Next Article in Journal
Macro- and Microelements and Radionuclides in the Mussel Mytilus galloprovincialis from Recreational and Harbor Sites of the Crimean Peninsula (The Black Sea)
Previous Article in Journal
The Effect of Salinity and Light Intensity on the Batch Cultured Cyanobacteria Anabaena sp. and Cyanothece sp.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phycoremediation: Use of Algae to Sequester Heavy Metals

1
Department of Environmental Sciences, Central University of Jharkhand, Ranchi 835205, India
2
Department of Biology and Global Environmental Sustainability, Oral Roberts University, Tulsa, OK 74171, USA
*
Authors to whom correspondence should be addressed.
Hydrobiology 2022, 1(3), 288-303; https://doi.org/10.3390/hydrobiology1030021
Submission received: 29 January 2022 / Revised: 1 June 2022 / Accepted: 10 June 2022 / Published: 1 July 2022

Abstract

:
Industrialization, natural processes, and urbanization have potentially accelerated the pace and the level of heavy metals (HMs) in soil and underground water. These HMs may be accumulated in plants and animals when they take up such contaminated water, and then make their way into human food chains. Several remediation technologies have been employed to take up HMs. Diverse conventional means such as ion exchange, electrolytic technologies, and chemical extraction have been employed in the past, but the majority of these techniques are not economical for extensive projects and they need stringent control and continuous monitoring. These technologies also have low efficiency for effective removal of HMs. In this context, algae offer an eco-friendly and sustainable alternative for remediation of HMs from polluted water. The accumulation of HMs by macro and microalgae is advantageous for phycoremediation compared to other approaches that are not economical and not environmentally friendly. So, there is an urgent necessity to refine the chances of accumulation of HMs in algae, employing the techniques of genetic engineering to create transgenic species for over-expressing metallothioneins and phytochelatins, which may form complexes with HMs and store them in vacuoles to make the maximum use of phytoaccumulation while also removing hazardous metals from the aquatic habitats. This review outlines the major sources of HMs, their adverse effects on humans, the potential of algae in phytoremediation (called phycoremediation), and their uptake mechanism of HMs.

1. Introduction

Above and below groundwater contamination by HMs is a cause of grave concern [1]. Natural processes (e.g., floods, wind, and volcanic eruption) and anthropogenic activities (e.g., industrial effluents) are some of the ways that HMs are released into the environment [2,3,4,5,6,7]. Precipitation causes HMs in soil and air to become mixed in water bodies [8]. HMs are persistent and non-biodegradable in nature; they jeopardize human health and have negative effects on ecosystems [9,10,11]. Excess HMs in freshwater render it unfit for its designated purpose [12,13]. Hence, strict measures must be taken to reduce the level of HMs in wastewater below permitted limits prior to their discharge into natural reservoirs of water.
HMs are toxic to plants, animals, and overall ecosystem health [14,15]. HMs are accumulated in plants particularly through their roots [14] and enter the food chain. Increased lead (Pb) concentration in soils can decrease productivity, resulting in plants developing dark green leaves, decreased foliage, drooping of aged leaves, and dwarf brown roots [16,17,18,19]. Excess uptake of chromium (Cr), cadmium (Cd), lead (Pb), and manganese (Mn) can result in elevated rates of transpiration [20]. HMs elevate the amount of reactive oxygen species (ROS) that may harm fish and other aquatic life [21,22]. Humans are exposed to HMs mostly through accumulation of these metals at high concentrations, such as through ingestion of foods grown in such polluted soil [14].
Leaching, hydrolysis, polymer micro-encapsulation, chemical extraction, precipitation, ion exchange, and electrolytic technologies are a few conventional techniques used for the removal of HMs from polluted sites [23,24]. A major drawback is that these techniques are either expensive or ineffective for large-scale projects. They also require constant monitoring and tedious control. Hence, bioremediation (biological treatment) is suggested as an environmentally friendly substitute for the successful removal of HMs from polluted sites.
Algae are non-vascular plants that range in size from single-celled or colonial microalgae or “phytoplankton” such as diatoms and dinoflagllates, to large multicellular macroalgae such as seaweed or kelp [25]. Every alga contains chlorophyll but the majority of them lack stems, vascular tissue, roots, and leaves [25]. Bioremediation by algae, called “phycoremediation”, has emerged as a viable method for removing HMs from wastewater [26,27,28]. Phytoremediation is the process of using plants to sequester toxic chemicals from soil and water [29,30]. Advantages of using phytoremediation include: economic feasibility—phytoremediation is an autotrophic system driven by solar energy, hence their management is easy, and the maintenance and installation cost is low; eco-friendly—it can decrease the exposure of the contaminants to the ecosystem; applicability—it may be used over a vast field and may be disposed easily; it checks leaching and erosion of metal by stabilizing HMs, decreasing the risk of proliferation of contaminants; it may also improve the fertility of soil by releasing several organic matters into the soil [31,32,33]. Advantages of phycoremediation (Figure 1) include economic viability [34]; algal biomass may be used over a period of years [35]; algae have the potential to sequester diverse contaminants (e.g., HMs, pesticides, inorganic and organic toxins) [36]; in dilute effluents, they are quite effective, due to their high surface area-to-volume ratio [37]; they are highly efficient at sequestering HMs [36]; they are suitable even in wastewater containing higher metal concentrations [37]; algae are appropriate for aerobic and anaerobic effluent treatment units [38]; and cultures of microalgae may be cultivated in large-scale reservoirs, open ponds, and also in the laboratory [38].

Other Ecosystem Services of Algae

Algae play a crucial role in aquatic ecosystems by constituting the energy base of the food web for all the aquatic fauna. As autotrophic organisms, algae transform H2O and CO2 to sugar by the process of photosynthesis [39]. Photosynthesis also produces oxygen as a by-product, helping in the survival of fish and other aquatic fauna. Algae are significant components of various biogeochemical cycles and are important members of aquatic environment food webs [40]. Furthermore, algae also harm several goods and services of the ecosystem [41]. Hence, algae may be employed to signify ecosystem goods and services that complement how algal indicators are also used to measure the pollutant levels and habitat changes [42].
Algae affect the quality of water, assist in sewage treatment, and generate biomass [43]. They may be employed to produce hydrogen, which is a clean fuel, and biodiesel. Few other ecosystem services include decreasing the CO2 load of the atmosphere, restoring contaminated ecosystems, global warming mitigation, etc. [44,45,46]. Algae may be important in understanding and fixing some environmental issues [47,48].
Algae are known to be an excellent renewable energy source due to their fast rate of growth and their capability to be grown in wastewater or waste land [49]. Many government agencies and industries are making efforts to reduce the operating costs and capital cost, and to make algae-based fuel production economically feasible [50]. Algae are fast-growing plants that can produce more biomass or oil per acre than other plants and crops. Algal lipids would be one of the best feedstocks because they have the capacity to sequester enormous volumes of CO2, and do not compete heavily for agricultural land or food prices (i.e., the “food vs. fuel” concern is not relevant). Biofuels, which principally include biodiesel, biogas, and bioethanol, have come a long way and are likely to become important sources of sustainable energy in the future [51].
The buildup of HMs by algae offers various advantages for phytoremediation over other, less cost-effective and environmentally friendly approaches. This review looks at how algal biomass may help remove HMs from water and wastewater. High HMs ion uptake capacity and low-cost cultivation, with appropriate environmental conditions such as time, pH, temperature, and contact, make algal biomass very useful for wastewater bioremediation. The review’s purpose is to offer an overview of HMs as a potential contaminant in aquatic ecosystems, the phytoremediation potential of algae, and the underlying mechanism of HMs accumulation.

2. Heavy Metal as a Potential Contaminant of Aquatic Ecosystems

Aquatic ecosystem contamination with HMs has escalated since the inception of the industrial revolution. In contrast to organic chemicals, most of HMs cannot breakdown into less toxic or non-toxic compounds [52]; hence, they may persist in the aquatic ecosystems. HMs present in bioavailable form may be accumulated by aquatic life and, in due course, move up the trophic levels [52]. Upon entering the biological systems, HMs may interfere with development, growth, and metabolism, leading to several chronic and severe effects [53]. Few HMs are nutritionally important at lower levels, e.g., copper (Cu), zinc (Zn), and iron (Fe), whereas others do not contribute to biological functions, e.g., Pb, mercury (Hg). Nevertheless, surplus exposure of HMs may lead to toxic effects (Figure 2) [54]. Hence, to ameliorate human and ecological health risk, efforts for remediation of HMs pollution need large-scale implementation.
The majority of inorganic pollutants comprise highly toxic metals (Pb, Cd, Zn, Hg, etc.), whereas few are non-metallic compounds of inorganic nature such as phosphates and nitrates. However, these are still damaging to the aquatic habitat and somewhat toxic to the biological system [12]. HMs have been known to attack cell organelles and components such as mitochondria, cell membrane, nuclei, lysosome, and endoplasmic reticulum, and a few enzymes that take part in damage repair, metabolism, and detoxification in biological systems [58]. Metallic ions can be associated with cellular nucleic acids and can cause DNA damage, which may cause modulation of the cell cycle, apoptosis, or carcinogenesis [10,15,59,60]. Previous research discovered that the generation of reactive oxygen species (ROS) and stress due to oxidation plays a central role in the carcinogenicity and toxicity of metals such as Cd [61], Cr [62], As [63,64], Hg [65], and Pb [63,66]. Due to their potency of toxicity, Cd, Cr, As, Hg, and Pb, rank among the most significant metals in terms of public health. These are systemic toxins that have been shown to cause various harm to organs, even at low doses. The United States Environmental Protection Agency (U.S. EPA) and the International Agency for Research on Cancer (IARC) have classified these metals as either “probable” or “known” carcinogens to humans [53].

3. Major Sources and Adverse Effects of Heavy Metals

The nature of HMs varies extensively in terms of their chemical properties, and HMs are used widely in machines, electronics, and high-tech applications. As a result, they find their way into the food chains of animals and humans from a wide range of man-made sources and from natural sources such as the geochemical weathering of rocks and soil [67]. Primary sources of contamination are landfill leachates, mining wastes, industrial wastewaters, municipal wastewater, and urban runoff, especially from the electronic, metal-finishing, and electroplating industries [68]. Many water bodies have concentrations of metal that are higher than the standards established to protect humans and other organisms. Metals can be transported with sediments and thus may bioaccumulate in the food chain.
Major sources of HMs that involve anthropogenic activities are waste disposal, agriculture, industry, etc. [69]. Nonpoint sources of pollution such as the unregulated release and disposal of domestic and industrial wastewater adversely affects organisms in aquatic ecosystems [70]. Even non-significant concentrations of HMs pose noticeable problems in fauna and flora [71]. HMs do not participate in the metabolism, but they are accumulated via several mechanisms that include bioconcentration, biomagnifications, and bioaccumulation [27]. Organisms exposed to HMs have a protective mechanism against HMs toxicity. Following a threshold level, the HMs content causes direct toxicity to aquatic wildlife and vegetation [72].
In aquatic environments, HMs are one of the most frequent contaminants. These metals give rise to a toxicity threat to animals and human beings, even at very low concentrations, and cause severe impacts at higher levels (Table 1). Lead is very toxic and affects the reproductive system, kidneys, and nervous system [73]. Exposure to Pb leads to encephalopathic symptoms and irreversible brain damage. Cd exposure leads to renal problems, liver damage, bone degeneration, and blood-related problems [74]. It has been reported that there is enough evidence that proves the carcinogenicity of Cd. High-level exposure to Cu dust causes irritation in the mouth, eyes, and nose, and can cause diarrhea and nausea. Continuous exposure can cause damage the kidney and even death. Cu is harmful to a variety of aquatic organisms, even in trace amounts [75]. Accumulation of HMs inside the plant cell leads to inhibition of photosynthesis, decreases algal growth, and increases the permeability of the plasmalemma, which leads to loss of cell solutes, breakdown of the of the cell membrane, enzyme inhibition, unusual morphological development, and loss of flagella in certain algae [76].

4. Phytoremediation Potential of Algae

Algae have similar photosynthetic pigments as those of higher plants but algae have greater photosynthetic efficiency, and algae release higher amounts of oxygen into water bodies, leading to aerobic breakdown of organic compounds [77]. Algae utilize waste as a source of nutrition and decrease pollutants by enzymatic and metabolic processes. HMs pollution and xenobiotic compounds may be volatilized, detoxified, and transformed by the algal metabolic pathways [78].
Mechanisms for removal of HMs include flocculation, absorption, sedimentation, precipitation, cation and anion exchange, microbiological activity, oxidation/reduction, and uptake and complexation [79]. Removal of HMs by microalgae is performed directly via two major mechanisms; one is metabolism, which is dependent on the uptake of HMs in cells at low concentrations, and the other is biosorption, which is a passive process of adsorption [80].
Phytoremediation is the process of remediating soil and aquatic systems with the help of plants, algae, or fungi to absorb HMs. Lately, the application of macro- and microalgae has drawn significant attention because of their capability to sequester toxic elements from the environment [81].
Algae possess various characteristics that make them excellent candidates for selective HMs removal, including include the ability to grow heterotrophically and autotrophically, high surface area/volume ratios, tolerance to HMs, phototaxy, and genetic manipulation potential and phytochelatin expression [82]. Cladophora and Enteromorpha have been used to monitor levels of HMs [83]. Macroalgae have been used widely monitor HMs contamination in marine environments.
The propensity of macroalgae to collect metals inside their tissues has led to their widespread usage as metal availability bioindicators [84,85,86]. Cyanophyta and Chlorophyta are hyper-accumulators of As and B [87], absorbing these metals from the contaminated ecosystems. These algae may be efficient phycoremediators and their presence in water may decrease As and B [87]. The Phaeophyta are promising accumulators of metals because of the presence of significant levels of alginates and sulphated polysaccharides inside their cell walls, for which metals have high affinity [88]. Nielsen et al. [89] observed that brown algae such as Fucus spp. sometimes dominate the vegetation of HM-contaminated habitats.
Microalgae are microscopic, aquatic organisms that thrive in both marine and freshwater environments, and have a mechanism of photosynthesis that is similar to that of land plants. Biomass-wise, they constitute the biggest group of primary producers, accountable for almost 50% of global photosynthesis [90]. They have molecular processes that distinguish between essential and unnecessary HMs for growth [91]. Researchers have widely focused on the benefits of using microalgae in the biosorption of metals. The advantages include: fast capability to uptake metal (compared to higher plants), saves time and money, environmentally friendly, ready availability, perpetual occurrence, reusable/recyclable, efficient, high surface-to-volume ratio, capability to bind as much as 10% of their biomass with increased selectivity (which increases their performance), no generation of toxic waste, advantageous in both continuous and batch systems, and employable to waters having variable concentrations of metals and other contaminants [92]. Monteiro et al. [92] supported the importance of microalgae in removing metals, even at trace amounts of contamination. Microalgae not only have a better efficacy for HMs remediation, but they also help with HMs recovery using only some basic desorption compounds. They can remediate HMs from multi-metal solutions [93,94]. Rajamani et al. [94] described the transgenic methods employed to increase the HMs binding capability of microalgae, including the application of fluorescent HMs biosensors devised by employing transgenic Chlamydomonas. Microalgal affinity for polyvalent metals aids in determining its potential application in purifying wastewater having dissolved metallic ions [95]. Chlorella and Scenedesmus are the microalgae of choice for metal removal. Brinza et al. [96] studied the ability of Chlorella salina, Chlamydomonas reinhardtii, C. vulgaris, Chlorella sorokiniana, Chlorella miniata, Scenedesmus quadricauda, Scenedesmus abundans, Spirogyra spp., Scenedesmus subspicatus, Spirulina platensis, Stigeoclonium tenue, Stichococcus bacillaris, and Porphyridium purpureum to absorb metals such as Al, Ca, Mg, Fe, Co, Cu, Sr, Pb, Mn, V, Ni, Cd, Se, Zn, Mo As, and K. [96]. Perales-Vela et al. [91] studied Cu, Cd, and U removal capabilities of P. tricornutum, Scenedesmus and Chlorella. They found that microalgae synthesize peptides can bind to HMs [91], resulting in the formation of organometallic complexes that are further positioned within the vacuoles to make it easier to manage the cytoplasmic concentration of HMs ions, nullifying the HMs’ potential harmful effects [97].
The marine alga Dunaliella tertioleca exhibited high content of phytochelatin attributed to its ability to hyperaccumulate Cd and Zn [98]. Rhodophyta, Chromophyta, and Chlorophyta have strong cytochrome P450 monooxygenase activity and are efficient against xenobiotic (foreign) substrates such as 2,3-dichlorobiphenyl, isoproturon and 3,4-chlorobiphenyl [99,100]. The action of Nitroreductases on nitroaromatic compounds such as Glycerol trinitrate (GTN), 2,4,6-trinitro toluene (TNT), and hexahydro-1,3,5-trinitro1,3,5-triazine (RDX) in some was indicated by several researchers who successfully used transgenic approaches to produce end-products such as nitrate, ammonium, and CO2 [100,101]. Scientists discovered an mt gene in the marine brown alga Fucus vesiculosus that was activated with Cu exposure, and the proteins produced efficiently bound to Cu and Cd [102]. Several species of algae have been extensively studied and found to bear excellent potential to remove HMs from contaminated water (Table 2).
Mei et al. [128] observed that Platymonas subcordiformis, a marine green microalga, is a hyperaccumulator of Sr (strontium). However, high Sr concentration leads to oxidative damage, as suggested by the surge in lipid peroxidation in the cell samples of algae and the decline in the chlorophyll contents and rate of growth. Few species of algae can convert phenylmercuric or mercuric ions into metallic mercury, which is then volatilized out of the cell and from the solution [129,130]. Phormidium, a blue-green alga (which are actually procaryotes), is a promising hyperaccumulator of HMs such as Zn, Cd, Pb, Cu, and Ni [131,132]. Caulerpa racemosa is a cheap biomaterial that could be used for removing boron (B) from contaminated water bodies [133,134]. Dunaliella salina (green microalgae) is a promising accumulator of Zn followed by Cu and Co, and the lowest accumulation was for Cd. This can be due to the significance of Zn as a hydrogen transporter in the process of photosynthesis [134]. Metallic cations form complexes with carboxyl and other functional groups in algae. Further research may reveal that some algal species have adsorption capacities for different HMs [135].

5. Mechanism of Phytoremediation by Algae

Many researchers have documented the capability of phytoplankton to take up HMs from water bodies [136,137,138,139]. Removal of HMs ions by microalgae occurs through biosorption and bioaccumulation.

5.1. Biosorption

In biosorption, HMs ions adhere to functional groups on the cell surface (Figure 3) [140,141]. Algal cell wall components such as fucoidan and alginate are the most important functional groups accountable for HMs biosorption [142,143]. HMs ions in wastewater are exchanged with elemental ions such as Na+, Ca2+, and K+ on the surface of algae. The feasibility of this process is contingent upon critical factors such as regeneration potential and metal selectivity. Chemical modification of the algal biomass, such as oxidation with potassium permanganate or crosslinking with epichlorohydrin, improves biosorption selectivity [144].
Biosorption is a physio-chemical property that results in removal of HMs by covalent or ionic bonding [145,146]. Various functional groups, such as SH, COO, RNH2, RS, RO, and OH, facilitate metal ion biosorption. These groups can be present on the cell surface and in the cytoplasm, particularly in vacuoles. Algal cell walls have a net negative charge because of the presence of PO43−, COO, and other functional groups that bond with metals through ion exchange. Ditylum brightwellii and other species of algae produce a special material known as Cu-ligands [146,147]. Carboxyl (COO) present in brown algal cell walls is the most profuse acidic functional group. Algae minimize damage induced by metals through exclusion and excretion of metals and by synthesizing proteins and other binding compounds such as glutathione (GSH) and metallothioneins (MTs) [147,148]. Other functional groups are found in algae with different cell wall compositions. The selectivity of HMs uptake is contingent upon the encapsulation of microalgae and its cellulose derivatives [147,149]. Desorption of the adsorbed HMs can occur at a lower pH of suspension. Reversible loading or unloading of adsorbed HMs is achievable by utilizing citric acid or HCl sorption. Experiments on biosorption of metals have been conducted with brown algae (e.g., Fucus vesiculosus and Laminaria japonica), blue-green algae (e.g., Oscillatoria sp. and Microcystis aeruginosa), and freshwater green microalgae (e.g., Scenedesmus sp., Chlorella sp., and Chlamydomonas sp.) [150]. Several technologies for the removal of HMs, including algal turf scrubbers and high-rate algal ponds, have been backed for pragmatic usage. These technologies, however, are currently insufficient for large-scale implementation. Success of phycoremediation is dependent primarily on the bioaccumulation and biosorption capabilities of the algae, with biosorption controlling bioremediation [151]. Algae are economical and efficient biosorbents because of their low nutrient requirements. Algal biosorption effectiveness is around 15.3–84.6% higher than that of other fungi and bacteria [142,152,153].
The ability of algae to metabolize and adsorb HMs is linked with their high surface-to-volume ratios, efficient metal uptake, storage systems, and the presence of high-affinity metal-binding groups on their cell surfaces [27]. In physical adsorption, metal ions in aqueous solution bind to polyelectrolytes on the algal cell wall through weak forces such as redox interaction, covalent bonding, biomineralization, and van der Waals forces [154]. The pH of the adsorbing media strongly influences adsorption of metal ion. By replacing the functional groups on metal cations, alkaline pH boosts their attraction and eventually increases their adsorption on cell surfaces that carry a negative charge, such as phosphate, polysaccharides, aminos, carboxyl groups of protein, and amino groups of nucleic acid [75].

5.2. Bioaccumulation

By the process of bioaccumulation, HMs ions are transported across cell membranes through passive and active transport systems and accumulate inside the cells. Extracellular and intracellular metal binding approaches (such as complexation, physical adsorption, ion exchanges and chelation) have been employed by algae to lessen toxicity caused by HMs [155].
These mechanisms efficiently transform toxic metals into less or non-toxic forms [153,155]. Detoxification of metal by algae is performed in various ways such as binding to a particular intracellular organelle, transport to particular cellular components such as polyphosphate vacuoles/bodies, flushing out into the solution by efflux pump, and with synthesis of class III metallothioneins or phytochelatins [91,156]. Phytochelatins are peptides that bind to metals and are small, having molecular weights in the range from 2 to 10 kDa. Phytochelatins are synthesized by the constitutive enzyme called phytochelatin synthase [157]. Their synthesis from g-glutamylcysteine, hydroxymethyl-glutathione, homo-glutathione or glutathione [158] is catalyzed by a trans-peptidase known as phytochelatin synthase (the constitutive enzyme) that needs post-translational activation by HMs [159,160]. All higher plants and most algae can synthesize phytochelatins [161]. Cobbett and Goldsbrough (2002) conducted kinetic studies demonstrating that the synthesis of phytochelatin is quick (within a few minutes) and is not dependent on de novo synthesis of protein [162]. A diverse range of metalloids and metals such as As, Cu, Cd, Pb, Ag, Sn, Hg, Zn, and Au may help in activating phytochelatin synthase both in vivo and in vitro [163]. Cd was found to be a potential activator of phytochelatin synthesis and was able to promote the synthesis of phytochelatins with more stable chains (i.e., up to PC5) [163,164].
Precipitation in sulfide, phosphate, or carbonate reduces HMs toxicity on living algae. Cladophora glomerata, a green alga, was capable of removing Pb, Cd, Ni, Cr, and V at 7.9, 0.1, 15.6, 1.7, and 37.7 mg kg−1, respectively, from a refinery sewage lagoon [147]. Fucus vesiculosus, a brown macroalga, exhibited a significant efficiency for HMs accumulation from contaminated saltwater, removing 65, 95, and 76% of Pb, Hg, and Cd, respectively. Bioconcentration factors for Cd, Pb, and Hg ranged from 600 to 2300, with complete metal removal from the solution [165].

6. Conclusions

HMs contamination of aquatic ecosystems is a serious concern because of its toxicity towards human health and all biota. Several species of algae have been identified as promising candidates for removal or detoxification of HMs and are potential economical substitutes to physicochemical techniques of remediation. Removal of HMs can be attained by bioaccumulation and biosorption. Microalgae, predominantly, possess several mechanisms for sequestering HMs ions and are therefore recognized as potential biosorbents. Many reports corroborate their supremacy over several other physiochemical and traditional methods and their usefulness in large-scale remediation of wastewater. Although algae have excellent potential to be used for HMs accumulation, they have some limitations such as low biomass, and sensitivity towards high concentration of HMs. These problems may be solved by some advanced interventions such as genetic modifications that enhance the biomass production. In addition, in the field, a suitable and sustainable method for selecting the most acceptable biosorbents and favorable physical circumstances, and identifying the primary problems, must be developed. However, it is critical to consider a variety of microalgal remediation methods as environmentally acceptable alternatives for a healthier environment. Genetic engineering can be used to enhance the creation of transgenic organisms that over-express phytochelatins and metallothioneins that can form complexes with HMs and translocate them into vacuoles, and aid phytoaccumulation and hazardous element removal from aquatic habitats. Microalgae strains show varied response and tolerance along with bioaccumulation capability towards HMs. Various functional groups and proteins, in addition to peptides, are accountable for metal-binding. Mechanisms such as volatilization, reduction, extracellular adsorption, complex formation, intracellular accumulation, bio-methylation, and ion exchange chelation are involved in bioremediation and biosorption of HMs. More efforts in the areas of immobilization techniques, genetic engineering, and integration, along with other techniques, are required to entirely explore the potential of microalgae in HMs phycoremediation and, simultaneously, the formation of value-added products.

Author Contributions

A. drafted the manuscript. K.B. and J.K. conceptualized the theme and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

Ankit is thankful to University Grants Commission, New Delhi, India for the award and financial assistance in form of Senior Research Fellowship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kobielska, P.A.; Howarth, A.J.; Farha, O.K.; Nayak, S. Metal-organic frameworks for heavy metal removal from water. Coord. Chem. Rev. 2018, 358, 92–107. [Google Scholar] [CrossRef]
  2. Gupta, A.; Joia, J.; Sood, A.; Sood, R.; Sidhu, C. Microbes as potential tool for remediation of heavy metals: A review. J. Microb. Biochem. Technol. 2016, 8, 364–372. [Google Scholar] [CrossRef] [Green Version]
  3. Masindi, V.; Muedi, K.L. Environmental contamination by heavy metals. Heavy Met. 2018, 10, 115–132. [Google Scholar]
  4. Schwartz, G.E.; Hower, J.C.; Phillips, A.L.; Rivera, N.; Vengosh, A.; Hsu-Kim, H. Ranking coal ash materials for their potential to leach arsenic and selenium: Relative importance of ash chemistry and site biogeochemistry. Environ. Eng. Sci. 2018, 35, 728–738. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, Y.; Duan, X.; Wang, L. Spatial distribution and source analysis of heavy metals in soils influenced by industrial enterprise distribution: Case study in Jiangsu Province. Sci. Total Environ. 2020, 710, 134953. [Google Scholar] [CrossRef]
  6. Wu, W.; Qu, S.; Nel, W.; Ji, J. The impact of natural weathering and mining on heavy metal accumulation in the karst areas of the Pearl River Basin, China. Sci. Total Environ. 2020, 734, 139480. [Google Scholar] [CrossRef]
  7. Wei, J.; Zheng, X.; Liu, J.; Zhang, G.; Zhang, Y.; Wang, C.; Liu, Y. The Levels, Sources, and Spatial Distribution of Heavy Metals in Soils from the Drinking Water Sources of Beijing, China. Sustainability 2021, 13, 3719. [Google Scholar] [CrossRef]
  8. Warmate, A.; Ideriah, T.; Tamunobereton ARI, I.T.; Inyang, U.U.; Ibaraye, T. Concentrations of heavy metals in soil and water receiving used engine oil in Port Harcourt, Nigeria. J. Ecol. Nat. Environ. 2011, 3, 54–57. [Google Scholar]
  9. Kwaansa-Ansah, E.E.; Nti, S.O.; Opoku, F. Heavy metals concentration and human health risk assessment in seven commercial fish species from Asafo Market, Ghana. Food Sci. Biotechnol. 2019, 28, 569–579. [Google Scholar] [CrossRef]
  10. Chen, X.; Wang, Y.; Wang, X.; Wang, M.; Liang, Y.; Zhu, G.; Jin, T. A nomogram for predicting the renal dysfunction in a Chinese population with reduction in cadmium exposure based on 8 years follow up study. Ecotoxicol. Environ. Saf. 2020, 191, 110251. [Google Scholar] [CrossRef]
  11. Huang, J.; Peng, S.; Mao, X.; Li, F.; Guo, S.; Shi, L.; Shi, Y.; Yu, H.; Zeng, G. Source apportionment and spatial and quantitative ecological risk assessment of heavy metals in soils from a typical Chinese agricultural county. Process Saf. Environ. Prot. 2019, 126, 339–347. [Google Scholar] [CrossRef]
  12. Dixit, S.; Singh, D.P. Phycoremediation: Future perspective of green technology. In Algae and Environmental Sustainability; Springer: New Delhi, India, 2015; pp. 9–21. [Google Scholar]
  13. Schuler, M.S.; Relyea, R.A. A review of the combined threats of road salts and heavy metals to freshwater systems. BioScience 2018, 68, 327–335. [Google Scholar] [CrossRef]
  14. Jordao, C.P.; Nascentes, C.C.; Cecon, P.R.; Fontes, R.L.F.; Pereira, J.L. Heavy metal availability in soil amended with composted urban solid wastes. Environ. Monit. Assess. 2006, 112, 309–326. [Google Scholar] [CrossRef] [PubMed]
  15. Balali-Mood, M.; Naseri, K.; Tahergorabi, Z.; Khazdair, M.R.; Sadeghi, M. Toxic Mechanisms of Five Heavy Metals: Mercury, Lead, Chromium, Cadmium, and Arsenic. Front. Pharmacol. 2021, 12, 643972. [Google Scholar] [CrossRef]
  16. Bhattacharyya, P.; Chakrabarti, K.; Chakraborty, A.; Tripathy, S.; Powell, M.A. Fractionation and bioavailability of Pb in municipal solid waste compost and Pb uptake by rice straw and grain under submerged condition in amended soil. J. Geosci. 2008, 12, 41–45. [Google Scholar] [CrossRef]
  17. Zulfiqar, U.; Farooq, M.; Hussain, S.; Maqsood, M.; Hussaind, M.; Ishfaqa, M.; Ahmada, M.; Anjumf, M.Z. Lead toxicity in plants: Impacts and remediation. J. Environ. Manag. 2019, 250, 109557. [Google Scholar] [CrossRef]
  18. Kohli, S.K.; Handa, N.; Bali, S.; Khanna, K.; Arora, S.; Sharma, A.; Bhardwaj, R. Current Scenario of Pb Toxicity in Plants: Unraveling Plethora of Physiological Responses. Rev. Environ. Contam. Toxicol. 2020, 249, 153–197. [Google Scholar]
  19. Giannakoula, A.; Therios, I.; Chatzissavvidis, C. Effect of Lead and Copper on Photosynthetic Apparatus in Citrus (Citrus aurantium L.) Plants. The Role of Antioxidants in Oxidative Damage as a Response to Heavy Metal Stress. Plants 2021, 10, 155. [Google Scholar] [CrossRef]
  20. Sharma, R.K.; Agrawal, M.; Marshall, F. Heavy metal contamination of soil and vegetables in suburban areas of Varanasi, India. Ecotoxicol. Environ. Saf. 2007, 66, 258–266. [Google Scholar] [CrossRef]
  21. Schützendübel, A.; Nikolova, P.; Rudolf, C.; Polle, A. Cadmium and H2O2 -induced oxidative stress in Populus × canescens roots. Plant Physiol. Biochem. 2002, 40, 577–584. [Google Scholar] [CrossRef]
  22. Woo, S.; Yum, S.; Park, H.S.; Lee, T.K.; Ryu, J.C. Effects of heavy metals on antioxidants and stress-responsive gene expression in Javanese medaka (Oryzias javanicus). CBP 2009, 149, 289–299. [Google Scholar] [CrossRef] [PubMed]
  23. Jais, N.M.; Mohamed, R.; Al-Gheethi, A.; Hashim, M.A. The dual roles of phycoremediation of wet market wastewater for nutrients and heavy metals removal and microalgae biomass production. Clean Technol. Environ. Policy 2017, 19, 37–52. [Google Scholar] [CrossRef]
  24. Kumar, V.; Nanda, M. Microalgae: A Promising Tool for Remediation of Heavy Metals. In Biostimulation Remediation Technologies for Groundwater Contaminants; IGI Global: Hershey, PA, USA, 2018; pp. 141–153. [Google Scholar]
  25. Anbuchezhian, R.; Karuppiah, V.; Li, Z. Prospect of marine algae for production of industrially important chemicals. In Algal Biorefinery: An Integrated Approach; Springer: Cham, Switzerland, 2015; pp. 195–217. [Google Scholar]
  26. Koul, B.; Sharma, K.; Shah, M.P. Phycoremediation: A sustainable alternative in wastewater treatment (WWT) regime. Environ. Technol. Innov. 2022, 25, 102040. [Google Scholar] [CrossRef]
  27. Ahmad, S.; Pandey, A.; Pathak, V.V.; Tyagi, V.V.; Kothari, R. Phycoremediation: Algae as eco-friendly tools for the removal of heavy metals from wastewaters. In Bioremediation of Industrial Waste for Environmental Safety; Springer: Berlin/Heidelberg, Germany, 2020; pp. 53–76. [Google Scholar]
  28. Poo, K.M.; Son, E.B.; Chang, J.S.; Ren, X.; Choi, Y.J.; Chae, K.J. Biochars derived from wasted marine macro-algae (Saccharina japonica and Sargassum fusiforme) and their potential for heavy metal removal in aqueous solution. J. Environ. Manag. 2018, 206, 364–372. [Google Scholar] [CrossRef]
  29. 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]
  30. Tiwari, J.; Ankit; Shweta; Kumar, S.; Korstad, J.; Bauddh, K. Ecorestoration of polluted aquatic ecosystems through rhizofiltration. In Phytomanagement of Polluted Sites; Elsevier: Amsterdam, The Netherlands, 2019; pp. 179–201. [Google Scholar]
  31. Aken, B.V.; Correa, P.A.; Schnoor, J.L. Phytoremediation of polychlorinated biphenyls: New trends and promises. Environ. Sci. Technol. 2009, 44, 2767–2776. [Google Scholar] [CrossRef] [Green Version]
  32. Wuana, R.A.; Okieimen, F.E. Heavy metals in contaminated soils: A review of sources, chemistry, risks and best available strategies for remediation. Int. Sch. Res. Not. 2011, 2011, 402647. [Google Scholar] [CrossRef] [Green Version]
  33. Jacob, J.M.; Karthik, C.; Saratale, R.G.; Kumar, S.S.; Prabakar, D.; Kadirvelu, K. Biological approaches to tackle heavy metal pollution: A survey of literature. J. Environ. Manag. 2018, 217, 56–70. [Google Scholar] [CrossRef]
  34. Kotrba, P. Microbial biosorption of metals—general introduction. In Microbial Biosorption of Metals; Kotrba, P., Mackova, M., Macek, T., Eds.; Springer: Dordrecht, The Netherlands, 2011; pp. 1–6. [Google Scholar]
  35. Darda, S.; Papalas, T.; Zabaniotou, A. Biofuels journey in Europe: Currently the way to low carbon economy sustainability is still a challenge. J. Clean. Prod. 2019, 208, 575–588. [Google Scholar] [CrossRef]
  36. Ajayan, K.V.; Selvaraju, M.; Thirugnanamoorthy, K. Growth and heavy metals accumulation potential of microalgae grown in sewage wastewater and petrochemical effluents. Pak. J. Biol. Sci. 2011, 14, 805–811. [Google Scholar] [CrossRef] [Green Version]
  37. Yu, Q.; Matheickal, J.T.; Yin, P.; Kaewsarn, P. Heavy metal uptake capacities of common marine macro algal biomass. Water Res. 1999, 33, 1534–1537. [Google Scholar] [CrossRef]
  38. Salama, E.S.; Kurade, M.B.; Abou-Shanab, R.A.; El-Dalatony, M.M.; Yang, I.-S.; Min, B.; Jeon, B.-H. Recent progress in microalgal biomass production coupled with wastewater treatment for biofuel generation. Renew. Sustain. Energy Rev. 2017, 79, 1189–1211. [Google Scholar] [CrossRef]
  39. Xia, A.; Herrmann, C.; Murphy, J.D. How do we optimize third-generation algal biofuels? Biofuels Bioprod. Biorefin. 2015, 9, 358–367. [Google Scholar] [CrossRef]
  40. Villar-Argaiz, M.; Medina-Sánchez, J.M.; Biddanda, B.A.; Carrillo, P. Predominant non-additive effects of multiple stressors on autotroph C: N: P ratios propagate in freshwater and marine food webs. Front. Microbiol. 2018, 9, 69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Tsikoti, C.; Genitsaris, S. Review of Harmful Algal Blooms in the Coastal Mediterranean Sea, with a Focus on Greek Waters. Diversity 2021, 13, 396. [Google Scholar] [CrossRef]
  42. Stevenson, J. Ecological assessments with algae: A review and synthesis. J. Phycol. 2014, 50, 437–461. [Google Scholar] [CrossRef]
  43. Singh, G.; Patidar, S.K. Development and applications of attached growth system for microalgae biomass production. BioEnergy Res. 2021, 14, 709–722. [Google Scholar] [CrossRef]
  44. Ahmad, A.; Banat, F.; Alsafar, H.; Hasan, S.W. Algae biotechnology for industrial wastewater treatment, bioenergy production, and high-value bioproducts. Sci. Total Environ. 2022, 806, 150585. [Google Scholar] [CrossRef]
  45. Mustafa, S.; Bhatti, H.N.; Maqbool, M.; Iqbal, M. Microalgae biosorption, bioaccumulation and biodegradation efficiency for the remediation of wastewater and carbon dioxide mitigation: Prospects, challenges and opportunities. J. Water Process. Eng. 2021, 41, 102009. [Google Scholar] [CrossRef]
  46. Malyan, S.K.; Bhatia, A.; Tomer, R.; Harit, R.C.; Jain, N.; Bhowmik, A.; Kaushik, R. Mitigation of yield-scaled greenhouse gas emissions from irrigated rice through Azolla, Blue-green algae, and plant growth–promoting bacteria. Environ. Sci. Pollut. Res. 2021, 28, 51425–51439. [Google Scholar] [CrossRef]
  47. Ramanan, R.; Kim, B.H.; Cho, D.H.; Oh, H.M.; Kim, H.S. Algae–bacteria interactions: Evolution, ecology and emerging applications. Biotechnol. Adv. 2016, 34, 14–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Rai, L.C.; Har, D.K.; Frieder, H.M.; Carl, J.S. Services of algae to the environment. J. Microbiol. Biotechnol. 2000, 10, 119–136. [Google Scholar]
  49. Alam, F.; Mobin, S.; Chowdhury, H. Third generation biofuel from algae. Procedia Eng. 2015, 105, 763–768. [Google Scholar] [CrossRef]
  50. Ullah, K.; Ahmad, M.; Sharma, V.K.; Lu, P.; Harvey, A.; Zafar, M.; Sultana, S. Assessing the potential of algal biomass opportunities for bioenergy industry: A review. Fuel 2015, 143, 414–423. [Google Scholar] [CrossRef]
  51. Energy, J.O.R. Retracted: Microalgae as a Renewable Source of Energy: A Niche Opportunity. J. Renew. Energy 2021, 2021, 9813285. [Google Scholar] [CrossRef]
  52. Gheorghe, S.; Stoica, C.; Vasile, G.G.; Nita-Lazar, M.; Stanescu, E.; Lucaciu, I.E. Metals toxic effects in aquatic ecosystems: Modulators of water quality. In Water Quality; Tutu, H., Ed.; InTech: Rijeka, Croatia, 2017; pp. 60–89. [Google Scholar]
  53. Tchounwou, P.B.; Yedjou, C.G.; Patlolla, A.K.; Sutton, D.J. Heavy metal toxicity and the environment. Exp. Suppl. 2012, 101, 133–164. [Google Scholar]
  54. Sarwar, N.; Imran, M.; Shaheen, M.R.; Ishaque, W.; Kamran, M.A.; Matloob, A.; Rehim, A.; Hussain, S. Phytoremediation strategies for soils contaminated with heavy metals: Modifications and future perspectives. Chemosphere 2017, 171, 710–721. [Google Scholar] [CrossRef]
  55. Briffa, J.; Sinagra, E.; Blundell, R. Heavy metal pollution in the environment and their toxicological effects on humans. Heliyon 2020, 6, 04691. [Google Scholar] [CrossRef]
  56. Heckathorn, S.A.; Mueller, J.K.; LaGuidice, S.; Zhu, B.; Barrett, T.; Blair, B.; Dong, Y. Chloroplast small heat-shock proteins protect photosynthesis during heavy metal stress. Am. J. Bot. 2004, 91, 1312–1318. [Google Scholar] [CrossRef]
  57. Le Gall, H.; Philippe, F.; Domon, J.M.; Gillet, F.; Pelloux, J.; Rayon, C. Cell wall metabolism in response to abiotic stress. Plants 2015, 4, 112–166. [Google Scholar] [CrossRef]
  58. Wang, S.; Shi, X. Molecular mechanisms of metal toxicity and carcinogenesis. Mol. Cell. Biochem. 2001, 222, 3–9. [Google Scholar] [CrossRef] [PubMed]
  59. Kasprzak, K.S. Oxidative DNA and protein damage in metal induced toxicity and carcinogenesis. Free Radic. Biol. Med. 2002, 32, 958–967. [Google Scholar] [CrossRef]
  60. Beyersmann, D.; Hartwig, A. Carcinogenic metal compounds: Recent insight into molecular and cellular mechanisms. Arch. Toxicol. 2008, 82, 493–512. [Google Scholar] [CrossRef] [PubMed]
  61. Tchounwou, P.B.; Ishaque, A.B.; Schneider, J. Cytotoxicity and transcriptional activation of stress genes in human liver carcinoma cells (HepG2) exposed to cadmium chloride. Mol. Cell. Biochem. 2001, 222, 21–28. [Google Scholar] [CrossRef]
  62. Patlolla, A.; Barnes, C.; Yedjou, C.; Velma, V.; Tchounwou, P.B. Oxidative stress, DNA damage and antioxidant enzyme activity induced by hexavalent chromium in Sprague Dawley rats. Environ. Toxicol. 2009, 24, 66–73. [Google Scholar] [CrossRef] [Green Version]
  63. Tchounwou, P.B.; Centeno, J.A.; Patlolla, A.K. Arsenic toxicity, mutagenesis and carcinogenesis—A health risk assessment and management approach. Mol. Cell. Biochem. 2004, 255, 47–55. [Google Scholar] [CrossRef]
  64. Yedjou, C.G.; Tchounwou, P.B. Oxidative stress in human leukemia cells (HL-60), human liver carcinoma cells (HepG2) and human Jerkat-T cells exposed to arsenic trioxide. Met. Ions Biol. Med. 2006, 9, 298–303. [Google Scholar]
  65. Sutton, D.J.; Tchounwou, P.B. Mercury induces the externalization of phosphatidylserine in human proximal tubule (HK-2) cells. Int. J. Environ. Res. Public Health 2007, 4, 138–144. [Google Scholar] [CrossRef] [Green Version]
  66. Tchounwou, P.B.; Yedjou, C.G.; Foxx, D.; Ishaque, A.; Shen, E. Lead induced cytotoxicity and transcriptional activation of stress genes in human liver carcinoma cells (HepG2). Mol. Cell. Biochem. 2004, 255, 161–170. [Google Scholar] [CrossRef]
  67. Gautam, R.K.; Sharma, S.K.; Mahiya, S.; Chattopadhyaya, M.C. Contamination of heavy metals in aquatic media: Transport, toxicity and technologies for remediation. In Heavy Metals in Water: Presence, Removal and Safety; RSC Publishing: London, UK, 2014; pp. 1–24. [Google Scholar]
  68. Chowdhary, P.; Hare, V.; Raj, A. Book review: Environmental pollutants and their bioremediation approaches. Front. Bioeng. Biotechnol. 2018, 6, 193. [Google Scholar] [CrossRef]
  69. Singh, R.L.; Singh, P.K. Global environmental problems. In Principles and Applications of Environmental Biotechnology for a Sustainable Future; Springer: Singapore, 2017; pp. 13–41. [Google Scholar]
  70. Khan, I.; Ali, M.; Aftab, M.; Shakir, S.; Qayyum, S.; Haleem, K.S.; Tauseef, I. Mycoremediation: A treatment for heavy metal-polluted soil using indigenous metallotolerant fungi. Environ. Monit. Assess. 2019, 191, 622. [Google Scholar] [CrossRef] [PubMed]
  71. Asati, A.; Pichhode, M.; Nikhil, K. Effect of heavy metals on plants: An overview. Int. J. Appl. Innov. Eng. Manag. 2016, 5, 56–66. [Google Scholar]
  72. Shukla, V.; Shukla, P.; Tiwari, A. Lead poisoning. India J. Med. Spec. 2018, 9, 146–149. [Google Scholar] [CrossRef]
  73. Fatima, G.; Raza, A.M.; Hadi, N.; Nigam, N.; Mahdi, A.A. Cadmium in human diseases: It’s more than just a mere metal. Indian J. Clin. Biochem. 2019, 34, 371–378. [Google Scholar] [CrossRef] [PubMed]
  74. Obasi, P.N.; Akudinobi, B.B. Potential health risk and levels of heavy metals in water resources of lead–zinc mining communities of Abakaliki, southeast Nigeria. Appl. Water Sci. 2020, 10, 184. [Google Scholar] [CrossRef]
  75. Zahra, N.; Kalim, I. Perilous effects of heavy metals contamination on human health. Pak. J. Anal. Environ. Chem. 2017, 18, 1–17. [Google Scholar] [CrossRef]
  76. Priyadarshini, E.; Priyadarshini, S.S.; Pradhan, N. Heavy metal resistance in algae and its application for metal nanoparticle synthesis. Appl. Microbiol. Biotechnol. 2019, 103, 3297–3316. [Google Scholar] [CrossRef]
  77. Majumder, S.; Gupta, S.; Raghuvanshi, S. Removal of dissolved metals by bioremediation. In Heavy Metals in Water: Presence, Removal and Safety; Sharma, S.K., Ed.; The Royal Society of Chemistry: Cambridge, UK, 2015; pp. 44–56. [Google Scholar]
  78. Kumar, A.; Yadav, A.N.; Mondal, R.; Kour, D.; Subrahmanyam, G.; Shabnam, A.A.; Khan, S.A.; Yadav, K.K.; Sharma, G.K.; Cabral-Pinto, M.; et al. Myco-remediation: A mechanistic understanding of contaminants alleviation from natural environment and future prospect. Chemosphere 2021, 284, 131325. [Google Scholar] [CrossRef]
  79. Peng, W.; Li, X.; Xiao, S.; Fan, W. Review of remediation technologies for sediments contaminated by heavy metals. J. Soils Sediments 2018, 18, 1701–1719. [Google Scholar] [CrossRef]
  80. Matagi, S.; Swaiand, D.; Mugabe, R. A review of heavy metal removal mechanisms in wetlands. Afr. J. Trop. Hydrobiol. Fish. 1998, 8, 23. [Google Scholar] [CrossRef] [Green Version]
  81. Mitra, N.; Rezvan, Z.; Seyed Ahmad, M.; Gharaie, M.; Hosein, M. Studies of water arsenic and boron pollutants and algae phytoremediation in three springs, Iran. Int. J. Ecosys. 2012, 2, 32–37. [Google Scholar] [CrossRef] [Green Version]
  82. Saini, S.; Dhania, G. Cadmium as an environmental pollutant: Ecotoxicological effects, health hazards, and bioremediation approaches for its detoxification from contaminated sites. In Bioremediation of Industrial Waste for Environmental Safety; Springer: Singapore, 2020; pp. 357–387. [Google Scholar]
  83. Al-Homaidan, A.A.; Al-Ghanayem, A.A.; Areej, A.H. Green algae as bioindicators of heavy metal pollution in Wadi Hanifah Stream, Riyadh, Saudi Arabia. Int. J. Water Resour. Arid. Environ. 2011, 1, 10. [Google Scholar]
  84. Bonanno, G.; Orlando-Bonaca, M. Trace elements in Mediterranean seagrasses and macroalgae. A review. Sci. Total Environ. 2018, 618, 1152–1159. [Google Scholar] [CrossRef] [PubMed]
  85. Gosavi, K.; Sammut, J.; Gifford, S.; Jankowski, J. Macroalgal biomonitors of trace metal contamination in acid sulfate soil aquaculture ponds. Sci. Total Environ. 2004, 324, 25–39. [Google Scholar] [CrossRef]
  86. Rainbow, P.S. Biomonitoring of heavy metal availability in the marine environment. Mar. Pollut. Bull. 1995, 31, 183–192. [Google Scholar] [CrossRef]
  87. Abdel-Shafy, H.I.; Mansour, M.S. Phytoremediation for the elimination of metals, pesticides, PAHs, and other pollutants from wastewater and soil. In Phytobiont and Ecosystem Restitution; Springer: Singapore, 2018; pp. 101–136. [Google Scholar]
  88. Davis, T.A.; Volesky, B.; Mucci, A. A review of the biochemistry of heavy metal biosorption by brown algae. Water Res. 2003, 37, 4311. [Google Scholar] [CrossRef]
  89. Nielsen, H.D.; Burridge, T.R.; Brownlee, C.; Brown, M.T. Prior exposure to Cu contamination influences the outcome of toxicological testing of Fucus serratus embryos. Mar. Pollut. Bull. 2005, 50, 1675–1680. [Google Scholar] [CrossRef]
  90. Priyadarshani, I.; Sahu, D.; Rath, B. Microalgal bioremediation: Current practices and perspectives. J. Biochem. Technol. 2011, 3, 299–304. [Google Scholar]
  91. Perales-Vela, H.V.; Peña-Castro, J.M.; Cañizares Villanueva, R.O. Heavy metal detoxification in eukaryotic microalgae. Chemosphere 2006, 64, 1–10. [Google Scholar] [CrossRef]
  92. Monteiro, C.M.; Castro, P.M.L.; Malcata, F.X. Metal uptake by microalgae: Underlying mechanisms and practical applications. Biotechnol. Prog. 2012, 28, 299–311. [Google Scholar] [CrossRef]
  93. Figueira, M.M.F.; Volesky, B.; Azarian, K.; Ciminelli, V.S.T. Multi-metal bio-sorption in a column using Sargassum biomass. In Biohydrometallurgy and the Environment Toward the Mining of the 21st Century (Part B): International Biohydrometallurgy Symposium-Proceedings; Amils, R., Ballester, A., Eds.; Elsevier Science: Amsterdam, The Netherlands, 1999; pp. 503–512. [Google Scholar]
  94. Rajamani, S.; Siripornadulsil, S.; Falcao, V.; Torres, M.A.; Colepicolo, P.; Sayre, R. Phycorremediation of heavy metals Using Transgenic Microalgae. In Transgenic Microalgae as Green Cell Factories; León, R., Galván, A., Fernández, E., Eds.; Springer: Berlin/Heidelberg, Germany, 2007; pp. 99–107. [Google Scholar]
  95. de-Bashan, L.E.; Bashan, Y. Immobilized micro algae for removing pollutants: Review of practical aspects. Bioresour. Technol. 2010, 101, 1611–1627. [Google Scholar] [CrossRef] [PubMed]
  96. Brinza, L.; Dring, M.J.; Gavrilescu, M. Marine micro and macroalgal species as biosorbents for heavy metals. Environ. Eng. Manag. J. 2007, 6, 237–251. [Google Scholar] [CrossRef]
  97. Mondal, M.; Halder, G.; Oinam, G.; Indrama, T.; Tiwari, O.N. Bioremediation of organic and inorganic pollutants using microalgae. In New and Future Developments in Microbial Biotechnology and Bioengineering; Elsevier: Amsterdam, The Netherlands, 2019; pp. 223–235. [Google Scholar]
  98. 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]
  99. Tsuji, N.; Hirayanagi, N.; Iwabe, O.; Namba, T.; Tagawa, M.; Miyamoto, S.; Miyasaka, H.; Takagi, M.; Hirata, K.; Miyamoto, K. Regulation of phytochelatin synthesis by zinc and cadmium in marine green alga Dunaliella tertiolecta. Phytochemistry 2003, 62, 453–459. [Google Scholar] [CrossRef]
  100. Pflugmacher, S.; Sandermann, H. Cytochrome P450 monooxygenases for fatty acids and xenobiotics in marine microalgae. Plant Physiol. 1998, 117, 123–128. [Google Scholar] [CrossRef] [Green Version]
  101. French, C.E.; Rosser, S.J.; Davies, G.J.; Nicklin, S.; Bruce, N.C. Biodegradation of explosives by transgenic plants expressing pentaerythritol tetranitrate reductase. Nat. Biotechnol. 1999, 17, 491–494. [Google Scholar] [CrossRef]
  102. Hannink, N.; Rosser, S.J.; French, C.E.; Basran, A.; Murray, J.A.H.; Nicklin, A.; Bruce, N.C. Phyto-detoxification of TNT by transgenic plants expressing a bacterial nitroreductase. Nat. Biotechnol. 2001, 19, 1168–1172. [Google Scholar] [CrossRef]
  103. De Filippis, L.F.; Pallaghy, C.K. Heavy Metals: Sources and Biological Effects. In Advances in Limnology Series: Algae and Water Pollution, E; Rai, L.C., Gaur, J.P., Soeder, C.J., Eds.; Scheizerbartsche Press: Stuttgart, Germany, 1994; pp. 31–77. [Google Scholar]
  104. Wang, T.C.; Weissman, J.C.; Ramesh, G.; Varadarajan, R.; Benemann, J.R. Bioremoval of Toxic Elements with Aquatic Plants and Algae. In Bioremediation of Recalcitrant Organics; Hinchee, R.E., Anderson, D.B., Hoeppel, R.E., Eds.; Battelle Press: Columbus, OH, USA, 1995; p. 65. [Google Scholar]
  105. Upadhyay, A.K.; Singh, R.; Singh, D.P. Phycotechnological approaches toward wastewater management. In Emerging and Eco-Friendly Approaches for Waste Management; Springer: Singapore, 2019; pp. 423–435. [Google Scholar]
  106. Bursali, E.A.; Cavas, L.; Seki, Y.; Bozkurt, S.S.; Yurdakoc, M. Sorption of boron by invasive marine seaweed: Caulerpa racemosa var. cylindracea. Chem. Eng. J. 2009, 150, 385–390. [Google Scholar] [CrossRef]
  107. Liu, Y.; Yang, S.; Tan, S.; Lin, Y.; Tay, J.H. Aerobic granules: A novel zinc biosorbent. Lett. Appl. Microbiol. 2002, 35, 548–551. [Google Scholar] [CrossRef]
  108. Jin-Fen, P.; Rong-Gen, L.; Li, M. A review of heavy metal adsorption by marine algae. Chin. J. Oceanol. Limnol. 2000, 18, 260–264. [Google Scholar] [CrossRef]
  109. Oukarroum, A. Alleviation of metal-induced toxicity in aquatic plants by exogenous compounds: A mini-review. Water Air Soil Pollut. 2016, 227, 204. [Google Scholar] [CrossRef]
  110. Abd El-Hameed, M.M.; Abuarab, M.; Al-Ansari, N.; Mottaleb, S.A.; Bakeer, G.A.; Gyasi-Agyei, Y.; Mokhtar, A. Phytoremediation of Contaminated Water by Cadmium (Cd) Using Two Cyanobacteria Species (Anabaena Variabilis and Nostoc Muscorum). Environ. Sci. Eur. 2021, 33, 1–16. [Google Scholar]
  111. Dönmez, G.C.; Aksu, Z.; Ozturk, A.; Kutsal, T. A comparative study on heavy metal biosorption characteristics of some algae. Process Biochem. 1999, 34, 885–892. [Google Scholar] [CrossRef]
  112. Ahad, R.I.A.; Goswami, S.; Syiem, M.B. Biosorption and equilibrium isotherms study of cadmium removal by Nostoc muscorum Meg 1: Morphological, physiological and biochemical alterations. 3 Biotech 2017, 7, 104. [Google Scholar] [CrossRef] [Green Version]
  113. Manzoor, F.; Karbassi, A.; Golzary, A. Removal of heavy metal contaminants from wastewater by using Chlorella vulgaris beijerinck: A review. Curr. Environ. Manag. (Former. Curr. Environ. Eng.) 2019, 6, 174–187. [Google Scholar] [CrossRef]
  114. Zhai, J.; Li, X.; Li, W.; Rahaman, M.H.; Zhao, Y.; Wei, B.; Wei, H. Optimization of biomass production and nutrients removal by Spirulina platensis from municipal wastewater. Ecol. Eng. 2017, 108, 83–92. [Google Scholar] [CrossRef]
  115. Shukla, R.; Pandey, A.K.; Mishra, K.N. The efficacy of modified cyanobacterial biomass to remove Cr (VI) ions from aqueous solution. Indian J. Sci. Res. 2017, 16, 31–34. [Google Scholar]
  116. Sood, A.; Renuka, N.; Prasanna, R.; Ahluwalia, A.S. Cyanobacteria as potential options for wastewater treatment. In Phytoremediation; Springer: Cham, Switzerland, 2015; pp. 83–93. [Google Scholar]
  117. Tran, H.T.; Vu, N.D.; Matsukawa, M.; Okajima, M.; Kaneko, T.; Ohki, K.; Yoshikawa, S. Heavy metal biosorption from aqueous solutions by algae inhabiting rice paddies in Vietnam. J. Environ. Chem. Eng. 2016, 4, 2529–2535. [Google Scholar] [CrossRef]
  118. Al-Homaidan, A.A.; Alabdullatif, J.A.; Al-Hazzani, A.A.; Al-Ghanayem, A.A.; Alabbad, A.F. Adsorptive removal of cadmium ions by Spirulina platensis dry biomass. Saudi J. Biol. Sci. 2015, 22, 795–800. [Google Scholar] [CrossRef] [Green Version]
  119. Anjana, K.; Kaushik, A.; Kiran, B.; Nisha, R. Biosorption of Cr (VI) by immobilized biomass of two indigenous strains of cyanobacteria isolated from metal contaminated soil. J. Hazard. Mater. 2007, 148, 383–386. [Google Scholar] [CrossRef]
  120. Singh, S.; Kumar, V. Mercury detoxification by absorption, mercuric ion reductase, and exopolysaccharides: A comprehensive study. Environ. Sci. Pollut. Res. 2020, 27, 27181–27201. [Google Scholar] [CrossRef] [PubMed]
  121. Apiratikul, R.; Pavasant, P. Batch and column studies of biosorption of heavy metals by Caulerpa lentillifera. Biores. Technol. 2008, 99, 2766–2777. [Google Scholar] [CrossRef] [PubMed]
  122. Lodeiro, P.; Herrero, R.; de Sastre Vicente, M.E. The use of protonated Sargassum muticum as biosorbent for cadmium removal in a fixed-bed column. J. Hazard. Mater. 2006, 137, 244–253. [Google Scholar] [CrossRef]
  123. Vilar, V.J.P.; Botelho, C.M.S.; Boaventura, R.A.R. Lead uptake by algae Gelidium and composite material particles in a packed bed column. Chem. Eng. J. 2008, 144, 420–430. [Google Scholar] [CrossRef]
  124. Senthilkumar, R.; Vijayaraghavan, K.; Thilakavathi, M.; Iyer, P.V.R.; Velan, M. Seaweeds for the remediation of wastewaters contaminated with Zinc (II) ions. J. Hazard. Mater. B 2006, 136, 791–799. [Google Scholar] [CrossRef]
  125. Izquierdo, M.; Gabaldon, C.; Marzal, P.; Alvarez Hornos, F.J. Modeling of copper fixed-bed biosorption from wastewater by Posidonia oceanica. Biores. Technol. 2010, 101, 510–517. [Google Scholar] [CrossRef] [PubMed]
  126. Chu, K.H.; Hashim, M.A. Copper biosorption on immobilized seaweed biomass: Column breakthrough characteristics. J. Environ. Sci. 2007, 19, 928–932. [Google Scholar] [CrossRef]
  127. Arco, M.M.; Hanela, S.; Duran, J.; Dos Santos Afonso, M. Biosorption of Cu(II), Zn(II), Cd(II) and Pb(II) by dead biomasses of green alga Ulva lactuca and the development of a sustainable matrix for adsorption implementation. J. Hazard. Mater. 2012, 213–214, 123–132. [Google Scholar] [CrossRef]
  128. Mei, L.; Xitao, X.; Renhao, X.; Zhili, L. Effects of strontium-induced stress on marine microalgae Platymonas subcordiformis (Chlorophyta: Volvocales). Chin. J. Oceanol. Limnol. 2006, 24, 154. [Google Scholar] [CrossRef]
  129. De Filippis, L.F. The effect of sublethal concentration of mercury and Zinc on Chlorella IV characteristics of a general enzyme system for metallic ions. Zeit Schr. Panzenphysiologie. 1978, 86, 339. [Google Scholar] [CrossRef]
  130. De Filippis, L.F.; Pallaghy, C.K. The effect of sub-lethal concentrations of mercury and zinc on chlorella: III. Development and possible mechanisms of resistance to metals. Zeit Schr. Panzenphysiologie. 1976, 79, 323–335. [Google Scholar] [CrossRef]
  131. Wang, B.; Wang, J.; Zhang, W.; Meldrum, D.R. Application of synthetic biology in cyanobacteria and algae. Front. Microbiol. 2012, 3, 344. [Google Scholar] [CrossRef] [Green Version]
  132. Verma, S.; Kuila, A. Bioremediation of heavy metals by microbial process. Environ. Technol. Innov. 2019, 14, 100369. [Google Scholar] [CrossRef]
  133. Singh, A.; Prasad, S.M. Remediation of heavy metal contaminated ecosystem: An overview on technology advancement. Int. J. Environ. Sci. Technol. 2015, 12, 353–366. [Google Scholar] [CrossRef]
  134. Chekroun, K.B.; Baghour, M. The role of algae in phytoremediation of heavy metals: A review. J. Mater. Environ. Sci. 2013, 4, 873–880. [Google Scholar]
  135. Javanbakht, V.; Alavi, S.A.; Zilouei, H. Mechanisms of heavy metal removal using microorganisms as biosorbent. Water Sci. Technol. 2014, 69, 1775–1787. [Google Scholar] [CrossRef] [PubMed]
  136. Jan, S.; Parray, J.A. Approaches to Heavy Metal Tolerance in Plants. Springer: Singapore, 2016.
  137. Lahiri, S.; Ghosh, D.; Bhakta, J.N. Role of microbes in eco-remediation of perturbed aquatic ecosystem. In Handbook of Research on Inventive Bioremediation Techniques; Bhakta, J., Ed.; IGI Global: Hershey, PA, USA, 2017; pp. 70–107. [Google Scholar]
  138. Danouche, M.; El Ghachtouli, N.; El Baouchi, A.; El Arroussi, H. Heavy metals phycoremediation using tolerant green microalgae: Enzymatic and non-enzymatic antioxidant systems for the management of oxidative stress. J. Environ. Chem. Eng. 2020, 8, 104460. [Google Scholar] [CrossRef]
  139. Danouche, M.; El Ghachtouli, N.; El Arroussi, H. Phycoremediation mechanisms of heavy metals using living green microalgae: Physicochemical and molecular approaches for enhancing selectivity and removal capacity. Heliyon 2021, 7, 07609. [Google Scholar] [CrossRef]
  140. Kumar, K.S.; Dahms, H.U.; Won, E.J.; Lee, J.S.; Shin, K.H. Microalgae—A promising tool for heavy metal remediation. Ecotoxicol. Environ. Saf. 2015, 113, 329–352. [Google Scholar] [CrossRef]
  141. Park, D.M. Bioadsorption of rare earth elements through cell surface display of lanthanide binding tags. Environ. Sci. Technol. 2016, 50, 2735–2742. [Google Scholar] [CrossRef]
  142. Anastopoulos, I.; Kyzas, G.Z. Progress in batch biosorption of heavy metals onto algae. J. Mol. Liq. 2015, 209, 77–86. [Google Scholar] [CrossRef]
  143. Zeraatkar, A.K.; Ahmadzadeh, H.; Talebi, A.F.; Moheimani, N.R.; McHenry, M.P. Potential use of algae for heavy metal bioremediation, a critical review. J. Environ. Manag. 2016, 181, 817–831. [Google Scholar] [CrossRef] [PubMed]
  144. Luo, F.; Liu, Y.; Li, X.; Xuan, Z.; Ma, J. Biosorption of lead ion by chemically-modified biomass of marine brown algae Laminaria japonica. Chemosphere 2006, 64, 1122–1127. [Google Scholar] [CrossRef] [PubMed]
  145. He, J.; Chen, J.P. A comprehensive review on biosorption of heavy metals by algal biomass: Materials, performances, chemistry, and modeling simulation tools. Bioresour. Technol. 2014, 160, 67–78. [Google Scholar] [CrossRef]
  146. Rijstenbil, J.W.; Gerringa, L.J.A. Interactions of algal ligands, metal complexation and availability, and cell responses of the diatom Ditylum brightwellii with a gradual increase in copper. Aquat. Toxicol. 2002, 56, 115–131. [Google Scholar] [CrossRef]
  147. Salama, E.S.; Roh, H.S.; Dev, S.; Khan, M.A.; Abou-Shanab, R.A.; Chang, S.W.; Jeon, B.H. Algae as a green technology for heavy metals removal from various wastewater. World J. Microbiol. Biotechnol. 2019, 35, 1–19. [Google Scholar] [CrossRef]
  148. Wang, S.; Vincent, T.; Faur, C.; Guibal, E. Alginate and algal-based beads for the sorption of metal cations: Cu (II) and Pb (II). Int. J. Mol. Sci. 2016, 17, 1453. [Google Scholar] [CrossRef] [Green Version]
  149. Khan, S.; Shamshad, I.; Waqas, M.; Nawab, J.; Ming, L. Remediating industrial wastewater containing potentially toxic elements with four freshwater algae. Ecol. Eng. 2017, 102, 536–541. [Google Scholar] [CrossRef]
  150. Furey, P.C.; Deininger, A.; Liess, A. Substratum-associated microbiota. Water Environ. Res. 2016, 88, 1637–1671. [Google Scholar] [CrossRef]
  151. Kanchana, S.; Jeyanthi, J.; Kathiravan, R.; Suganya, K. Biosorption of heavy metals using algae: A review. Int. J. Pharma Med. Biol. Sci. 2014, 3, 1. [Google Scholar]
  152. Sweetly, J. Macroalgae as a potentially low-cost biosorbent for heavy metal removal: A review. Int. J. Pharm. Biol. Arch. 2014, 5, 17–26. [Google Scholar]
  153. Mustapha, M.U.; Halimoon, N. Microorganisms and biosorption of heavy metals in the environment: A review paper. J. Microb. Biochem. Technol. 2015, 7, 253–256. [Google Scholar] [CrossRef]
  154. Perpetuo, E.A.; Souza, C.B.; Nascimento, C.A.O. Engineering bacteria for bioremediation. In Progress in Molecular and Environmental Bioengineering from Analysis and Modeling to Technology Applications; Carpi, A., Ed.; Tech Publishers: Rijeka, Croatia, 2011; pp. 605–632. [Google Scholar]
  155. Mantzorou, A.; Navakoudis, E.; Paschalidis, K.; Ververidis, F. Microalgae: A potential tool for remediating aquatic environments from toxic metals. Int. J. Environ. Sci. Technol. 2018, 16, 1815–1830. [Google Scholar] [CrossRef]
  156. Tripathi, S.; Poluri, K.M. Adaptive and tolerance mechanism of microalgae in removal of cadmium from wastewater. In Algae; Springer: Singapore, 2021; pp. 63–88. [Google Scholar]
  157. Torres, M.A.; Barros, M.P.; Campos, S.C.G.; Pinto, E.; Rajamani, S.; Sayre, R.T.; Colepicolo, P. Biochemical biomarkers in algae and marine pollution: A review. Ecotoxicol. Environ. Saf. 2008, 71, 1–15. [Google Scholar] [CrossRef]
  158. Hayashi, Y.; Nakagawa, C.W.; Mutoh, N.; Isobe, M.; Goto, T. Two pathways in the biosynthesis of cadystins (gammaEC) nG in the cell-free system of the fission yeast. Biochem. Cell Biol. 1991, 69, 115–121. [Google Scholar] [CrossRef]
  159. Klapheck, S.; Schlunz, S.; Bergmann, L. Synthesis of phytochelatins and homo-phytochelatins in Pisum sativum L. Plant Physiol. 1995, 107, 515–521. [Google Scholar] [CrossRef] [Green Version]
  160. Chen, J.J.; Zhou, J.M.; Goldsbrough, P.B. Characterization of phytochelatin synthase from tomato. Physiol. Planta 1997, 101, 165–172. [Google Scholar] [CrossRef]
  161. Ahner, B.A.; Kong, S.; Morel, F.M.M. Phytochelatin production in marine algae. An interspecific comparison. Limnol. Oceanogr. 1995, 40, 649–657. [Google Scholar] [CrossRef]
  162. Cobbett, C.; Goldsbrough, P. Phytochelatins and metallothioneins: Roles in heavy metal detoxification and homeostasis. Annu. Rev. Plant Biol. 2002, 53, 159–182. [Google Scholar] [CrossRef] [Green Version]
  163. Bačkor, M.; Pawlik-Skowrońska, B.; Bud’ová, J.; Skowroński, T. Response to copper and cadmium stress in wild type and copper tolerant strains of the lichen alga Trebouxi aerici: Metal accumulation, toxicity and non-protein thiols. Plant Growth Regul. 2007, 52, 17–27. [Google Scholar] [CrossRef]
  164. Chmielewská, E.; Medved, J. Bioaccumulation of heavy metals by green algae Cladophora glomerata in a refinery sewage lagoon. Croat. Chem. Acta 2001, 74, 135–145. [Google Scholar]
  165. Henriques, B. Bioaccumulation of Hg, Cd and Pb by Fucus vesiculosus in single and multi-metal contamination scenarios and its effect on growth rate. Chemosphere 2017, 171, 208–222. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Multiple benefits of growing algae in wastewater.
Figure 1. Multiple benefits of growing algae in wastewater.
Hydrobiology 01 00021 g001
Figure 2. Deleterious effects of HMs [3,55,56,57].
Figure 2. Deleterious effects of HMs [3,55,56,57].
Hydrobiology 01 00021 g002
Figure 3. Phytoremediation mechanism of algae.
Figure 3. Phytoremediation mechanism of algae.
Hydrobiology 01 00021 g003
Table 1. Adverse effect of heavy metal on humans and plants. ROS, reactive oxygen species.
Table 1. Adverse effect of heavy metal on humans and plants. ROS, reactive oxygen species.
MetalAdverse EffectsReference
AsCarcinogenic effects
Hyperpigmentation, melanosis and keratosis in humans
Genotoxic, as it leads to the generation of ROS and causes lipid peroxidation
Immunotoxic
Modulates co-receptor expression
Causes Black foot disease
[64,65]
HgMutagenic effects
Minamata disease
Hampers cholesterol
[66,67]
CdThis leads to severe bone and kidney damage in humans
Anemia, bronchitis, emphysema,
Acute toxic effects in children
[68,69,70]
ZnCauses anemia
Phytotoxic
Leads to a decrease in muscular coordination
Causes pain in the abdomen
[71]
CuPhytotoxic
Damages a range of aquatic fauna
Corrosion and mucosal irritation
Disturbs the central nervous system and can lead to depression
[65]
CrIrritates gastrointestinal mucosa
Nephritis and death in humans at higher doses of Cr (VI)
[72]
NiHigh concentration may lead to DNA damage
Negative effect on fauna
Causes phytotoxicity
[65]
PbPhytotoxic
High concentration may lead to metabolic poison
Toxic to humans, aquatic fauna and livestock
Hypertension leading to brain damage
May lead to fatigue irritability, anemia and behavioral changes in children
[66,73,74]
Table 2. Heavy metal removal potential of algal species.
Table 2. Heavy metal removal potential of algal species.
AlgaeMetal RemovedDescription of Metal-Rich
Surrounding
References
Anacystis nidulansCuSolution of metal[103]
Tolypothrix tenuisCdAqueous solution[104]
Synechocystis sp., Scenedesmus obliquus,
and Chlorella vulgaris
Cr (VI), Ni, CuAqueous solution[105]
Nostoc muscorumCd and CuMulti metal solution[106]
N. rivularis and Nostoc linckiaCd and ZnSewage water[107]
Chlorella vulgarisNi and CuSingle and binary metal solution[108]
Spirulina sp.Trace elementCopper smelter and refinery effluent[109]
Aulosira fertilissimaNi and CrFree-cell condition[110]
Anabaena, subcylindrical, and Nostoc muscorumMn, Co, Pb and CuIndustrial wastewater and sewage[111]
Cladophora fascicularisPb and CuAqueous solution[112]
Spirulina platensis (Spi SORB)CuColumn reactor system[113]
Chroococcus sp. and Nostoc calcicolaCrMetal-contaminated soil[114]
Gloeocapsa and LyngbyaCrContaminated sites[115]
Aphanothece flocculosa and Spirulina platensisHgWet biomass[116]
Gloeothece sp. strain PCC 6909CuWastewater[117]
Hapalosiphon welwitschii NägelCdMetal solution[118]
Phormidium sp. NTMS02 and Oscillatoria sp. NTMS01Cr (VI)Aqueous solution[119]
Caulerpa racemosa and Sargassum wightiiCd, Pb, Cr (III and VI)Aqueous solution[120]
Caulerpa lentilliferaCu (II), Pb (II), Cd (II),Aqueous solution[121]
Sargassum spCu (II), Cd (II), Ni (II), Fe (III)Aqueous solution[122]
Gelidium sp.Pb (II), Cu (II), Cr (III)Aqueous solution[123]
Ulva reticulatCu (II), Co (II), Ni (II), Zn (II)Aqueous solution[124]
Posidonia oceanicaCu (II)Aqueous solution[125]
Sargassum bacculariaCu (II)Aqueous solution[126]
Ulva lactucaCu (II), Zn (II), Cd (II), Pb (II)Aqueous solution[127]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ankit; Bauddh, K.; Korstad, J. Phycoremediation: Use of Algae to Sequester Heavy Metals. Hydrobiology 2022, 1, 288-303. https://doi.org/10.3390/hydrobiology1030021

AMA Style

Ankit, Bauddh K, Korstad J. Phycoremediation: Use of Algae to Sequester Heavy Metals. Hydrobiology. 2022; 1(3):288-303. https://doi.org/10.3390/hydrobiology1030021

Chicago/Turabian Style

Ankit, Kuldeep Bauddh, and John Korstad. 2022. "Phycoremediation: Use of Algae to Sequester Heavy Metals" Hydrobiology 1, no. 3: 288-303. https://doi.org/10.3390/hydrobiology1030021

Article Metrics

Back to TopTop