma and can be exploited in a reductive sulfur cycle by transforming into SO42− or SO32− in plant cell walls. The final products are cysteine or other organic substances. Likewise, NOX, the predecessor of the photochemical reaction, can enter plant systems through deposition on foliage or roots. The factors affecting NOX penetration into leaves are species, the age of the plant, NOX concentration and other environmental factors.

5.3. Removal of Pesticides

Plant–microbe interactions help in the degradation of xenobiotics in the root zone of plants through the secretion of degrading enzymes [114]. For instance, bacteria segregated from insecticide-polluted soils possess the ability to break down atrazine [115]. Some of the suitable plants with specific pesticide degradation capacities are given in Table 8.

5.4. Reclamation of Abandoned Mine Sites

Reclamation of affected mining sites can restore soil fertility and recreational values, thereby increasing areas of rangeland and promoting the formation of wildlife habitats [68]. The major approaches involved in the rehabilitation of abandoned mine locations are metal accretion, translocation, accumulation and phytostabilization processes. Plants suitable for mine spoil treatment [116] are listed in Table 9.

5.5. Biodrainage

Biodrainage is the process of removing excess surface and subsurface water in waterlogged areas using vegetation. Plant species suitable for the biodrainage technique should possess the following characteristics: they should be able to withstand prolonged waterlogging, have a high transpiration rate and water use efficiency and be salt-tolerant and perennial. Examples include Eucalyptus tereticornis, Anthocephalus cadamba, Acacia nilotica, Pongamia pinnata, Bambusa bamboos, Prosopis juliflora, etc.

6. Advantages and Limitations of Phytoremediation

Phytoremediation has the following advantages [20,32,117]: (i) it can be carried out in both in situ and ex situ environments, (ii) the technology is amenable to a diverse variety of inorganic and organic substances, (iii) it is suited for remediation of large areas of terrestrial and aquatic ecosystems and can be simply implemented, (iv) the technique is suitable for places with shallow pollutants, (v) it is cost-effective compared to conventional methods as it does not require procurement of large machines, (vi) it is easy to implement and maintain and accepted by the public, (vii) growing trees on the contaminated sites makes them aesthetic and pleasing to the eyes and (viii) this approach can also recover the fertility of soil by releasing organic matter to the soil surface. The green technology of phytoremediation also has the following limitations: (i) the phytoremediation process takes years to rehabilitate a contaminated site, (ii) it is slower than conventional methods, (iii) the problem of accumulation of contaminants in plant fruit and other edible parts of vegetables and crops arises, (iv) the toxicity and biological availability of the biodegradation compounds are not identified, (v) the technology is controlled by soil and climatic factors and (vi) the process is not as effective for regions with a high concentration of pollutants. In order to overcome the aforementioned disadvantages and to improve the phytoremediation technology, genetically modified plants [118], microbial inoculants [119], microbiologically induced carbonate precipitation (MICP) techniques [120,121,122] and chelate-assisted approaches [32] have been emphasized.

7. Conclusions

The increasing population is creating the necessity for more developmental activities to address hunger and supplement livelihoods. Industrial development seems to be a viable option to address these two important needs, but increases in the toxicity of contaminants beyond the permissible limits due to these industries can threaten the biological systems of terrestrial and aquatic ecosystems, thus implying the need for environmental remediation/restoration. Hence, phytoremediation can be employed as a sustained, eco-friendly and cost-effective technology to rehabilitate these toxic pollutants by using green plants. Furthermore, phytoremediation mechanisms vary with the type of pollutant, the media in which remediation occurs and the scope of application. Phytoremediation could emerge as a holistic approach to environmental amelioration and livelihood security. The study also concluded that edible plants should be avoided because of the problem of accumulation of contaminants in plant tissues, from where they can enter into the food chain and cause chronic health issues. For effective treatment of polluted soil, air and water resources, identification of pollutant types and the underlying biodegradation method and selection of a plant species is critical. However, measures should be taken to increase awareness regarding bioremediation technology among native growers while focusing on fast-growing native plant species. Scientists, planners and policymakers, in collaboration with local inhabitants and concerned stakeholders, should focus on determining the opportunities for risk-free alternative techniques to restore polluted ecosystems.

Author Contributions

Conceptualization, S.K. and T.H.M.; Methodology, S.K., T.H.M. and N.A.P.; Data collection: S.K. and T.H.M.; writing—original draft preparation, S.K., T.H.M., S.M., J.A.M. and P.A.S.; Writing—review and editing, M.U.Z., A.K., S.K., N.A.P., R.K., A.K. and R.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are thankful to the Registrar of SKUAST-K, Srinagar, Jammu and Kashmir, for thoughtful comments and for providing the necessary facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Antoniadis, V.; Shaheen, S.M.; Levizou, E.; Shahid, M.; Niazi, N.K.; Vithanage, M.; Ok, Y.S.; Bolan, N.; Rinklebe, J. A critical prospective analysis of the potential toxicity of trace element regulation limits in soils worldwide: Are they protective concerning health risk assessment?—A review. Environ. Int. 2019, 127, 819–847. [Google Scholar] [CrossRef]
  2. Mehmood, M.A.; Ganie, S.A.; Bhat, R.A.; Rashid, A.; Rehman, S.; Dar, G.H.; Gulzar, A. Assessment of Some Trace Elements and Heavy Metals in river Jhelum of Kashmir Valley. Int. J. Theor. Appl. Sci. 2018, 10, 27–31. [Google Scholar]
  3. Rafique, N.; Tariq, S.R. Distribution and source apportionment studies of heavy metals in soil of cotton/wheat fields. Environ. Monit. Assess. 2016, 188, 309. [Google Scholar] [CrossRef]
  4. Iqbal, M.; Iqbal, N.; Bhatti, A.I.; Ahmad, N.; Zahid, M. Response surface methodology application in optimization of cadmium adsorption by shoe waste: A good option of waste mitigation by waste. Ecol. Eng. 2016, 88, 265–275. [Google Scholar] [CrossRef]
  5. Farahat, E.; Linderholm, H.W. The effect of long-term wastewater irrigation on accumulation and transfer of heavy metals in Cupressus sempervirens leaves and adjacent soils. Sci. Total Environ. 2015, 15, 1–7. [Google Scholar] [CrossRef]
  6. Shukla, K.P.; Sharma, S.; Kumar, N.; Singh, V.; Bisht, S.; Kumar, V. Rhizoremediation: A Promising Rhizosphere Technology. In Applied Bioremediation—Active and Passive Approaches; Patil, Y.B., Rao, P., Eds.; IntechOpen Limited: London, UK, 2013. [Google Scholar] [CrossRef] [Green Version]
  7. Mandal, A.; Purakayastha, T.J.; Ramana, S.; Neenu, S.; Bhaduri, D.; Chakraborty, K.; Manna, M.C.; Rao, A.S. Status on Phytoremediation of Heavy Metals in India—A Review. Int. J. Bio-Resour. Stress Manag. 2014, 5, 553–560. [Google Scholar] [CrossRef]
  8. Chen, B.; Stein, A.F.; Castell, N.; Gonzalez-Castanedo, Y.; Sanchez de la Campa, A.M.; de la Rosa, J.D. Modeling and evaluation of urban pollution events of atmospheric heavy metals from a large Cu-smelter. Sci. Total Environ. 2016, 1, 17–25. [Google Scholar] [CrossRef] [PubMed]
  9. Muradoglu, F.; Gundogdu, M.; Ercisli, S.; Encu, T.; Balta, F.; Jaafar, H.Z.E.; Zia-Ul-Haq, M. Cadmium toxicity affects chlorophyll a and b content, antioxidant enzyme activities and mineral nutrient accumulation in strawberry. Biol. Res. 2015, 48, 11. [Google Scholar] [CrossRef] [Green Version]
  10. Kristanti, R.A.; Ngu, W.J.; Yuniarto, A.; Hadibarata, T. Rhizofiltration for Removal of Inorganic and Organic Pollutants in Groundwater: A Review. Biointerface Res. Appl. Chem. 2021, 11, 12326–12347. [Google Scholar] [CrossRef]
  11. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. [Google Scholar] [CrossRef]
  12. Ahmaruzzaman, M. Industrial wastes as low-cost potential adsorbents for the treatment of wastewater laden with heavy metals. Adv. Colloid Interface Sci. 2011, 10, 36–59. [Google Scholar] [CrossRef]
  13. Yaqoob, A.; Nasim, F.H.; Sumreen, A.; Munawar, N.; Zia, M.A.; Choudhary, M.S.; Ashraf, M. Current scenario of phytoremediation: Progresses and limitations. Int. J. Biosci. 2019, 14, 191–206. [Google Scholar] [CrossRef]
  14. Vamerali, T.; Bandiera, M.; Mosca, G. Field crops for phytoremediation of metal-contaminated land. A review. Environ. Chem. Lett. 2010, 8, 1–17. [Google Scholar] [CrossRef]
  15. Robinson, B.H.; Lombi, E.; Zhao, F.J.; Mcgrath, S.P. Uptake and distribution of nickel and other metals in the hyperaccumulator Berkheya Coddii. New Phytol. 2003, 158, 279–285. [Google Scholar] [CrossRef] [Green Version]
  16. Borisev, M.; Pajevic, S.; Nikolic, N.; Pilipovic, A.; Krstic, B.; Orlovic, S. Phytoextraction of Cd, Ni, and Pb Using Four Willow Clones (Salix spp.). Pol. J. Environ. Stud. 2009, 18, 553–561. [Google Scholar]
  17. Tripathi, V.; Edrisi, S.A.; Abhilash, P.C. Towards the coupling of phytoremediation with bioenergy production. Renew. Sustain. Energy Rev. 2016, 57, 1386–1389. [Google Scholar] [CrossRef]
  18. Bokhari, S.H.; Ahmad, I.; Mahmood-Ul-Hassan, M.; Mohammad, A. Phytoremediation potential of Lemna minor L. for heavy metals. Int. J. Phytoremediat. 2016, 18, 25–32. [Google Scholar] [CrossRef]
  19. Wani, K.A.; Sofi, Z.M.; Malik, J.A.; Wani, J.A. Phytoremediation of Heavy Metals Using Salix (Willows). In Bioremediation and Biotechnology; Bhat, R., Hakeem, K., Dervash, M., Eds.; Springer: Cham, Switzerland, 2020; p. 2. [Google Scholar] [CrossRef]
  20. Farraji, H.; Zaman, N.Q.; Tajuddin, R.; Faraji, H. Advantages and disadvantages of phytoremediation: A concise review. Int. J. Environ. Sci. Technol. (IJEST) 2016, 2, 69–75. [Google Scholar]
  21. Pilon-Smits, E. Phytoremediation. Annu. Reviewof. Plant Biol. 2005, 56, 15–39. [Google Scholar] [CrossRef]
  22. Cunningham, S.D.; Anderson, T.A.; Schwab, A.P.; Hsu, F.C. Phytoremediation of Soils Contaminated with Organic Pollutants. In Advances in Agronomy; Academic Press Inc.: Cambridge, MA, USA, 1996; Volume 56, pp. 55–114. [Google Scholar] [CrossRef]
  23. Ghosh, M.; Singh, S.P. A Review on Phytoremediation of Heavy Metals and Utilization of Its Byproducts. Appl. Ecol. Environ. Res. 2005, 3, 1–18. [Google Scholar] [CrossRef]
  24. Hutchinson, S.L.; Banks, M.K.; Schwab, A.P. Phytoremediation of aged petroleum sludge: Effect of inorganic fertilizer. J. Environ. Qual. 2001, 30, 395–403. [Google Scholar] [CrossRef]
  25. Joner, E.; Leyval, C. Phytoremediation of organic pollutants using mycorrhizal plants: A new aspect of rhizosphere interactions. Agronomie 2003, 23, 495–502. [Google Scholar] [CrossRef]
  26. Awa, S.H.; Hadibarata, T. Removal of Heavy Metals in Contaminated Soil by Phytoremediation Mechanism: A review. Water Air Soil Pollut. 2020, 231, 47. [Google Scholar] [CrossRef]
  27. Limmer, M.; Burken, J. Phytovolatilization of Organic Contaminants. Environ. Sci. Technol. 2016, 50, 6632–6643. [Google Scholar] [CrossRef] [PubMed]
  28. Bhat, S.A.; Bashir, O.; Haq, S.A.U.; Amin, T.; Rafiq, A.; Ali, M.; Americo-Pinheiro, J.H.P.; Sher, F. Phytoremediation of heavy metals in soil and water: An eco-friendly, sustainable and multidisciplinary approach. Chemosphere 2022, 303, 100774. [Google Scholar] [CrossRef]
  29. Kumar, A.; Kumar, A.; Mms, C.-P.; Chaturvedi, A.K.; Shabnam, A.A.; Subrahmanyam, G.; Mondal, R.; Gupta, D.K.; Malyan, S.K.; Kumar, S.S.; et al. Lead Toxicity: Health Hazards, Influence on Food Chain, and Sustainable Remediation Approaches. Int. J. Environ. Res. Public Health 2020, 17, 2179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Tangahu, B.V.; Abdullah, S.R.S.; Basri, H.; Idris, M.; Anuar, N.; Mukhlisin, M. A Review on Heavy Metals (As, Pb, and Hg) Uptake by Plants through Phytoremediation. Int. J. Chem. Eng. 2011, 2011, 939161. [Google Scholar] [CrossRef]
  31. Babu, S.M.O.F.; Hossain, M.B.; Rahman, M.S.; Rahman, M.; Ahmed, A.S.S.; Hasan, M.M.; Rakib, A.; Emran, T.B.; Xiao, J.; Simal-Gandara, J. Phytoremediation of Toxic Metals: A Sustainable Green Solution for Clean Environment. Appl. Sci. 2021, 11, 10348. [Google Scholar] [CrossRef]
  32. Yan, A.; Wang, Y.; Tan, S.N.; Mohd –Yusof, M.L.; Ghosh, S.; Chen, Z. Phytoremediation: A Promising Approach for Revegetation of Heavy Metal-Polluted Land. Front. Plant Sci. 2020, 11, 359. [Google Scholar] [CrossRef]
  33. Finnegan, P.M.; Chen, W. Arsenic toxicity: The effects on plant metabolism. Front. Physiol. 2012, 6, 182. [Google Scholar] [CrossRef] [Green Version]
  34. Mandal, P. An insight of environmental contamination of arsenic on animal health. Emerg. Contam. 2017, 3, 17–22. [Google Scholar] [CrossRef]
  35. Han, J.; Park, H.; Kim, J.; Jeong, D.; Kang, J. Toxic effects of arsenic on growth, hematological parameters, and plasma components of starry flounder, Platichthys stellatus, at two water temperature conditions. Fish. Aquat. Sci. 2019, 22, 3. [Google Scholar] [CrossRef]
  36. Kumar, A.; Subrahmanyam, G.; Mondal, R.; Cabral-Pinto, M.; Shabnam, A.A.; Jigyasu, D.K.; Malyan, S.K.; Fagodiya, R.K.; Khan, S.A.; Yu, Z.-G. Bio-remediation approaches for alleviation of cadmium contamination in natural resources. Chemosphere 2021, 268, 128855. [Google Scholar] [CrossRef] [PubMed]
  37. Environmental Protection Agency (EPA). Aquatic Life Ambient Water Quality Criteria Update for Cadmium—2016; U.S. Environmental Protection Agency Office of Water Office of Science and Technology: Washington, DC, USA, 2016. Available online: http://www.epa.gov/wqc/aquatic-life-criteria-cadmium (accessed on 11 December 2022).
  38. Genchi, G.; Sinicropi, M.S.; Lauria, G.; Carocci, A.; Catalano, A. The Effects of Cadmium Toxicity. Int. J. Environ. Res. Public Health 2020, 26, 3782. [Google Scholar] [CrossRef] [PubMed]
  39. Haider, F.U.; Liqun, C.; Coulter, J.A.; Cheema, S.A.; Wu, J.; Zhang, R.; Wenjun, M.; Farooq, M. Cadmium toxicity in plants: Impacts and remediation strategies. Ecotoxicol. Environ. Saf. 2011, 211, 111887. [Google Scholar] [CrossRef]
  40. Outridge, P.M.; Scheuhammer, A.M. Bioaccumulation and toxicology of chromium: Implications for wildlife. Rev. Environ. Contam. Toxicol. 1993, 130, 31–77. [Google Scholar] [CrossRef]
  41. Aslam, S.; Yousafzai, A.M. Chromium toxicity in fish: A review article. J. Entomol. Zool. Stud. 2017, 5, 1483–1488. [Google Scholar]
  42. Kapoor, R.T.; Bani- Mfarrej, M.F.; Alam, P.; Rinklebe, J.; Ahmad, P. Accumulation of chromium in plants and its repercussion in animals and humans. Environ. Pollut. 2022, 15, 119044. [Google Scholar] [CrossRef] [PubMed]
  43. Wani, Z.A.; Ahmad, Z.; Asgher, M.; Bhat, J.A.; Sharma, M.; Kumar, A.; Sharma, V.; Kumar, A.; Pant, S.; Lukatkin, A.S.; et al. Phytoremediation of Potentially Toxic Elements: Role, Status and Concerns. Plants 2023, 12, 429. [Google Scholar] [CrossRef]
  44. Wang, Z.; He, N.; Wang, Y.; Zhang, J. Effects of Copper on Organisms: A Review. Adv. Mater. Res. 2013, 726–731, 340–343. [Google Scholar] [CrossRef]
  45. Environmental Protection Agency (EPA). Aquatic Life Ambient Freshwater Quality Criteria—Copper 2007 Revision; U.S. Environmental Protection Agency Office of Water Office of Science and Technology: Washington, DC, USA, 2007.
  46. Blakley, B.R. Lead Poisoning in Animals. In Toxicology—MSD Veterinary Manual; Saskatoon, SK, Canada, 2022; pp. 1–5. Available online: https://www.msdvetmanual.com/toxicology/lead-poisoning/lead-poisoning-in-animals (accessed on 27 January 2023).
  47. Milatovic, D.; Gupta, R.C. Manganese. In Veterinary Toxicology, 3rd ed.; Gupta, R.C., Ed.; Academic Press: Cambridge, MA, USA, 2018; pp. 445–454. [Google Scholar] [CrossRef]
  48. Rodrigues, G.Z.P.; de Souza, M.S.; Gehlen, G. Impacts Caused by Manganese in the Aquatic Environments of Brazil. In Pollution of Water Bodies in Latin America; Gomez-Olivan, L., Ed.; Springer: Cham, Switzerland, 2019. [Google Scholar] [CrossRef]
  49. Zheng, N.; Wang, S.; Dong, W.; Hua, X.; Li, Y.; Song, X.; Chu, Q.; Hou, S.; Li, Y. The Toxicological Effects of Mercury Exposure in Marine Fish. Bull Environ. Contam. Toxicol. 2019, 102, 714–720. [Google Scholar] [CrossRef]
  50. Gworek, B.; Dmuchowski, W.; Baczewska-Dąbrowska, A.H. Mercury in the terrestrial environment: A review. Environ. Sci. Eur. 2020, 32, 1–19. [Google Scholar] [CrossRef]
  51. Bittner, F. Molybdenum metabolism in plants and crosstalk to iron. Front. Plant Sci. 2014, 5, 28. [Google Scholar] [CrossRef] [Green Version]
  52. Novotny, J.A.; Peterson, C.A. Molybdenum. Adv. Nutr. 2018, 9, 272–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Al-Attar, A.M. The Influences of Nickel Exposure on Selected Physiological Parameters and Gill Structure in the Teleost Fish, Oreochromis niloticus. J. Biol. Sci. 2007, 7, 77–85. Available online: https://scialert.net/abstract/?doi=jbs.2007.77.85 (accessed on 12 November 2022). [CrossRef] [Green Version]
  54. Kumar, A.; Jigyasu, D.K.; Subrahmanyam, G.; Mondal, R.; Shabnam, A.A.; Cabral-Pinto, M.; Malyan, S.K.; Chaturvedi, A.K.; Gupta, D.K.; Fagodiya, R.K.; et al. Nickel in terrestrial biota: Comprehensive review on contamination, toxicity, tolerance and its remediation approaches. Chemosphere 2021, 275, 129996. [Google Scholar] [CrossRef] [PubMed]
  55. Hassan, M.U.; Chattha, M.U.; Khan, I.; Chattha, M.B.; Aamer, M.; Nawaz, M.; Ali, A.; Khan, M.A.U.; Khan, T.A. Nickel toxicity in plants: Reasons, toxic effects, tolerance mechanisms, and remediation possibilities—A review. Environ. Sci. Pollut. Res. Int. 2019, 26, 12673–12688. [Google Scholar] [CrossRef] [PubMed]
  56. Plum, L.M.; Rink, L.; Haase, H. The essential toxin: Impact of zinc on human health. Int. J. Environ. Res. Public Health 2010, 7, 1342–1365. [Google Scholar] [CrossRef] [Green Version]
  57. Kaur, H.; Garg, N. Zinc toxicity in plants: A review. Planta 2021, 253, 129. [Google Scholar] [CrossRef]
  58. Mehes-Smith, M.; Nkongolo, K.; Cholew, E. Coping Mechanisms of Plants to Metal Contaminated Soil. In Environmental Change and Sustainability; Silvern, S., Young, S., Eds.; Intech Open: London, UK, 2013. [Google Scholar] [CrossRef] [Green Version]
  59. Baker, A.J.M.; Walker, P.L. Ecophysiology of metal uptake by tolerant plants: Heavy metal tolerance in plants. In Evolutionary Aspects; Shaw, A.J., Ed.; CRC Press: Boca Raton, FL, USA, 1990. [Google Scholar]
  60. Mganga, N.; Manoko, M.L.K.; Rulangaranga, Z.K. Classification of Plants According to Their Heavy Metal Content around North Mara Gold Mine, Tanzania: Implication for Phytoremediation. Tanzan. J. Sci. 2011, 37, 109–119. [Google Scholar]
  61. Cunningham, S.D.; Ow, D.W. Promises and Prospects of Phytoremediation. Plant Physiol. 1996, 110, 715–719. [Google Scholar] [CrossRef] [PubMed]
  62. Rascio, N.; Navari-Izzo, F. Heavy Metal Hyperaccumulating Plants: How and Why Do They Do It? And What Makes Them So Interesting? Plant Sci. 2011, 180, 169–181. [Google Scholar] [CrossRef]
  63. Bian, X.; Cui, J.; Tang, B.; Yang, L. Chelant-Induced Phytoextraction of Heavy Metals from Contaminated Soils: A Review. Pol. J. Environ. Stud. 2018, 27, 2417–2424. [Google Scholar] [CrossRef] [Green Version]
  64. Bolan, N.S.; Park, J.H.; Robinson, B.; Naidu, R.; Huh, K.Y. Phytostabilization: A Green Approach to Contaminant Containment. In Advances in Agronomy; Sparks, D.L., Ed.; Academic Press: Cambridge, MA, USA, 2011; Volume 112, pp. 145–204. [Google Scholar] [CrossRef]
  65. Singh, S.; Karwadiya, J.; Srivastava, S.; Patra, P.K.; Venugopalan, V.P. Potential of indigenous plant species for phytoremediation of arsenic contaminated water and soil. Ecol. Eng. 2022, 175, 106476. [Google Scholar] [CrossRef]
  66. Patel, S.J.; Bhattacharya, P.; Banu, S.; Bai, L.; Namratha. Phytoremediation of Copper and Lead by Using Sunflower, Indian Mustard and Water Hyacinth Plants. Int. J. Sci. Res. (IJSR) 2015, 4, 113–115. [Google Scholar]
  67. Bhattacharya, E.; Biswas, S.M. First Report of the Hyperaccumulating Potential of Cadmium and Lead by Cleome rutidosperma DC. With a Brief Insight into the Chemical Vocabulary of its Roots. Front. Environ. Sci. 2022, 10, 87. [Google Scholar] [CrossRef]
  68. Reddy, L.C.S.; Reddy, K.V.R.; Humane, S.K.; Damodaram, B. Accumulation of Chromium in Certain plant Species Growing on Mine Dump from Byrapur, Karnataka, India. Res. J. Chem. Sci. 2012, 2, 17–20. [Google Scholar]
  69. Kumari, P.; Kumar, P.; Kumar, T. An overview of phytomining: A metal extraction process from plant species. J. Emerg. Technol. Innov. Res. (JETIR) 2019, 6, 1367–1376. [Google Scholar]
  70. Capuana, M. A review of the performance of woody and herbaceous ornamental plants for phytoremediation in urban areas. Iforest—Biogeosci. For. 2020, 13, 139–151. [Google Scholar] [CrossRef] [Green Version]
  71. Malik, J.A.; Wani, A.A.; Wani, K.A.; Bhat, M.A. Role of White Willow (Salix alba L.) for Cleaning Up the Toxic Metal Pollution. In Bioremediation and Biotechnology; Hakeem, K., Bhat, R., Qadri, H., Eds.; Springer: Cham, Switzerland, 2020. [Google Scholar] [CrossRef]
  72. Kumar, A.; Mishra, S.; Pandey, R.; Yu, Z.G.; Kumar, M.; Khoo, K.S.; Thakur, T.K.; Show, P.L. Microplastics in terrestrial ecosystems: Un-ignorable impacts on soil characterises, nutrient storage and its cycling. TrAC Trends Anal. Chem. 2023, 158, 116869. [Google Scholar] [CrossRef]
  73. Lee, S.; Ka, J.O.; Gyu, S.H. Growth promotion of Xanthium italicum by application of rhizobacterial isolates of Bacillus aryabhattai in microcosm soil. J. Microbiol. 2012, 50, 45–49. [Google Scholar] [CrossRef] [PubMed]
  74. Pimviriyakul, P.; Wongnate, T.; Tinikul, R.; Chaiyen, P. Microbial degradation of halogenated aromatics: Molecular mechanisms and enzymatic reactions. Microb. Biotechnol. 2020, 13, 67–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Sood, S.; Sharma, A.; Sharma, N.; Kanwar, S.S. Carboxylesterases: Sources, Characterization and Broader Applications. Insights Enzyme Res. 2016, 1, 1–11. [Google Scholar] [CrossRef]
  76. Raunio, H.; Kuusisto, M.; Juvonen, R.O.; Pentikäinen, O.T. Modeling of interactions between xenobiotics and cytochrome P450 (CYP) enzymes. Front. Pharmacol. 2015, 6, 123. [Google Scholar] [CrossRef] [PubMed]
  77. Lee, M.D.; Odom, J.M.; Buchanan, R.J. New perspectives on microbial dehalogenation of chlorinated solvents: Insights from the field. Annu. Rev. Microbiol. 1998, 52, 423–452. [Google Scholar] [CrossRef]
  78. Akram, M.S.; Rashid, N.; Basheer, S. Physiological and molecular basis of plants tolerance to linear halogenated hydrocarbons. In Handbook of Bioremediation; Hasanuzzaman, M., Prasad, M.N.V., Eds.; Academic Press: Cambridge, MA, USA, 2021; pp. 591–602. [Google Scholar] [CrossRef]
  79. Karich, A.; Ullrich, R.; Scheibner, K.; Hofrichter, M. Fungal Unspecific Peroxygenases Oxidize the Majority of Organic EPA Priority Pollutants. Front. Microbiol. 2017, 8, 1463. [Google Scholar] [CrossRef] [Green Version]
  80. Christian, V.; Shrivastava, R.; Shukla, D.; Modi, H.A.; Vyas, B.R.M. Degradation of xenobiotic compounds by lignin-degrading white-rot fungi: Enzymology and mechanisms involved. Indian J. Exp. Biol. 2005, 43, 301–312. [Google Scholar]
  81. Balcazar-Lopez, E.; Mendez-Lorenzo, L.H.; Batista-Garcia, R.A.; Esquivel-Naranjo, U.; Ayala, M.; Kumar, V.V.; Savary, O.; Cabana, H.; Herrera-Estrella, A.; Folch-Mallol, J.L. Xenobiotic Compounds Degradation by Heterologous Expression of a Trametes sanguineus Laccase in Trichoderma atroviride. PLoS ONE 2016, 5, e147997. [Google Scholar] [CrossRef] [Green Version]
  82. Schaffner, A.; Messner, B.; Langebartels, C.; Sandermann, H. Genes and Enzymes for In-Planta Phytoremediation of Air, Water and Soil. Acta Biotechnol. 2002, 22, 141–152. [Google Scholar] [CrossRef]
  83. Ogunyemi, A.K.; Buraimoh, O.M.; Ogunyemi, B.C.; Samuel, T.A.; Ilori, M.O.; Amund, O.O. Nitrilase gene detection and nitrile metabolism in two bacterial strains associated with waste streams in Lagos, Nigeria. Bull Natl. Res. Cent. 2022, 46, 151. [Google Scholar] [CrossRef]
  84. Kitts, C.L.; Green, C.E.; Otley, R.A.; Alvarez, M.A.; Unkefer, P.A. Type I nitroreductases in soil enterobacteria reduce TNT (2,4,6-trinitrotoluene) and RDX (hexahydro-1,3,5-trinitro-1,3,5-triazine). Can. J. Microbiol. 2000, 46, 278–282. [Google Scholar] [CrossRef] [PubMed]
  85. Chen, Q.; Wang, C.H.; Deng, S.K.; Wu, Y.D.; Li, Y.; Yao, L.; Jiang, J.D.; Yan, X.; He, J.; Li, S.P. Novel three-component Rieske non-heme iron oxygenase system catalyzing the N-dealkylation of chloroacetanilide herbicides in sphingomonads DC-6 and DC-2. Appl. Environ. Microbiol. 2014, 80, 5078–5085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Karagianni, E.P.; Kontomina, E.; Davis, B.; Kotseli, B.; Tsirka, T.; Garefalaki, V.; Sim, E.; Glenn, A.E.; Boukouvala, S. Homologues of xenobiotic metabolizing N-acetyltransferases in plant-associated fungi: Novel functions for an old enzyme family. Sci. Rep. 2015, 5, 12900. [Google Scholar] [CrossRef] [Green Version]
  87. Kolhe, P.M.; Ingle, S.T.; Wagh, N.D. Degradation of Phenol Containing Wastewater by Advance Catalysis System—A Review. Annu. Res. Rev. Biol. 2015, 8, 1–15. [Google Scholar] [CrossRef]
  88. Ambreen, S.; Yasmin, A.; Aziz, S. Isolation and characterization of organophosphorus phosphatases from Bacillus thuringiensis MB497 capable of degrading Chlorpyrifos, Triazophos and Dimethoate. Heliyon 2020, 6, e04221. [Google Scholar] [CrossRef]
  89. Peco, J.D.; Higueras, P.; Campos, J.A.; Esbri, 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. Lorestani, B.; Yousefi, N.; Cheraghi, M.; Farmany, A. Phytoextraction and phytostabilization potential of plants grown in the vicinity of heavy metal-contaminated soils: A case study at an industrial town site. Environ. Monit. Assess. 2013, 185, 10217–10223. [Google Scholar] [CrossRef]
  91. Wu, B.; Peng, H.; Sheng, M.; Luo, H.; Wang, X.; Zhang, R.; Xu, F.; Xu, H. Evaluation of phytoremediation potential of native dominant plants and spatial distribution of heavy metals in abandoned mining area in Southwest China. Ecotoxicol. Environ. Saf. 2021, 220, 112368. [Google Scholar] [CrossRef] [PubMed]
  92. Zeng, P.; Guo, Z.; Cao, X.; Xiao, X.; Liu, Y.; Shi, L. Phytostabilization potential of ornamental plants grown in soil contaminated with cadmium. Int. J. Phytoremediat. 2018, 20, 311–320. [Google Scholar] [CrossRef]
  93. Siyar, R.; Doulati-Ardejani, F.; Norouzi, P.; Maghsoudy, S.; Yavarzadeh, M.; Taherdangkoo, R.; Butscher, C. Phytoremediation Potential of Native Hyperaccumulator Plants Growing on Heavy Metal-Contaminated Soil of Khatunabad Copper Smelter and Refinery, Iran. Water 2022, 14, 3597. [Google Scholar] [CrossRef]
  94. Greipsson, S. Phytoremediation. Nat. Educ. Knowl. 2011, 3, 7. [Google Scholar]
  95. Newman, L.A.; Reynolds, C.M. Phytodegradation of organic compounds. Curr. Opin. Biotechnol. 2004, 15, 225–230. [Google Scholar] [CrossRef] [PubMed]
  96. Guarino, F.; Miranda, A.; Castiglione, S.; Cicatelli, A. Arsenic phytovolatilization and epigenetic modifications in Arundo donax L. assisted by a PGPR consortium. Chemosphere 2020, 251, 126310. [Google Scholar] [CrossRef] [PubMed]
  97. Nedjimi, B. Phytoremediation: A sustainable environmental technology for heavy metals decontamination. SN Appl. Sci. 2021, 3, 286. [Google Scholar] [CrossRef]
  98. Narayanan, M.; Davis, L.C.; Erickson, L.E. Fate of volatile chlorinated organic compounds in a laboratory chamber with alfalfa plants. Environ. Sci. Technol. 1995, 29, 2437–2444. [Google Scholar] [CrossRef]
  99. Ashraf, M.; Ozturk, M.; Ahmad, M.S.A. (Eds.) Toxins and their phytoremediation. In Plant Adaptation and Phytoremediation; Springer: Berlin/Heidelberg, Germany, 2010; pp. 1–32. [Google Scholar]
  100. Woraharn, S.; Meeinkuirt, W.; Phusantisampan, T.; Avakul, P. Potential of ornamental monocot plants for rhizofiltration of cadmium and zinc in hydroponic systems. Environ. Sci. Pollut. Res. 2021, 28, 35157–35170. [Google Scholar] [CrossRef]
  101. Han, Y.; Lee, J.; Kim, C.; Park, J.; Lee, M.; Yang, M. Uranium Rhizofiltration by Lactuca sativa, Brassica campestris L., Raphanus sativus L., Oenanthe javanica under Different Hydroponic Conditions. Minerals 2021, 11, 41. [Google Scholar] [CrossRef]
  102. Kafle, A.; Timilsina, A.; Gautam, A.; Adhikari, K.; Bhattarai, A.; Aryal, N. Phytoremediation: Mechanisms, plant selection and enhancement by natural and synthetic agents. Environ. Adv. 2022, 8, 100203. [Google Scholar] [CrossRef]
  103. Jha, P.; Jobby, R.; Desai, N.S. Remediation of textile azo dye acid red 114 by hairy roots of Ipomoea carnea Jacq. and assessment of degraded dye toxicity with human keratinocyte cell line. J. Hazard. Mater. 2016, 311, 158–167. [Google Scholar] [CrossRef]
  104. Doty, S.L.; Shang, T.Q.; Wilson, A.M.; Moore, A.L.; Newman, L.A.; Strand, S.E.; Gordon, M.P. Metabolism of the soil and groundwater contaminants, ethylene dibromide and trichloroethylene, by the tropical leguminous tree, Leuceana leucocephala. Water Res. 2003, 37, 441–449. [Google Scholar] [CrossRef]
  105. Saddhe, A.A.; Manuka, R.; Nikalje, G.C.; Penna, S. Halophytes as a Potential Resource for Phytodesalination. In Handbook of Halophytes; Grigore, M.N., Ed.; Springer: Cham, Switzerland, 2020. [Google Scholar] [CrossRef]
  106. Rozema, E.R.; Gordon, R.J.; Zheng, Y. Plant species for the removal of Na+ and Cl from greenhouse nutrient solution. Hort. Sci. 2014, 49, 1071–1075. [Google Scholar] [CrossRef]
  107. Sarath, N.G.; Sruthi, P.; Shackira, A.M.; Puthur, J.T. Halophytes as effective tool for phytodesalination and land reclamation. In Frontiers in Plant-Soil Interaction; Academic Press: Cambridge, MA, USA, 2021; pp. 459–494. [Google Scholar] [CrossRef]
  108. Lastiri-Hernandez, M.A.; Alvarez-Bernal, D.; Bermudez-Torres, K.; Cardenas, G.C.; Ceja-Torres, L.F. Phytodesalination of a moderately saline soil combined with two inorganic amendments. Bragantia 2019, 78, 579–586. [Google Scholar] [CrossRef] [Green Version]
  109. Zorrig, W.; Rabhi, M.; Ferchichi, S.; Smaouti, A.; Abdelly, C. Phytodesalination: A Solution for Salt-affected Soils in Arid and Semi-arid Regions. J. Arid. Land Stud. 2012, 22, 299–302. [Google Scholar]
  110. Gatliff, E.G. Vegetative Remediation Process Offers Advantages over Traditional Pump and-Treat Technologies. Remediation 1994, 4, 343–352. [Google Scholar] [CrossRef]
  111. Ahmadpour, P.; Ahmadpour, F.; Mahmud, T.M.M.; Abdu, A.; Soleimani, M.; Tayefeh, F.H. Phytoremediation of heavy metals: A green technology. Afr. J. Biotechnol. 2012, 11, 14036–14043. [Google Scholar]
  112. Hasnaoui, S.E.; Fahr, M.; Keller, C.; Levard, C.; Angeletti, B.; Chaurand, P.; Triqui, Z.E.A.; Guedira, A.; Rhazi, L.; Colin, F.; et al. Screening of Native Plants Growing on a Pb/Zn Mining Area in Eastern Morocco: Perspectives for Phytoremediation. Plants 2020, 9, 1458. [Google Scholar] [CrossRef] [PubMed]
  113. Singh, S.N.; Tripathi, R.D. Environmental Bioremediation Technologies; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2007; pp. 1–47. [Google Scholar]
  114. Henderson, K.L.; Belden, J.B.; Zhao, S.; Coats, J.R. Phytoremediation of pesticide wastes in soil. Z. Für Nat. 2006, 61, 213–221. [Google Scholar] [CrossRef] [Green Version]
  115. Zhao, S.; Arthur, E.L.; Coats, J.R. Influence of microbial inoculation (Pseudomonas sp. strain ADP), the enzyme atrazine chlorohydrolase, and vegetation on the degradation of atrazine and metolachlor in soil. J. Agric. Food Chem. 2003, 51, 3043–3048. [Google Scholar] [CrossRef] [Green Version]
  116. Prasad, M.N.V. Phytoremediation in India. In Phytoremediation: Methods and Review; Neil Willey Center for Research in Plant Science, University of the West of England: Bristol, UK, 2007; pp. 435–454. [Google Scholar]
  117. Tripathi, S.; Singh, V.K.; Srivastava, P.; Singh, R.; Devi, R.S.; Kumar, A.; Bhadouria, R. Phytoremediation of organic pollutants: Current status and future directions. In Abatement of Environmental Pollutants; Singh, P., Kumar, A., Borthakur, A., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 81–105. [Google Scholar] [CrossRef]
  118. 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]
  119. Raklami, A.; Meddich, A.; Oufdou, K.; Baslam, M. Plants-Microorganisms-Based Bioremediation for Heavy Metal Cleanup: Recent Developments, Phytoremediation Techniques, Regulation Mechanisms, and Molecular Responses. Int. J. Mol. Sci. 2022, 23, 5031. [Google Scholar] [CrossRef]
  120. Song, H.W.; Wang, C.C.; Kumar, A.; Ding, Y.; Li, S.; Bai, X.; Liu, T.; Wang, J.L.; Zhang, Y.L. Removal of Pb2+ and Cd2+ from contaminated water using novel microbial material (Scoria@UF1). J. Environ. Chem. Eng. 2021, 9, 106495. [Google Scholar] [CrossRef]
  121. Kumar, A.; Song, H.W.; Mishra, S.; Zhang, W.; Zhang, Y.L.; Zhang, Q.R.; Yu, Z.G. Application of microbial-induced carbonate precipitation (MICP) techniques to remove heavy metal in the natural environment: A critical review. Chemosphere 2023, 318, 137894. [Google Scholar] [CrossRef] [PubMed]
  122. Song, H.; Kumar, A.; Zhang, Y. Microbial-induced carbonate precipitation prevents Cd2+ migration through the soil profile. Sci. Total Environ. 2022, 844, 157167. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Characteristics of the ideal plants for phytoremediation.
Figure 1. Characteristics of the ideal plants for phytoremediation.
Water 15 01498 g001
Figure 2. Mechanism of phytoextraction.
Figure 2. Mechanism of phytoextraction.
Water 15 01498 g002
Figure 3. Rhizodegradation techniques: (a) chemical structures of different organic contaminants, (b) sources of PAH contamination, (c) mechanisms of rhizoremediation, and (d) factors affecting the rate of rhizodegradation.
Figure 3. Rhizodegradation techniques: (a) chemical structures of different organic contaminants, (b) sources of PAH contamination, (c) mechanisms of rhizoremediation, and (d) factors affecting the rate of rhizodegradation.
Water 15 01498 g003
Figure 4. (a) Mechanisms of phytostabilization and (b) factors affecting the process of phytostabilization.
Figure 4. (a) Mechanisms of phytostabilization and (b) factors affecting the process of phytostabilization.
Water 15 01498 g004
Figure 5. Mechanisms of (a) phytovolatilization, (b) rhizofiltration, and (c) phytodegradation.
Figure 5. Mechanisms of (a) phytovolatilization, (b) rhizofiltration, and (c) phytodegradation.
Water 15 01498 g005
Figure 6. Phytodesalination process: (a) mechanism of phytodesalination and (b) factors affecting the mechanism of phytodesalination.
Figure 6. Phytodesalination process: (a) mechanism of phytodesalination and (b) factors affecting the mechanism of phytodesalination.
Water 15 01498 g006aWater 15 01498 g006b
Figure 7. Mechanism of phytohydraulics.
Figure 7. Mechanism of phytohydraulics.
Water 15 01498 g007
Figure 8. Phytoremediation applications.
Figure 8. Phytoremediation applications.
Water 15 01498 g008
Table 1. Various types of phytoremediation methods [28].
Table 1. Various types of phytoremediation methods [28].
Phytoremediation TypeContaminant NatureMediumMechanismScope of Application
Phytoextraction/phytoaccumulationInorganicsSoil, waterHyperaccumulation Moderately polluted sites
Rhizodegradation/phytostimulationOrganicsSoil Breakdown inside the rhizosphere through microbial activityPolycyclic aromatic hydrocarbon (PAH) contaminants
PhytostabilizationInorganicsSoilImmobilization Mining contamination
Phytovolatilization/phytoevaporationOrganics/inorganicsSoil, waterVolatilizationVolatile contaminants
RhizofiltrationOrganics/inorganicsWater Rhizosphere accumulationWastewater
Phytodegradation/phytotransformationComplex organicsWater, soil Breakdown inside the plant through metabolic processesSoil and wastewater contamination
PhytodesalinationOrganics/inorganicsSoil, waterNa hyperaccumulationSodic soil and water
Phytohydraulics Organics/inorganicsGround waterUptake, sequestering and degradation of groundwater contaminantsShallow contaminated sites
Table 2. Sources of heavy metal contamination and metals’ adverse impacts on terrestrial and aquatic life systems.
Table 2. Sources of heavy metal contamination and metals’ adverse impacts on terrestrial and aquatic life systems.
Name of the MetalSourceImpactsReferences
TerrestrialAquatic
HumansPlantsAnimals
Arsenic (As)Fuel burning, pesticides, painting, wood treatment, geothermal processes, natural forces, thermal power plants, smeltingCancer, skin lesions and increased deaths Inhibits root extension and proliferationAbdominal pain, vomiting and diarrhea; subsequently, rapid circulatory collapseBioaccumulation, physiological and biochemical disorders[33,34,35,36]
Cadmium (Cd)Fertilizers, fuel combustion, electroplating, smelting operations, batteries, dead batteries, paint sludgeLung, prostate, nasopharynx, pancreas, and kidney cancers, as well as itai-itaiReduces uptake and translocation of nutrients and water and disrupts metabolism Kidney, lung, bone, liver, blood and nervous system are affectedIncrease in mortality rates, deleterious effects on growth and reproduction systems[37,38,39]
Chromium (Cr)Mining operations, pesticide application, industrial coolant liquids, dyes, timber treatment, leather tanning, chromium salt manufacturingCancer, asthma, nose sores, skin infections, kidney and liver problemsInfluences crop growth rate, productivity and quality of grainsDetrimental effects on wild birds and mammalsCytotoxicity and detrimental impact[40,41,42,43]
Copper (Cu)Fertilizers, pigments, fungicides, mining, painting, electrical sources, lumber treatment, electroplating, smelting practices Inflammation, cancer and anemia Growth and development are blockedDisheveled feathers, gizzard erosion, intestinal inflammation, hematochezia and damaged kidneysAdverse effects on survival, growth and reproduction[44,45]
Lead (Pb)Metal products, paints, e-waste products, batteries, petrol additives, preservatives, ceramics, thermal power plants, bangle industryWeakness, hypertension, brain and kidney damage, impotence and miscarriage Poor germination, inhibits root growth and biomass productionSalivation, lack of vision, spastic twitching of eyelids, muscle tremors, jaw champing and convulsionsOxidative stress, neurotoxin, bioaccumulation [29,43,46]
Manganese (Mn)Application of fertilizerDeficits and neurodegenerative diseases, including a disorder called manganismTriggers oxidative stress and disrupt photosynthesisReduced feed intake and growth rate and lethargyIntestinal inflammatory damages, genotoxicity and oxidative stress[47,48]
Mercury (Hg)Fumigants, geothermal processes, fluorescent lights, chlor-alkali plants, hospital waste (broken thermometers, sphygmomanometers, barometers), thermal power plantsLoss of memory, kidney and nervous system problems and weakened hearing and vision abilityGrowth retardationAnorexia, stomatitis, vomiting, diarrhea, shock, pain and dehydrationTeratogenic, reproductive and neuro-toxicity [43,46,49,50]
Molybdenum (Mo)Fertilizer, spent catalystsPain in joints, gout-like signs and high blood levels of uric acidReduces seedling growth, yellowish leavesInduces secondary copper deficiency in animalsTransformations in the forms of aquatic biota systems and instability in fundamental activities[51,52]
Nickel (Ni)Alloys, mine tailings, battery manufacturers, smelting processes, thermal power systemsAllergy, kidney disorders, cardiovascular fibrosis, lung and nasal cancerLower seed germination, growth, biomass and final yield Lung disorders in rodents and affects liver, kidney, blood and reproduction processes in rats and mice Inhibition of respiration, ionoregulatory destruction and enhanced oxidative stress[53,54,55]
Zinc (Zn)Dyes, paints, fertilizers, galvanization processes, lumber treatment, mining, electroplating, smelting practicesNausea, back pain, vomiting, anemia and lethargy Slows down photosynthetic and respiratory rates and leads to unbalanced mineral nutrition Vomiting, diarrhea, depression, damage to red blood cells and lack of appetiteKills fish by destroying gill tissues[56,57]
Table 3. List of potential plants used for phytoextraction technology.
Table 3. List of potential plants used for phytoextraction technology.
Plant Species Heavy MetalAccumulation(A)/
Translocation (T)
Literature Cited
Althernanthera ficoidesAsA, T[65]
Brassica junceaPb A[66]
Cleome rutidosperma DCCd and PbA[67]
Helianthus annusCuA[66]
Hyptis suaveolensCrA[68]
Berkheya coddiiNiA[69]
Phragmites australisNi, Mo, Se and Cu A/T[43]
Populus speciesCdA[70]
Ricinus communisNi A[70]
Salix speciesCd and ZnA, T[71]
Senna siameaPb A[70]
Table 4. Important enzymes associated with the process of rhizodegradation.
Table 4. Important enzymes associated with the process of rhizodegradation.
EnzymeTarget PollutantBiodegradation PathwayReferences
Aromatic dehalogenaseChlorinated aromatics (DDT, PCBs, etc.)Hydrolytic dehalogenation[74]
CarboxylesterasesXenobioticsHydrolysis [75]
Cytochrome P450Xenobiotics (PCBs)Oxidation, reduction, hydrolysis and conjugation[76]
DehalogenaseChlorinated solvents and ethyleneDehalogenation [77]
Glutathione s-transferaseXenobioticsDehalogenation[78]
PeroxygenasesXenobioticsOxygenations and oxidations[79]
PeroxidasesXenobioticsOxidation and reduction[80]
LaccaseXenobiotics, degradation of explosivesOxidation[81]
N-glucosyl transferasesVarious xenobioticsConjugation[82]
NitrilaseHerbicidesDegradation of nitrile[83]
NitroreductaseExplosives (RDX and TNT)Reduction[84]
N-malonyl transferasesXenobioticsConjugation[82]
O-demethylaseAlachlor, metalachorN-dealkylation [85]
O-glucosyl transferasesXenobioticsConjugation[82]
O-malonyl transferasesXenobioticsHydrolysis[86]
PeroxdasePhenolsElimination or reduction[87]
PhosphataseOrganophosphatesHydrolase [88]
Table 5. Suitable plant species for phytostabilization.
Table 5. Suitable plant species for phytostabilization.
Plants PollutantMechanism References
Arundo donaxNi, Pb, HgDeposition[89]
Atriplex portulacoidesZnAdsorption[43]
Cirsium arvensePb, Mn, Zn Absorption/adsorption[90]
Conyza CanadensisCr, Ni, Cu, Pb, CdAccumulation[91]
Euonymus japonicusCdDeposition[92]
Launaea acanthodesNi, MoAccumulation[93]
Populus deltoidsAsDeposition[89]
Ricinus communisCd, Cu, Mn, ZnAbsorption/adsorption[43]
Salix purpureaAsDeposition[89]
Table 7. Examples of plant species suitable for the phytodegradation and phytodesalination.
Table 7. Examples of plant species suitable for the phytodegradation and phytodesalination.
Species PollutantBiodegradation PathwayReferences
Phytodegradation
Elodea CanadensisDDTCatalytic degradation[102]
Ipomoea carneaTextile azo dyesRedox reaction[103]
Populus spp.Trichloroethylene (TCE)Catalytic degradation[94]
Leucaena leucocephalaEthylene dibromideReduction [104]
Pueraria thunbergianaDDTDehalogenation [102]
Phytodesalination
Andropogon gerardiiNa+EC reduction and salt accumulation[105]
Atriplex prostrateNa+, ClAccumulate Na+ and Cl[106]
Phragmites australisNa+Extraction [105]
Typha latifoliaNa+, ClAccumulate Na+ and Cl[106]
Table 8. Examples of plants involved in the remediation of pesticides.
Table 8. Examples of plants involved in the remediation of pesticides.
Name of the Plant Pesticide
Populus spp.Atrazine
Corbicula flumineaCarbaryl, diazinon, carbofuran, glycophosphate, coumphos, parathion
Oryza sativaCarbaryl, parathion, atrazine, carbofuran, diazinon, coumphos, glycophosphate
Bassia scopariaAtrazine
Salix spp.2,4,5-T, 2,4-D, aldrin
Myriophyllum aquaticumOrgano-phosphate pesticides, halogenated pesticides
Elodea CanadensisOrgano-phosphate pesticides, halogenated pesticides
Spirodela oligorrhizaOrgano-phosphate pesticides
Table 9. Examples of plants involved in the reclamation of mining sites.
Table 9. Examples of plants involved in the reclamation of mining sites.
Category of Mine Spoils Plants
Coal mine spoils Eucalyptus hybrid, Pongamia pinnata, Acacia nilotica
Limestone mine spoils Salix tetrasperma, Acacia catechu, Leucaena leucocephala
Copper, tungiston, mica, limestone and marble mine spoils Prosopis juliflora, Acacia Senegal, Cynodon dactylon
Iron ore wasteLeucaena leucocephala
Manganese, haematite and magnetite spoilAlbizia lebeck
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khan, S.; Masoodi, T.H.; Pala, N.A.; Murtaza, S.; Mugloo, J.A.; Sofi, P.A.; Zaman, M.U.; Kumar, R.; Kumar, A. Phytoremediation Prospects for Restoration of Contamination in the Natural Ecosystems. Water 2023, 15, 1498. https://doi.org/10.3390/w15081498

AMA Style

Khan S, Masoodi TH, Pala NA, Murtaza S, Mugloo JA, Sofi PA, Zaman MU, Kumar R, Kumar A. Phytoremediation Prospects for Restoration of Contamination in the Natural Ecosystems. Water. 2023; 15(8):1498. https://doi.org/10.3390/w15081498

Chicago/Turabian Style

Khan, Shaista, Tariq H. Masoodi, Nazir A. Pala, Shah Murtaza, Javeed A. Mugloo, Parvez A. Sofi, Musaib U. Zaman, Rupesh Kumar, and Amit Kumar. 2023. "Phytoremediation Prospects for Restoration of Contamination in the Natural Ecosystems" Water 15, no. 8: 1498. https://doi.org/10.3390/w15081498

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