MRecent advances encompassing molecular tools for genome editing, together with the advent of multi-omics technologies and computational approaches, have permitted the design of tailor-made microalgal cell factories with significantly improved heavy metal bioremediation capacity.
Heavy metals (HMs) are an integral constituent of the biosphere; they are naturally recycled in the environment through various biotic and abiotic processes, as part of biogeochemical cycles [1][2][1,2]. However, the dramatic rise in urbanization and industrialization has led to the release of alarmingly toxic levels of HMs, along with many other organic and inorganic pollutants in the environment. Aside from geochemical processes beyond one’s control, as is the case of erosion, atmospheric deposition, infiltration, thermal spring activity, and volcanic eruptions, HMs are increasingly penetrating aquatic systems, as a consequence of a wide range of anthropogenic activities, for example.g., discharges of untreated effluents from mining, spontaneous leaching from intensive agriculture, petroleum refining, improved performance of petroleum-based fuels, refuse burning, electroplating, printing, power generation, and other such activities carried out by (fine) chemical and metallurgical industries in the manufacture of microelectronic devices, paints, plastics, batteries, cosmetics, and medical equipment [3][4][3,4]. Conventional methods to remove HMs from effluents are chemical precipitation, solvent extraction, ion exchange, evaporation, adsorption, nanofiltration, ultrafiltration, reverse osmosis, and electrochemical treatments; unfortunately, they often prove inefficient in terms of energy input, environmental footprint, capital investment, and operational costs. As a consequence, huge amounts of non-treated domestic and industrial waste end up in the natural environment [2][4][5][2,4,14]. Bioremediation comes as a tool to avoid such dumping (bearing several well-documented advantages [6][7][8][9][15–18]); it is defined at large as the application of biological organisms and their components to degrade, transform, sequester, mobilize, or contain environmental contaminants present in the soil, water, or air [10][19]. Various species of plants, bacteria, fungi, yeasts, and microalgae, as well as dead biomass-derived therefrom, have shown a great potential for bioremediation of HM-derived pollution in aqueous media—through metal binding and uptake [11][12][13][20–22]. Among these, phycoremediation (for i.example., use of microalgae for mitigation of organic and inorganic contamination) offers several advantages; hence, it has accordingly undergone intensive investigation for the large-scale remediation of industrial and domestic effluents, as well as HM-contaminated sites and water bodies [14][15][12,23]. Microalgae are photoautotrophic eukaryotic microorganisms, which account for most of the biologically sequestered trace metals in aquatic environments [16] [24]. Their unique metabolic plasticity, inherent capacity to grow on nonarable lands and in wastewater, using just solar light as the source of energy and atmospheric CO2 as the carbon source, and relatively high rate of cell division and growth account for such widespread occurrence; the exceptional retention capacity of HMs also contributes to make them the ideal platform to develop the next-generation technologies for wastewater treatment. Microalgae cells can accumulate HMs up to 10% of their biomass, owing to their large surface-to-volume ratio, coupled with their efficient metal binding, uptake, metabolization, and storage mechanisms [12][13][12,22]. Despite their outstanding potential, current technical and economic constraints—associated with the underlying upstream and downstream processes—have hampered the large-scale use of microalgae in HM bioremediation. Even though the integration of microalgae-mediated wastewater treatment with energy production appears logical and inevitable, the strains most commonly employed, essentially retrieved from nature in their native state, lack the robustness required by sustainable, large-scale scenarios bearing a commercial interest. Recent advances in genetic and protein engineering, complemented by an essentially unanimous orientation toward a holistic approach to the engineering of biological systems, at the expense of bioinformatic and omics tools, may soon allow for the tailor-made design of microbial cell factories for a number of end- or start-products while lowering the risks and concerns over the putative adverse effects associated to use of genetically modified organisms (GMOs). Further to knowledge on the most obvious routes to enhance energy load (stemming from hydrocarbons) and improving growth rate and photosynthetic ability in microalgae cells, effective engineering of (sustainable) cell factories, for efficient HM bioremediation, will call for work on specific genes and traits identified as relevant.
Heavy metals adversely affect the physiological health of, and may even cause severe toxicity to microalgae cells, owing to attenuation of the bioactivity of proteins, lipids, nucleic acids, pigments, and other molecules, or the generation of excessive reactive oxygen species (ROS); more specifically, they act by impairing photosynthetic machinery, inhibiting enzyme activities, and/or ceasing cell division [17][29], or else by inhibiting the normal function of the thylakoid membrane and chlorophyll biosynthesis, acidifying cytoplasm, or damaging the cell membrane [18][19][20][30–32]. Similar to other life forms, microalgae have developed, through evolution, several intracellular and extracellular adaptive mechanisms for the mitigation of HM toxicity. For instance, the physicochemical properties of the microalgal cell wall and extracellular polymeric substances (EPS) allow the binding of HM ions to functional groups on their surface, in a process generally known as biosorption [21][33]. As the interface between the intracellular compartment and external environment, the constitutive macromolecules of the cell wall possess various negatively charged functional groups, for example.g., amino, hydroxyl, carboxyl, sulfhydryl, sulfate, phosphate, carbonyl, amide, imidazole, thioether, and phenol; said moieties can bind to ions from the surrounding medium, in the absence of steric or conformational barriers [22][23][24][25][34–37]. The molecular mechanisms behind the biosorption of HMs onto the cell wall and EPS include ion exchange, chelation and complexation, hydroxide condensation, covalent binding, redox interaction, biomineralization, and precipitation of insoluble metal complexes, through electrostatic, van der Waals, or hydrophobic interactions of positively charged HM cations with negatively charged groups present on the cell surface [1][26][27][1,20,38]. The adsorption capacity of the microalgal cell wall is a metabolism-independent process; hence, it is primarily affected by such environmental factors as pH, temperature, contact time, and concentration of HM and competing ions [4][28][4,39]. Given the numerous reports on fluidity and evolution of cell wall components (for example.g., fatty acids) in response to external stimuli, a yet unknown metabolic background to this phenomenon seems to exist. On the other hand, biosorption of HMs onto EPS is regulated by the cell itself via changes in the properties of such biopolymers, as required by the nature of metabolic stress, for examplei.e., metal toxicity in the situation under scrutiny [40,41]. Both the cell wall and EPS provide, indeed, an extracellular protective layer to the cell that prevents the harmful effects of HMs, if transported into the intracellular compartment; in this fashion, cellular integrity is maintained. The distinct physiology of existing species of microalgae then accounts for the differences found in the composition and structure of such outer structures, which, in turn, drive their species- and even strain-dependent HM-biosorption capacities [1][11][1,20]. Secretion of metal-chelating proteins and specific organic acids, and subsequent endocytosis of the organometallic complexes formed, is another mechanism for extracellular HM-bioremediation, reported in microalgae cells [29][42]. Since HMs are hydrophilic in nature, a requirement exists for certain carrier molecules that facilitate their transport into the cells. Once in the cytoplasm, HM toxicity is overcome by resorting to unique metabolic mechanisms—some of which have been well-documented. Several metal efflux pumps do regulate the algal membrane permeability, by actively transporting HMs into and out of the cell [30][31][32][43–45]; the net metal flux is accordingly reduced, and may even affect the chemical speciation of the HMs, due to expulsion of trace metal complexes [33][46]. Another strategy followed by microalgae is increased expression of metal-binding amino acids, peptides, and proteins, such as metallothioneins (MTs), phytochelatins (PCs), glutathione (GSH), proline, histidine, and glutamate [34][47]. These organometallic complexes are typically transported in, but partitioned into vacuoles—so as to neutralize the otherwise toxic effects of HMs in the cytoplasm [5]. In the acidic environment of vacuoles, HMs are released from their organic carrier; while the latter may be transported back to the cytosol, HMs are most likely stabilized and chelated by sulfides or organic acids in said vacuoles [35][36][37][10,48,49]. Depending on the prevailing environmental conditions, microalgae may attain a high polyphosphate content, suitable for binding divalent HM cations, and drive them into vacuoles for further sequestration. In addition to vacuoles, excess intracellular HM loads can be transported for eventual sequestration in such other organelles as mitochondria and chloroplasts [32][45], meaning that the expression level of metal transporters in the membrane of those organelles will play an important role in determining HM-removal specificity, capacity, and rate by microalgae strains. As happens with other cell types, microalgae respond to HM-generated oxidative stress by controlling the cell redox state and overexpressing heat shock proteins (HSPs) [38][39][50,51]. MicroRNAs (miRNAs) also serve as key components of the gene regulatory network involved in cellular HM mitigation. They contribute to post-transcriptional cleavage and translational inhibition of target mRNAs, or methylation of target DNAs to regulate a particular response, aimed at maintaining cellular homeostasis, by triggering complexation of excess HMs, defense against oxidative stress, and signal transduction for biological control purposes [40][52]. Despite the scarce information available on the subject, metal-responsive transcription factors (TFs) appear to activate multiple genes responsible for HM uptake, transport, and detoxification, thus, establishing a global resistance network against HM toxicity in the cell. Therefore, identification and characterization of the set of TFs able to regulate HM stress will be of the utmost importance, in attempts to develop transgenic microalgae with improved bioremediation potential [17][41][29,53]. Sexual reproduction, expression of metal-modifying enzymes, and phenotypic plasticity are alternative mechanisms entertained by microalgae as survival tools against HM toxicity [37][49]. For instance, transcriptomic analysis of Chlorella vulgaris, following exposure to Cu cations, unfolded a marked increase in intracellular carotenoids and proline contents, and in activity of such antioxidative enzymes as catalase, peroxidase, polyphenol oxidase, and superoxide dismutase. Photosystem II (PSII) and CO2 assimilation are apparently inhibited in microalgae, as a response to metal stress; a relative reduction in growth rate and cell density has been reported for various species of microalgae upon exposure to high concentrations of Cu. A severe drop in protein levels, in parallel to an enhanced rate of carbohydrate biosynthesis, has been demonstrated in microalgae cells in the presence of HMs [42][43][54,55]. The aforementioned coordinated response of microalgae cells to metal toxicity is a result of crosstalk among different molecular networks, urged by the need for keeping cellular hemostasis. Similar to other cellular functions, the key role played by such signaling molecules as phytohormones, Ca itself, and kinases is worthy of further research [29,56].
Despite being one of the most sophisticated biological systems regarding resistance to and transformation of heavy metals, the evidence shows that further improvements are possible in such phycoremediation ability—namely through specifically designed genetic modifications. Acceleration, yet under tight control, of the underlying forces of natural evolution will likely play an important role in theour quest toward restoration of the lost ecological balance while furthering knowledge of the processes supporting life. Genetic engineering holds immense potential, yet careful experimentation and implementation are vital. Harnessing this potential and converting it into a useful technology for a better future, demands rational approaches, rather than prohibitive regulations tout court and strict banning; carefully designed bench- and industrial-scale efforts, complemented by open and transparent communication between science and technology stakeholders, and to the society at large, constitute the only reasonable and effective path thereto.
1. 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, e07609. [CrossRef]
2. Sibi, G. Factors influencing heavy metal removal by microalgae—A review. J. Crit. Rev. 2019, 6, 29–32.
3. Kumar, V.; Kothari, R.; Kumari, S.; Kumar, P. Sustainable approaches towards wastewater treatment using algal technology along with management of post-harvest biomass. In Advances in Environmental Pollution Management: Wastewater Impacts and Treatment Technologies, 1st ed.; Agro Environ Media: Haridwar, India, 2020; Volume 1, pp. 174–187.
4. Salam, K.A. Towards sustainable development of microalgal biosorption for treating effluents containing heavy metals. Biofuel Res. J. 2019, 6, 948–961. [CrossRef]
5. Balzano, S.; Sardo, A.; Blasio, M.; Chahine, T.B.; dell’Anno, F.; Sansone, C.; Brunet, C. Microalgal metallothioneins and phytochelatins and their potential use in bioremediation. Front. Microbiol. 2020, 11, 517. [CrossRef] [PubMed]
6. Dung, T.T.T.; Cappuyns, V.; Swennen, R.; Phung, N.K. From geochemical background determination to pollution assessment of heavy metals in sediments and soils. Rev. Environ. Sci. Bio/Technol. 2013, 12, 335–353. [CrossRef]
7. Singh, M.; Kumar, J.; Singh, S.; Singh, V.; Prasad, S.; Singh, M. Adaptation strategies of plants against heavy metal toxicity: A short review. Biochem. Pharmacol. 2015, 4, 161.
8. Mishra, S.; Bharagava, R.N.; More, N.; Yadav, A.; Zainith, S.; Mani, S.; Chowdhary, P. Heavy metal contamination: An alarming threat to environment and human health. In Environmental Biotechnology for Sustainable Future; Springer: Singapore, 2019; pp. 103–125.
9. Bruland, K.; Lohan, M. Controls of trace metals in seawater. In Treatise on Geochemistry, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2003; pp. 23–47.
10. Sunitha, M.S.; Prashant, S.; Kumar, S.A.; Rao, S.; Narasu, M.L.; Kishor, P.K. Cellular and molecular mechanisms of heavy metal tolerance in plants: A brief overview of transgenic plants overexpressing phytochelatin synthase and metallothionein genes. Plant Cell Biotechnol. Mol. Biol. 2013, 14, 33–48.
11. Nriagu, J.O.; Pacyna, J.M. Quantitative assessment of worldwide contamination of air, water and soils by trace metals. Nature 1988, 333, 134–139. [CrossRef] [PubMed]
12. Tripathi, S.; Arora, N.; Gupta, P.; Pruthi, P.A.; Poluri, K.M.; Pruthi, V. Microalgae: An emerging source for mitigation of heavy metals and their potential implications for biodiesel production. In Advanced Biofuels; Elsevier: Amsterdam, The Netherlands, 2019; pp. 97–128.
13. Tchounwou, P.B.; Yedjou, C.G.; Patlolla, A.K.; Sutton, D.J. Heavy metal toxicity and the environment. Mol. Clin. Environ. Toxicol. 2012, 101, 133–164.
14. Chan, A.; Salsali, H.; McBean, E. Heavy metal removal (copper and zinc) in secondary effluent from wastewater treatment plants by microalgae. ACS Sustain. Chem. Eng. 2014, 2, 130–137. [CrossRef]
15. Macek, T.; Mackova, M. Potential of biosorption technology. In Microbial Biosorption of Metals; Springer: Berlin/Heidelberg, Germany, 2011; pp. 7–17.
16. Ahalya, N.; Ramachandra, T.; Kanamadi, R. Biosorption of heavy metals. Reseach J. Chem. Environ. 2003, 7, 71–79.
17. Gadd, G.M. Biosorption: Critical review of scientific rationale, environmental importance and significance for pollution treatment. J. Chem. Technol. Biotechnol. Int. Res. Process Environ. Clean Technol. 2009, 84, 13–28. [CrossRef]
18. Eccles, H. Treatment of metal-contaminated wastes: Why select a biological process? Trends Biotechnol. 1999, 17, 462–465. [CrossRef]
19. Hazen, T.C. Bioremediation. In The Microbiology of the Terrestrial Deep Subsurface; CRC Press: Boca Raton, FL, USA, 2018; pp. 247–266.
20. 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. [CrossRef]
21. Fu, F.;Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. [CrossRef]
22. Kaplan, D. Absorption and adsorption of heavy metals by microalgae. Handb. Microalgal Cult. Appl. Phycol. Biotechnol. 2013, 2, 602–611.
23. John, J. Phycoremediation. In Modern Trends in Applied Aquatic Ecology; Springer: Berlin/Heidelberg, Germany, 2003; pp. 133–147.
24. Rajamani, S.; Siripornadulsil, S.; Falcão, V.; Torres, M.; Colepicolo, P.; Sayre, R. Phycoremediation of heavy metals using transgenic microalgae. In Transgenic Microalgae Green Cell Factories; Springer: Berlin/Heidelberg, Germany, 2007; pp. 99–109.
25. Singh, P.; Gupta, S.K.; Guldhe, A.; Rawat, I.; Bux, F. Microalgae isolation and basic culturing techniques. In Handbook of Marine Microalgae; Elsevier: Amsterdam, The Netherlands, 2015; pp. 43–54.
26. Biswas, H.; Bandyopadhyay, D. Physiological responses of coastal phytoplankton (Visakhapatnam, SW Bay of Bengal, India) to experimental copper addition. Mar. Environ. Res. 2017, 131, 19–31. [CrossRef] [PubMed]
27. Levy, J.L.; Angel, B.M.; Stauber, J.L.; Poon,W.L.; Simpson, S.L.; Cheng, S.H.; Jolley, D.F. Uptake and internalisation of copper by three marine microalgae: Comparison of copper-sensitive and copper-tolerant species. Aquat. Toxicol. 2008, 89, 82–93. [CrossRef]
28. Letry, K.A.; Castro, E.D.; Gupta, S.K.; Kumar, M. Industrial wastewater-based algal biorefineries: Application constraints and future prospects. In Application of Microalgae in Wastewater Treatment; Springer: Berlin/Heidelberg, Germany, 2019; pp. 371–392.
29. Tripathi, S.; Poluri, K.M. Heavy metal detoxification mechanisms by microalgae: Insights from transcriptomics analysis. Environ. Pollut. 2021, 285, 117443. [CrossRef]
30. Cheng, S.Y.; Show, P.-L.; Lau, B.F.; Chang, J.-S.; Ling, T.C. New prospects for modified algae in heavy metal adsorption. Trends Biotechnol. 2019, 37, 1255–1268. [CrossRef]
31. Samadani, M.; Perreault, F.; Oukarroum, A.; Dewez, D. Effect of cadmium accumulation on green algae Chlamydomonas reinhardtii and acid-tolerant Chlamydomonas CPCC 121. Chemosphere 2018, 191, 174–182. [CrossRef]
32. Dal Corso, G. Heavy metal toxicity in plants. In Plants and Heavy Metals; Springer: Berlin/Heidelberg, Germany, 2012; pp. 1–25.
33. Spain, O.; Plöhn, M.; Funk, C. The cell wall of green microalgae and its role in heavy metal removal. Physiol. Plant. 2021, 173, 526–535. [CrossRef] [PubMed]
34. Saavedra, R.; Muñoz, R.; Taboada, M.E.; Vega, M.; Bolado, S. Comparative uptake study of arsenic, boron, copper, manganese and zinc from water by different green microalgae. Bioresour. Technol. 2018, 263, 49–57. [CrossRef]
35. Leong, Y.K.; Chang, J.-S. Bioremediation of heavy metals using microalgae: Recent advances and mechanisms. Bioresour. Technol. 2020, 303, 122886. [CrossRef] [PubMed]
36. Javanbakht, V.; Alavi, S.A.; Zilouei, H. Mechanisms of heavy metal removal using microorganisms as biosorbent. Water Sci. Technol. 2014, 69, 1775–1787. [CrossRef]
37. Singh, D.V.; Bhat, R.A.; Upadhyay, A.K.; Singh, R.; Singh, D. Microalgae in aquatic environments: A sustainable approach for remediation of heavy metals and emerging contaminants. Environ. Technol. Innov. 2021, 21, 101340. [CrossRef]
38. Zhang, J.; Zhou, F.; Liu, Y.; Huang, F.; Zhang, C. Effect of extracellular polymeric substances on arsenic accumulation in Chlorella pyrenoidosa. Sci. Total Environ. 2020, 704, 135368. [CrossRef]
39. 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. [CrossRef] [PubMed]
40. Naveed, S.; Li, C.; Lu, X.; Chen, S.; Yin, B.; Zhang, C.; Ge, Y. Microalgal extracellular polymeric substances and their interactions with metal(loid)s: A review. Crit. Rev. Environ. Sci. Technol. 2019, 49, 1769–1802. [CrossRef]
41. Ubando, A.T.; Africa, A.D.M.; Maniquiz-Redillas, M.C.; Culaba, A.B.; Chen, W.-H.; Chang, J.-S. Microalgal biosorption of heavy metals: A comprehensive bibliometric review. J. Hazard. Mater. 2021, 402, 123431. [CrossRef] [PubMed]
42. Arunakumara, K.; Zhang, X. Heavy metal bioaccumulation and toxicity with special reference to microalgae. J. Ocean Univ. China 2008, 7, 60–64. [CrossRef]
43. 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, 15, 1815–1830. [CrossRef]
44. García-García, J.D.; Sánchez-Thomas, R.; Moreno-Sánchez, R. Bio-recovery of non-essential heavy metals by intra- and extracellular mechanisms in free-living microorganisms. Biotechnol. Adv. 2016, 34, 859–873. [CrossRef]
45. Monteiro, C.M.; Castro, P.M.; Malcata, F.X. Metal uptake by microalgae: Underlying mechanisms and practical applications. Biotechnol. Prog. 2012, 28, 299–311. [CrossRef] [PubMed]
46. Worms, I.; Simon, D.F.; Hassler, C.; Wilkinson, K. Bioavailability of trace metals to aquatic microorganisms: Importance of chemical, biological and physical processes on biouptake. Biochimie 2006, 88, 1721–1731. [CrossRef]
47. Hirata, K.; Tsuji, N.; Miyamoto, K. Biosynthetic regulation of phytochelatins, heavy metal-binding peptides. J. Biosci. Bioeng. 2005, 100, 593–599. [CrossRef] [PubMed]
48. Penen, F.; Isaure, M.P.; Dobritzsch, D.; Bertalan, I.; Castillo-Michel, H.; Proux, O.; Gontier, E.; le Coustumer, P.; Schaumlöffel, D. Pools of cadmium in Chlamydomonas reinhardtii revealed by chemical imaging and XAS spectroscopy. Metallomics 2017, 9, 910–923. [CrossRef]
49. Perales-Vela, H.V.; Peña-Castro, J.M.; Cañizares-Villanueva, R.O. Heavy metal detoxification in eukaryotic microalgae. Chemosphere 2006, 64, 1–10. [CrossRef]
50. Wang, W.; Vinocur, B.; Shoseyov, O.; Altman, A. Role of plant heat-shock proteins and molecular chaperones in the abiotic stress response. Trends Plant Sci. 2004, 9, 244–252. [CrossRef]
51. Tukaj, S.; Tukaj, Z. Distinct chemical contaminants induce the synthesis of Hsp70 proteins in green microalgae Desmodesmus subspicatus: Heat pretreatment increases cadmium resistance. J. Therm. Biol. 2010, 35, 239–244. [CrossRef]
52. Gielen, H.; Remans, T.; Vangronsveld, J.; Cuypers, A. MicroRNAs in metal stress: Specific roles or secondary responses? Int. J. Mol. Sci. 2012, 13, 15826–15847. [CrossRef] [PubMed]
53. Yang, X.E.; Jin, X.F.; Feng, Y.; Islam, E. Molecular mechanisms and genetic basis of heavy metal tolerance/hyperaccumulation in plants. J. Integr. Plant Biol. 2005, 47, 1025–1035. [CrossRef]
54. Kebeish, R.; el-Ayouty, Y.; Husain, A. Effect of copper on growth, bioactive metabolites, antioxidant enzymes and photosynthesis-related gene transcription in Chlorella vulgaris. World J. Biol. Biol. Sci. 2014, 2, 34–43.
55. Li, M.; Hu, C.; Zhu, Q.; Chen, L.; Kong, Z.; Liu, Z. Copper and zinc induction of lipid peroxidation and effects on antioxidant enzyme activities in the microalga Pavlova viridis (Prymnesiophyceae). Chemosphere 2006, 62, 565–572. [CrossRef] [PubMed]
56. Nguyen, H.N.; Kisiala, A.B.; Emery, R.N. The roles of phytohormones in metal stress regulation in microalgae. J. Appl. Phycol.
2020, 32, 3817–3829. [CrossRef]
57. Kumar, G.; Shekh, A.; Jakhu, S.; Sharma, Y.; Kapoor, R.; Sharma, T.R. Bioengineering of microalgae: Recent advances, perspectives, and regulatory challenges for industrial application. Front. Bioeng. Biotechnol. 2020, 8, 914. [CrossRef]
58. Guihéneuf, F.; Khan, A.; Tran, L.-S.P. Genetic engineering: A promising tool to engender physiological, biochemical, and molecular stress resilience in green microalgae. Front. Plant Sci. 2016, 7, 400. [CrossRef]
59. Fajardo, C.; de Donato, M.; Carrasco, R.; Martínez-Rodríguez, G.; Mancera, J.M.; Fernández-Acero, F.J. Advances and challenges in genetic engineering of microalgae. Rev. Aquac. 2020, 12, 365–381. [CrossRef]
60. Poliner, E.; Clark, E.; Cummings, C.; Benning, C.; Farre, E.M. A high-capacity gene stacking toolkit for the oleaginous microalga, Nannochloropsis oceanica CCMP1779. Algal Res. 2020, 45, 101664. [CrossRef]
61. Crozet, P.; Navarro, F.J.; Willmund, F.; Mehrshahi, P.; Bakowski, K.; Lauersen, K.J.; Pérez-Pérez, M.-E.; Auroy, P.; Gorchs Rovira, A.; Sauret-Gueto, S. Birth of a photosynthetic chassis: A MoClo toolkit enabling synthetic biology in the microalga Chlamydomonas reinhardtii. ACS Synth. Biol. 2018, 7, 2074–2086. [CrossRef]
62. Adler-Agnon, Z.; Leu, S.; Zarka, A.; Boussiba, S.; Khozin-Goldberg, I. Novel promoters for constitutive and inducible expression of transgenes in the diatom Phaeodactylum tricornutum under varied nitrate availability. J. Appl. Phycol. 2018, 30, 2763–2772. [CrossRef]
63. Shin, J.-H.; Choi, J.; Jeon, J.; Kumar, M.; Lee, J.; Jeong, W.-J.; Kim, S.-R. The establishment of new protein expression system using N starvation inducible promoters in Chlorella. Sci. Rep. 2020, 10, 12713. [CrossRef] [PubMed]
64. Doron, L.; Segal, N.A.; Shapira, M. Transgene expression in microalgae—From tools to applications. Front. Plant Sci. 2016, 7, 505. [CrossRef] [PubMed]
65. Ng, I.S.; Keskin, B.B.; Tan, S.I. A critical review of genome editing and synthetic biology applications in metabolic engineering of microalgae and cyanobacteria. Biotechnol. J. 2020, 15, 1900228. [CrossRef]
66. Sinha, R.; Shukla, P. Current trends in protein engineering: Updates and progress. Curr. Protein Pept. Sci. 2019, 20, 398–407. [CrossRef]
67. Yang, K.K.; Wu, Z.; Arnold, F.H. Machine-learning-guided directed evolution for protein engineering. Nat. Methods 2019, 16, 687–694. [CrossRef] [PubMed]
68. Pillai, S.; Behra, R.; Nestler, H.; Suter, M.J.-F.; Sigg, L.; Schirmer, K. Linking toxicity and adaptive responses across the transcriptome, proteome, and phenotype of Chlamydomonas reinhardtii exposed to silver. Proc. Natl. Acad. Sci. USA 2014, 111, 3490–3495. [CrossRef] [PubMed]
69. Mishra, A.; Medhi, K.; Malaviya, P.; Thakur, I.S. Omics approaches for microalgal applications: Prospects and challenges. Bioresour. Technol. 2019, 291, 121890. [CrossRef]
70. El-Sheekh, M.; el-Dalatony, M.; Thakur, N.; Zheng, Y.; Salama, E.-S. Role of microalgae and cyanobacteria in wastewater treatment: Genetic engineering and omics approaches. Int. J. Environ. Sci. Technol. 2021, 18, 1–22. [CrossRef]
71. Hainline, B. Algomics for the development of a sustainable microalgae biorefinery. Single Cell Biol. 2016, 5, 2.
72. May, P.; Wienkoop, S.; Kempa, S.; Usadel, B.; Christian, N.; Rupprecht, J.; Weiss, J.; Recuenco-Munoz, L.; Ebenhöh, O.; Weckwerth, W. Metabolomics- and proteomics-assisted genome annotation and analysis of the draft metabolic network of Chlamydomonas reinhardtii. Genetics 2008, 179, 157–166. [CrossRef] [PubMed]
73. Kanehisa, M.; Goto, S. KEGG: Kyoto encyclopedia of genes and genomes. Nucleic Acids Res. 2000, 28, 27–30. [CrossRef]
74. Lopez, D.; Casero, D.; Cokus, S.J.; Merchant, S.S.; Pellegrini, M. Algal functional annotation tool: A web-based analysis suite to functionally interpret large gene lists using integrated annotation and expression data. BMC Bioinform. 2011, 12, 1–10. [CrossRef] [PubMed]
75. Goodstein, D.M.; Shu, S.; Howson, R.; Neupane, R.; Hayes, R.D.; Fazo, J.; Mitros, T.; Dirks, W.; Hellsten, U.; Putnam, N. Phytozome: A comparative platform for green plant genomics. Nucleic Acids Res. 2012, 40, D1178–D1186. [CrossRef]
76. Dal’Molin, C.G.; Nielsen, L.K. Algae genome-scale reconstruction, modelling and applications. In The Physiology of Microalgae; Springer: Berlin/Heidelberg, Germany, 2016; pp. 591–598.
77. Boyle, N.R.; Morgan, J.A. Flux balance analysis of primary metabolism in Chlamydomonas reinhardtii. BMC Syst. Biol. 2009, 3, 1–14. [CrossRef] [PubMed]
78. Ding, N.; Wang, L.; Kang, Y.; Luo, K.; Zeng, D.; Man, Y.B.; Zhang, Q.; Zeng, L.; Luo, J.; Jiang, F. The comparison of transcriptomic response of green microalga Chlorella sorokiniana exposure to environmentally relevant concentration of cadmium (II) and 4-n-nonylphenol. Environ. Geochem. Health 2020, 42, 2881–2894. [CrossRef]
79. Jamers, A.; Blust, R.; de Coen,W.; Griffin, J.L.; Jones, O.A. Copper toxicity in the microalga Chlamydomonas reinhardtii: An integrated approach. Biometals 2013, 26, 731–740. [CrossRef] [PubMed]
80. Jamers, A.; Blust, R.; de Coen, W.; Griffin, J.L.; Jones, O.A. An omics based assessment of cadmium toxicity in the green alga Chlamydomonas reinhardtii. Aquat. Toxicol. 2013, 126, 355–364. [CrossRef] [PubMed]
81. Olsson, S.; Puente-Sanchez, F.; Gómez, M.J.; Aguilera, A. Transcriptional response to copper excess and identification of genes involved in heavy metal tolerance in the extremophilic microalga Chlamydomonas acidophila. Extremophiles 2015, 19, 657–672. [CrossRef] [PubMed]
82. Sanchez, L.R.S.; Cao, E.P. Metagenomic analysis reveals the presence of heavy metal response genes from cyanobacteria thriving in Balatoc Mines, Benguet Province, Philippines. Philipp. J. Sci. 2019, 148, 71–82.
83. Ye, J.; Song, Z.; Wang, L.; Zhu, J. Metagenomic analysis of microbiota structure evolution in phytoremediation of a swine lagoon wastewater. Bioresour. Technol. 2016, 219, 439–444. [CrossRef] [PubMed]
84. Kumari, P.; Kumar, Y. Bioinformatics and computational tools in bioremediation and biodegradation of environmental pollutants. In Bioremediation for Environmental Sustainability; Elsevier: Amsterdam, The Netherlands, 2021; pp. 421–444.
85. Hanikenne, M.; Krämer, U.; Demoulin, V.; Baurain, D. A comparative inventory of metal transporters in the green alga Chlamydomonas reinhardtii and the red alga Cyanidioschizon merolae. Plant Physiol. 2005, 137, 428–446. [CrossRef]
86. Rosakis, A.; Köster, W. Transition metal transport in the green microalga Chlamydomonas reinhardtii—Genomic sequence analysis. Res. Microbiol. 2004, 155, 201–210. [CrossRef] [PubMed]
87. Blaby-Haas, C.E.; Merchant, S.S. The ins and outs of algal metal transport. Biochim. Biophys. Acta Mol. Cell Res. 2012, 1823, 1531–1552. [CrossRef] [PubMed]
88. Nevo, Y.; Nelson, N. The NRAMP family of metal-ion transporters. Biochim. Biophys. Acta Mol. Cell Res. 2006, 1763, 609–620. [CrossRef] [PubMed]
89. Lu, J.; Ma, Y.; Xing, G.; Li, W.; Kong, X.; Li, J.; Wang, L.; Yuan, H.; Yang, J. Revelation of microalgae’s lipid production and resistance mechanism to ultra-high Cd stress by integrated transcriptome and physiochemical analyses. Environ. Pollut. 2019, 250, 186–195. [CrossRef]
90. Puente-Sánchez, F.; Díaz, S.; Penacho, V.; Aguilera, A.; Olsson, S. Basis of genetic adaptation to heavy metal stress in the acidophilic green alga Chlamydomonas acidophila. Aquat. Toxicol. 2018, 200, 62–72. [CrossRef] [PubMed]
91. Rosakis, A.; Köster, W. Divalent metal transport in the green microalga Chlamydomonas reinhardtii is mediated by a protein similar to prokaryotic Nramp homologues. Biometals 2005, 18, 107–120. [CrossRef] [PubMed]
92. Beauvais-Flück, R.; Slaveykova, V.I.; Cosio, C. Cellular toxicity pathways of inorganic and methyl mercury in the green microalga Chlamydomonas reinhardtii. Sci. Rep. 2017, 7, 8034. [CrossRef]
93. Hirooka, S.; Hirose, Y.; Kanesaki, Y.; Higuchi, S.; Fujiwara, T.; Onuma, R.; Era, A.; Ohbayashi, R.; Uzuka, A.; Nozaki, H. Acidophilic green algal genome provides insights into adaptation to an acidic environment. Proc. Natl. Acad. Sci. USA 2017, 114, E8304–E8313. [CrossRef]
94. Wang, Z.; Gui, H.; Luo, Z.; Zhen, Z.; Yan, C.; Xing, B. Dissolved organic phosphorus enhances arsenate bioaccumulation and biotransformation in Microcystis aeruginosa. Environ. Pollut. 2019, 252, 1755–1763. [CrossRef] [PubMed]
95. Ibuot, A.; Dean, A.P.; McIntosh, O.A.; Pittman, J.K. Metal bioremediation by CrMTP4 over-expressing Chlamydomonas reinhardtii
in comparison to natural wastewater-tolerant microalgae strains. Algal Res. 2017, 24, 89–96. [CrossRef]
96. Osundeko, O.; Dean, A.P.; Davies, H.; Pittman, J.K. Acclimation of microalgae to wastewater environments involves increased oxidative stress tolerance activity. Plant Cell Physiol. 2014, 55, 1848–1857. [CrossRef] [PubMed]
97. Rosenzweig, A.C.; Argüello, J.M. Toward a molecular understanding of metal transport by P1B-Type ATPases. Curr. Top. Membr. 2012, 69, 113–136. [PubMed]
98. Williams, L.E.; Mills, R.F. P1B-ATPases—An ancient family of transition metal pumps with diverse functions in plants. Trends Plant Sci. 2005, 10, 491–502. [CrossRef] [PubMed]
99. Ibuot, A.; Webster, R.E.; Williams, L.E.; Pittman, J.K. Increased metal tolerance and bioaccumulation of zinc and cadmium in Chlamydomonas reinhardtii expressing a AtHMA4 C-terminal domain protein. Biotechnol. Bioeng. 2020, 117, 2996–3005. [CrossRef]
100. Bækgaard, L.; Mikkelsen, M.D.; Sørensen, D.M.; Hegelund, J.N.; Persson, D.P.; Mills, R.F.; Yang, Z.; Husted, S.; Andersen, J.P.; Buch-Pedersen, M.J. A combined zinc/cadmium sensor and zinc/cadmium export regulator in a heavy metal pump. J. Biol. Chem. 2010, 285, 31243–31252. [CrossRef] [PubMed]
101. Verret, F.; Gravot, A.; Auroy, P.; Preveral, S.; Forestier, C.; Vavasseur, A.; Richaud, P. Heavy metal transport by AtHMA4 involves the N-terminal degenerated metal binding domain and the C-terminal His11 stretch. FEBS Lett. 2005, 579, 1515–1522. [CrossRef]
102. Ceasar, S.A.; Lekeux, G.; Motte, P.; Xiao, Z.; Galleni, M.; Hanikenne, M. Di-cysteine residues of the Arabidopsis thaliana HMA4 C-terminus are only partially required for cadmium transport. Front. Plant Sci. 2020, 11, 560. [CrossRef]
103. Lekeux, G.; Laurent, C.; Joris, M.; Jadoul, A.; Jiang, D.; Bosman, B.; Carnol, M.; Motte, P.; Xiao, Z.; Galleni, M. Di-cysteine motifs in the C-terminus of plant HMA4 proteins confer nanomolar affinity for zinc and are essential for HMA4 function in vivo. J. Exp. Bot. 2018, 69, 5547–5560. [CrossRef]
104. Ramírez-Rodríguez, A.E.; Bañuelos-Hernández, B.; García-Soto, M.J.; Govea-Alonso, D.G.; Rosales-Mendoza, S.; Alfaro de la Torre, M.C.; Monreal-Escalante, E.; Paz-Maldonado, L.M. Arsenic removal using Chlamydomonas reinhardtii modified with the gene acr3 and enhancement of its performance by decreasing phosphate in the growing media. Int. J. Phytoremed. 2019, 21, 617–623. [CrossRef] [PubMed]
105. Zhang, H.; Zeng, R.; Chen, D.; Liu, J. A pivotal role of vacuolar H+-ATPase in regulation of lipid production in Phaeodactylum tricornutum. Sci. Rep. 2016, 6, 31319. [CrossRef] [PubMed]
106. Choi, H.I.; Hwang, S.-W.; Kim, J.; Park, B.; Jin, E.; Choi, I.-G.; Sim, S.J. Augmented CO2 tolerance by expressing a single H+-pump
enables microalgal valorization of industrial flue gas. Nat. Commun. 2021, 12, 1–15. [CrossRef] [PubMed]
107. Simon, D.F.; Descombes, P.; Zerges, W.; Wilkinson, K.J. Global expression profiling of Chlamydomonas reinhardtii exposed to trace levels of free cadmium. Environ. Toxicol. Chem. Int. J. 2008, 27, 1668–1675. [CrossRef] [PubMed]
108. Khatiwada, B.; Hasan, M.T.; Sun, A.; Kamath, K.S.; Mirzaei, M.; Sunna, A.; Nevalainen, H. Proteomic response of Euglena gracilis to heavy metal exposure—Identification of key proteins involved in heavy metal tolerance and accumulation. Algal Res. 2020,
45, 101764. [CrossRef]
109. Kotrba, P.; Mackova, M.; Macek, T. Transgenic approaches to improve phytoremediation of heavy metal polluted soils. In biomanagement of metal-contaminated soils; Springer: New York, NY, USA, 2011; pp. 409–438.
110. Lee, J.; Bae, H.; Jeong, J.; Lee, J.-Y.; Yang, Y.-Y.; Hwang, I.; Martinoia, E.; Lee, Y. Functional expression of a bacterial heavy metal transporter in Arabidopsis enhances resistance to and decreases uptake of heavy metals. Plant Physiol. 2003, 133, 589–596. [CrossRef]
111. Korenkov, V.; Hirschi, K.; Crutchfield, J.D.; Wagner, G.J. Enhancing tonoplast Cd/H antiport activity increases Cd, Zn, and Mn tolerance, and impacts root/shoot Cd partitioning in Nicotiana tabacum L. Planta 2007, 226, 1379–1387. [CrossRef]
112. Hirschi, K.D.; Korenkov, V.D.; Wilganowski, N.L.; Wagner, G.J. Expression of Arabidopsis CAX2 in tobacco. Altered metal accumulation and increased manganese tolerance. Plant Physiol. 2000, 124, 125–134. [CrossRef]
113. Van der Zaal, B.J.; Neuteboom, L.W.; Pinas, J.E.; Chardonnens, A.N.; Schat, H.; Verkleij, J.A.; Hooykaas, P.J. Overexpression of a novel Arabidopsis gene related to putative zinc-transporter genes from animals can lead to enhanced zinc resistance and accumulation. Plant Physiol. 1999, 119, 1047–1056. [CrossRef]
114. Zheng, C.; Aslam, M.; Liu, X.; Du, H.; Xie, X.; Jia, H.; Huang, N.; Tang, K.; Yang, Y.; Li, P. Impact of Pb on Chlamydomonas reinhardtii at physiological and transcriptional levels. Front. Microbiol. 2020, 11, 1443. [CrossRef]
115. Puente-Sánchez, F.; Olsson, S.; Aguilera, A. Comparative transcriptomic analysis of the response of Dunaliella acidophila (Chlorophyta) to short-term cadmium and chronic natural metal-rich water exposures. Microb. Ecol. 2016, 72, 595–607. [CrossRef] [PubMed]
116. Li, Z.-S.; Lu, Y.-P.; Zhen, R.-G.; Szczypka, M.; Thiele, D.J.; Rea, P.A. A new pathway for vacuolar cadmium sequestration in Saccharomyces cerevisiae: YCF1-catalyzed transport of bis (glutathionato) cadmium. Proc. Natl. Acad. Sci. USA 1997, 94, 42–47. [CrossRef]
117. Song, W.-Y.; Sohn, E.J.; Martinoia, E.; Lee, Y.J.; Yang, Y.-Y.; Jasinski, M.; Forestier, C.; Hwang, I.; Lee, Y. Engineering tolerance and accumulation of lead and cadmium in transgenic plants. Nat. Biotechnol. 2003, 21, 914–919. [CrossRef]
118. Yazaki, K.; Yamanaka, N.; Masuno, T.; Konagai, S.; Kaneko, S.; Ueda, K.; Sato, F. Heterologous expression of a mammalian ABC transporter in plant and its application to phytoremediation. Plant Mol. Biol. 2006, 61, 491–503. [CrossRef] [PubMed]
119. Ortiz, D.F.; Ruscitti, T.; McCue, K.F.; Ow, D.W. Transport of metal-binding peptides by HMT1, a fission yeast ABC-type vacuolar membrane protein. J. Biol. Chem. 1995, 270, 4721–4728. [CrossRef] [PubMed]
120. Ortiz, D.; Kreppel, L.; Speiser, D.; Scheel, G.; McDonald, G.; Ow, D. Heavy metal tolerance in the fission yeast requires an ATP-binding cassette-type vacuolar membrane transporter. EMBO J. 1992, 11, 3491–3499. [CrossRef]
121. Cobbett, C.; Goldsbrough, P. Phytochelatins and metallothioneins: Roles in heavy metal detoxification and homeostasis. Annu. Rev. Plant Biol. 2002, 53, 159–182. [CrossRef] [PubMed]
122. Capdevila, M.; Atrian, S. Metallothionein protein evolution: A miniassay. J. Biol. Inorg. Chem. 2011, 16, 977–989. [CrossRef] [PubMed]
123. Leszczyszyn, O.I.; White, C.R.J.; Blindauer, C.A. The isolated Cys 2 His 2 site in EC metallothionein mediates metal-specific protein folding. Mol. Biosyst. 2010, 6, 1592–1603. [CrossRef] [PubMed]
124. Ziller, A.; Fraissinet-Tachet, L. Metallothionein diversity and distribution in the tree of life: A multifunctional protein. Metallomics 2018, 10, 1549–1559. [CrossRef]
125. Gaur, J.; Rai, L. Heavy metal tolerance in algae. In Algal Adaptation to Environmental Stresses; Springer: New York, NY, USA, 2001; pp. 363–388.
126. Ahner, B.A.; Wei, L.; Oleson, J.R.; Ogura, N. Glutathione and other low molecular weight thiols in marine phytoplankton under metal stress. Mar. Ecol. Prog. Ser. 2002, 232, 93–103. [CrossRef]
127. Meister, A.; Anderson, M.E. Glutathione. Annu. Rev. Biochem. 1983, 52, 711–760. [CrossRef]
128. 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. Phytochem. 2003, 62, 453–459. [CrossRef]
129. Hirata, K.; Tsujimoto, Y.; Namba, T.; Ohta, T.; Hirayanagi, N.; Miyasaka, H.; Zenk, M.H.; Miyamoto, K. Strong induction of phytochelatin synthesis by zinc in marine green alga, Dunaliella tertiolecta. J. Biosci. Bioeng. 2001, 92, 24–29. [CrossRef]
130. Suárez, C.; Torres, E.; Pérez-Rama, M.; Herrero, C.; Abalde, J. Cadmium toxicity on the freshwater microalga Chlamydomonas moewusii Gerloff: Biosynthesis of thiol compounds. Environ. Toxicol. Chem. 2010, 29, 2009–2015. [CrossRef] [PubMed]
131. Li, C.; Zheng, C.; Fu, H.; Zhai, S.; Hu, F.; Naveed, S.; Zhang, C.; Ge, Y. Contrasting detoxification mechanisms of Chlamydomonas reinhardtii under Cd and Pb stress. Chemosphere 2021, 274, 129771. [CrossRef] [PubMed]
132. Kobayashi, I.; Fujiwara, S.; Saegusa, H.; Inouhe, M.; Matsumoto, H.; Tsuzuki, M. Relief of arsenate toxicity by Cd-stimulated phytochelatin synthesis in the green alga Chlamydomonas reinhardtii. Mar. Biotechnol. 2006, 8, 94–101. [CrossRef]
133. Abboud, P.; Wilkinson, K.J. Role of metal mixtures (Ca, Cu and Pb) on Cd bioaccumulation and phytochelatin production by Chlamydomonas reinhardtii. Environ. Pollut. 2013, 179, 33–38. [CrossRef] [PubMed]
134. Pawlik-Skowro´ nska, B. Correlations between toxic Pb effects and production of Pb-induced thiol peptides in the microalga Stichococcus bacillaris. Environ. Pollut. 2002, 119, 119–127. [CrossRef]
135. Gómez-Jacinto, V.; García-Barrera, T.; Gómez-Ariza, J.L.; Garbayo-Nores, I.; Vílchez-Lobato, C. Elucidation of the defence mechanism in microalgae Chlorella sorokiniana under mercury exposure. Identification of Hg–phytochelatins. Chem. Biol. Interact. 2015, 238, 82–90. [CrossRef]
136. Nagalakshmi, N.; Prasad, M. Responses of glutathione cycle enzymes and glutathione metabolism to copper stress in Scenedesmus bijugatus. Plant Sci. 2001, 160, 291–299. [CrossRef]
137. Gekeler, W.; Grill, E.; Winnacker, E.-L.; Zenk, M.H. Algae sequester heavy metals via synthesis of phytochelatin complexes. Arch. Microbiol. 1988, 150, 197–202. [CrossRef]
138. Liang Zhu, Y.; Pilon-Smits, E.A.; Jouanin, L.; Terry, N. Overexpression of glutathione synthetase in Indian mustard enhances cadmium accumulation and tolerance. Plant Physiol. 1999, 119, 73–80. [CrossRef]
139. Piña-Olavide, R.; Paz-Maldonado, L.M.; de La Torre, M.C.A.; García-Soto, M.J.; Ramírez-Rodríguez, A.E.; Rosales-Mendoza, S.; Bañuelos-Hernández, B.; de la-Cruz, R.F.G. Increased removal of cadmium by Chlamydomonas reinhardtii modified with a synthetic gene for -glutamylcysteine synthetase. Int. J. Phytoremediat. 2020, 22, 1269–1277. [CrossRef] [PubMed]
140. Wang, Y.; Cheng, Z.Z.; Chen, X.; Zheng, Q.; Yang, Z.M. CrGNAT gene regulates excess copper accumulation and tolerance in Chlamydomonas reinhardtii. Plant Sci. 2015, 240, 120–129. [CrossRef] [PubMed]
141. Martínez, M.; Bernal, P.; Almela, C.; Vélez, D.; García-Agustín, P.; Serrano, R.; Navarro-Aviñó, J. An engineered plant that accumulates higher levels of heavy metals than Thlaspi caerulescens, with yields of 100 times more biomass in mine soils. Chemosphere 2006, 64, 478–485. [CrossRef]
142. Chaurasia, N.; Mishra, Y.; Rai, L.C. Cloning expression and analysis of phytochelatin synthase (pcs) gene from Anabaena sp. PCC 7120 offering multiple stress tolerance in Escherichia coli. Biochem. Biophys. Res. Commun. 2008, 376, 225–230. [CrossRef]
143. Picault, N.; Cazalé, A.; Beyly, A.; Cuiné, S.; Carrier, P.; Luu, D.; Forestier, C.; Peltier, G. Chloroplast targeting of phytochelatin synthase in Arabidopsis: Effects on heavy metal tolerance and accumulation. Biochimie 2006, 88, 1743–1750. [CrossRef]
144. Pal, R.; Rai, J. Phytochelatins: Peptides involved in heavy metal detoxification. Appl. Biochem. Biotechnol. 2010, 160, 945–963. [CrossRef]
145. Leszczyszyn, O.I.; Imam, H.T.; Blindauer, C.A. Diversity and distribution of plant metallothioneins: A review of structure, properties and functions. Metallomics 2013, 5, 1146–1169. [CrossRef]
146. Gutierrez, J.C.; Amaro, F.; Martín-González, A. From heavy metal-binders to biosensors: Ciliate metallothioneins discussed. BioEssays 2009, 31, 805–816. [CrossRef] [PubMed]
147. Gutiérrez, J.C.; Amaro, F.; Diaz, S.; de Francisco, P.; Cubas, L.; Martín-González, A. Ciliate metallothioneins: Unique microbial eukaryotic heavy-metal-binder molecules. J. Biol. Inorg. Chem. 2011, 16, 1025–1034. [CrossRef]
148. Kawakami, S.K.; Gledhill, M.; Achterberg, E.P. Production of phytochelatins and glutathione by marine phytoplankton in response to metal stress 1. J. Phycol. 2006, 42, 975–989. [CrossRef]
149. Cai, X.-H.; Brown, C.; Adhiya, J.; Traina, S.J.; Sayre, R.T. Growth and heavy metal binding properties of transgenic Chlamydomonas expressing a foreign metallothionein gene. Int. J. Phytoremediat. 1999, 1, 53–65. [CrossRef]
150. He, Z.; Siripornadulsil, S.; Sayre, R.T.; Traina, S.J.;Weavers, L.K. Removal of mercury from sediment by ultrasound combined with biomass (transgenic Chlamydomonas reinhardtii). Chemosphere 2011, 83, 1249–1254. [CrossRef]
151. Domínguez-Solís, J.R.; López-Martín, M.C.; Ager, F.J.; Ynsa, M.D.; Romero, L.C.; Gotor, C. Increased cysteine availability is essential for cadmium tolerance and accumulation in Arabidopsis thaliana. Plant Biotechnol. J. 2004, 2, 469–476. [CrossRef]
152. Kawashima, C.G.; Noji, M.; Nakamura, M.; Ogra, Y.; Suzuki, K.T.; Saito, K. Heavy metal tolerance of transgenic tobacco plants over-expressing cysteine synthase. Biotechnol. Lett. 2004, 26, 153–157. [CrossRef]
153. Freeman, J.L.; Salt, D.E. The metal tolerance profile of Thlaspi goesingense is mimicked in Arabidopsis thaliana heterologously expressing serine acetyl-transferase. BMC Plant Biol. 2007, 7, 1–10. [CrossRef]
154. Wangeline, A.L.; Burkhead, J.L.; Hale, K.L.; Lindblom, S.D.; Terry, N.; Pilon, M.; Pilon-Smits, E.A. Overexpression of ATP sulfurylase in Indian mustard: Effects on tolerance and accumulation of twelve metals. J. Environ. Qual. 2004, 33, 54–60. [CrossRef]
155. Torres, E.; Cid, A.; Fidalgo, P.; Herrero, C.; Abalde, J. Long-chain class III metallothioneins as a mechanism of cadmium tolerance in the marine diatom Phaeodactylum tricornutum Bohlin. Aquat. Toxicol. 1997, 39, 231–246. [CrossRef]
156. Pérez-Rama, M.; López, C.H.; Alonso, J.A.; Vaamonde, E.T. Class III metallothioneins in response to cadmium toxicity in the marine microalga Tetraselmis suecica (Kylin) Butch. Environ. Toxicol. Chem. Int. J. 2001, 20, 2061–2066. [CrossRef]
157. Tsednee, M.; Castruita, M.; Salomé, P.A.; Sharma, A.; Lewis, B.E.; Schmollinger, S.R.; Strenkert, D.; Holbrook, K.; Otegui, M.S.; Khatua, K. Manganese co-localizes with calcium and phosphorus in Chlamydomonas acidocalcisomes and is mobilized in manganese-deficient conditions. J. Biol. Chem. 2019, 294, 17626–17641. [CrossRef] [PubMed]
158. Goodenough, U.; Heiss, A.A.; Roth, R.; Rusch, J.; Lee, J.-H. Acidocalcisomes: Ultrastructure, biogenesis, and distribution in microbial eukaryotes. Protist 2019, 170, 287–313. [CrossRef] [PubMed]
159. Ruiz, F.A.; Marchesini, N.; Seufferheld, M.; Docampo, R. The polyphosphate bodies of Chlamydomonas reinhardtii possess a proton-pumping pyrophosphatase and are similar to acidocalcisomes. J. Biol. Chem. 2001, 276, 46196–46203. [CrossRef]
160. Sanz-Luque, E.; Bhaya, D.; Grossman, A.R. Polyphosphate: A multifunctional metabolite in cyanobacteria and algae. Front. Plant Sci. 2020, 11, 938. [CrossRef] [PubMed]
161. Wang, W.-X.; Dei, R.C. Metal stoichiometry in predicting Cd and Cu toxicity to a freshwater green alga Chlamydomonas reinhardtii. Environ. Pollut. 2006, 142, 303–312. [CrossRef] [PubMed]
162. Rao, N.N.; Gómez-García, M.R.; Kornberg, A. Inorganic polyphosphate: Essential for growth and survival. Annu. Rev. Biochem. 2009, 78, 605–647. [CrossRef]
163. Gao, F.;Wu, H.; Zeng, M.; Huang, M.; Feng, G. Overproduction, purification, and characterization of nanosized polyphosphate bodies from Synechococcus sp. PCC 7002. Microb. Cell Factories 2018, 17, 1–12. [CrossRef] [PubMed]
164. Zhu, J.; Loubéry, S.; Broger, L.; Zhang, Y.; Lorenzo-Orts, L.; Utz-Pugin, A.; Fernie, A.R.; Young-Tae, C.; Hothorn, M. A genetically validated approach for detecting inorganic polyphosphates in plants. Plant J. 2020, 102, 507–516. [CrossRef]
165. Gerasimait˙ e, R.; Sharma, S.; Desfougeres, Y.; Schmidt, A.; Mayer, A. Coupled synthesis and translocation restrains polyphosphate to acidocalcisome-like vacuoles and prevents its toxicity. J. Cell Sci. 2014, 127, 5093–5104. [CrossRef]
166. Aksoy, M.; Pootakham, W.; Grossman, A.R. Critical function of a Chlamydomonas reinhardtii putative polyphosphate polymerase subunit during nutrient deprivation. Plant Cell 2014, 26, 4214–4229. [CrossRef]
167. Chia, M.A.; Lombardi, A.T.; da Graça Gama Melão, M.; Parrish, C.C. Combined nitrogen limitation and cadmium stress stimulate total carbohydrates, lipids, protein and amino acid accumulation in Chlorella vulgaris (Trebouxiophyceae). Aquat. Toxicol. 2015, 160, 87–95. [CrossRef] [PubMed]
168. Yang, S.-L.; Lan, S.-S.; Gong, M. Hydrogen peroxide-induced proline and metabolic pathway of its accumulation in maize seedlings. J. Plant Physiol. 2009, 166, 1694–1699. [CrossRef]
169. Szabados, L.; Savouré, A. Proline: A multifunctional amino acid. Trends Plant Sci. 2010, 15, 89–97. [CrossRef]
170. Farago, M.; ME, F.;WA, M. Plants which accumulate metals. IV. A possible copper-proline complex from the roots of Armeria maritima. Inorg. Chim. Acta 1979, 32, L93–L94. [CrossRef]
171. Siripornadulsil, S.; Traina, S.; Verma, D.P.S.; Sayre, R.T. Molecular mechanisms of proline-mediated tolerance to toxic heavy metals in transgenic microalgae. Plant Cell 2002, 14, 2837–2847. [CrossRef] [PubMed]
172. Kolodyazhnaya, Y.S.; Titov, S.; Kochetov, A.; Trifonova, E.; Romanova, A.; Komarova, M.; Koval, V.; Shumny, V. Tobacco transformants expressing antisense sequence of proline dehydrogenase gene possess tolerance to heavy metals. Russ. J. Genet. 2007, 43, 825–828. [CrossRef]
173. Zheng, Q.; Cheng, Z.Z.; Yang, Z.M. HISN3 mediates adaptive response of Chlamydomonas reinhardtii to excess nickel. Plant Cell Physiol. 2013, 54, 1951–1962. [CrossRef]
174. Ma, J.F.; Ryan, P.R.; Delhaize, E. Aluminium tolerance in plants and the complexing role of organic acids. Trends Plant Sci. 2001, 6, 273–278. [CrossRef]
175. Boominathan, R.; Doran, P.M. Organic acid complexation, heavy metal distribution and the effect of ATPase inhibition in hairy roots of hyperaccumulator plant species. J. Biotechnol. 2003, 101, 131–146. [CrossRef]
176. Berg, C.V.d.; Wong, P.; Chau, Y. Measurement of complexing materials excreted from algae and their ability to ameliorate copper toxicity. J. Fish. Board Can. 1979, 36, 901–905. [CrossRef]
177. McKnight, D.M.; Morel, F.M. Release of weak and strong copper-complexing agents by algae 1. Limnol. Oceanogr. 1979, 24, 823–837. [CrossRef]
178. Yen, H.-W.; Chen, P.-W.; Hsu, C.-Y.; Lee, L. The use of autotrophic Chlorella vulgaris in chromium (VI) reduction under different reduction conditions. J. Taiwan Inst. Chem. Eng. 2017, 74, 1–6. [CrossRef]
179. Lee, L.; Hsu, C.-Y.; Yen, H.-W. The effects of hydraulic retention time (HRT) on chromium (VI) reduction using autotrophic cultivation of Chlorella vulgaris. Bioprocess Biosyst. Eng. 2017, 40, 1725–1731. [CrossRef] [PubMed]
180. Dhankher, O.P.; Li, Y.; Rosen, B.P.; Shi, J.; Salt, D.; Senecoff, J.F.; Sashti, N.A.; Meagher, R.B. Engineering tolerance and hyperaccumulation of arsenic in plants by combining arsenate reductase and -glutamylcysteine synthetase expression. Nat. Biotechnol. 2002, 20, 1140–1145. [CrossRef] [PubMed]
181. Yin, X.; Wang, L.; Duan, G.; Sun, G. Characterization of arsenate transformation and identification of arsenate reductase in a green alga Chlamydomonas reinhardtii. J. Environ. Sci. 2011, 23, 1186–1193. [CrossRef]
182. Arora, N.; Gulati, K.; Tripathi, S.; Pruthi, V.; Poluri, K.M. Algae as a budding tool for mitigation of arsenic from aquatic systems. In Mechanisms of Arsenic Toxicity and Tolerance in Plants; Springer: Berlin/Heidelberg, Germany, 2018; pp. 269–297.
183. Wang, Y.; Wang, S.; Xu, P.; Liu, C.; Liu, M.; Wang, Y.; Wang, C.; Zhang, C.; Ge, Y. Review of arsenic speciation, toxicity and metabolism in microalgae. Rev. Environ. Sci. Bio/Technol. 2015, 14, 427–451. [CrossRef]
184. Levy, J.L.; Stauber, J.L.; Adams, M.S.; Maher, W.A.; Kirby, J.K.; Jolley, D.F. Toxicity, biotransformation, and mode of action of arsenic in two freshwater microalgae (Chlorella sp. and Monoraphidium arcuatum). Environ. Toxicol. Chem. Int. J. 2005, 24, 2630–2639. [CrossRef] [PubMed]
185. Hasegawa, H.; Sohrin, Y.; Seki, K.; Sato, M.; Norisuye, K.; Naito, K.; Matsui, M. Biosynthesis and release of methylarsenic compounds during the growth of freshwater algae. Chemosphere 2001, 43, 265–272. [CrossRef]
186. Karadjova, I.B.; Slaveykova, V.I.; Tsalev, D.L. The biouptake and toxicity of arsenic species on the green microalga Chlorella salina in seawater. Aquat. Toxicol. 2008, 87, 264–271. [CrossRef]
187. Kelly, D.J.; Budd, K.; Lefebvre, D.D. Biotransformation of mercury in pH-stat cultures of eukaryotic freshwater algae. Arch. Microbiol. 2007, 187, 45–53. [CrossRef]
188. Huang, C.-C.; Chen, M.-W.; Hsieh, J.-L.; Lin, W.-H.; Chen, P.-C.; Chien, L.-F. Expression of mercuric reductase from Bacillus megaterium MB1 in eukaryotic microalga Chlorella sp. DT: An approach for mercury phytoremediation. Appl. Microbiol. Biotechnol. 2006, 72, 197–205. [CrossRef]
189. Bizily, S.P.; Kim, T.; Kandasamy, M.K.; Meagher, R.B. Subcellular targeting of methylmercury lyase enhances its specific activity for organic mercury detoxification in plants. Plant Physiol. 2003, 131, 463–471. [CrossRef] [PubMed]
190. Bizily, S.P.; Rugh, C.L.; Meagher, R.B. Phytodetoxification of hazardous organomercurials by genetically engineered plants. Nat. Biotechnol. 2000, 18, 213–217. [CrossRef]
191. Yang, H.; Nairn, J.; Ozias-Akins, P. Transformation of peanut using a modified bacterial mercuric ion reductase gene driven by an actin promoter from Arabidopsis thaliana. J. Plant Physiol. 2003, 160, 945–952. [CrossRef]
192. Heaton, A.C.; Rugh, C.L.; Kim, T.;Wang, N.J.; Meagher, R.B. Toward detoxifying mercury-polluted aquatic sediments with rice genetically engineered for mercury resistance. Environ. Toxicol. Chem. Int. J. 2003, 22, 2940–2947. [CrossRef] [PubMed]
193. Czakó-Vér, K.; Batiè, M.; Raspor, P.; Sipiczki, M.; Pesti, M. Hexavalent chromium uptake by sensitive and tolerant mutants of Schizosaccharomyces pombe. FEMS Microbiol. Lett. 1999, 178, 109–115. [CrossRef]
194. Rugh, C.L.; Wilde, H.D.; Stack, N.M.; Thompson, D.M.; Summers, A.O.; Meagher, R.B. Mercuric ion reduction and resistance in transgenic Arabidopsis thaliana plants expressing a modified bacterial merA gene. Proc. Natl. Acad. Sci. USA 1996, 93, 3182–3187.
[CrossRef] [PubMed]
195. Rugh, C.L.; Senecoff, J.F.; Meagher, R.B.; Merkle, S.A. Development of transgenic yellow poplar for mercury phytoremediation. Nat. Biotechnol. 1998, 16, 925–928. [CrossRef]
196. Che, D.; Meagher, R.B.; Heaton, A.C.; Lima, A.; Rugh, C.L.; Merkle, S.A. Expression of mercuric ion reductase in Eastern cottonwood (Populus deltoides) confers mercuric ion reduction and resistance. Plant Biotechnol. J. 2003, 1, 311–319. [CrossRef]
197. Lyyra, S.; Meagher, R.B.; Kim, T.; Heaton, A.; Montello, P.; Balish, R.S.; Merkle, S.A. Coupling two mercury resistance genes in Eastern cottonwood enhances the processing of organomercury. Plant Biotechnol. J. 2007, 5, 254–262. [CrossRef] [PubMed]
198. Khan, M.S.; Zaidi, A.; Goel, R.; Musarrat, J. Biomanagement of Metal-Contaminated Soils, 1st ed.; Springer: New York, NY, USA, 2011; pp. 409–438.
199. Hussein, H.S.; Ruiz, O.N.; Terry, N.; Daniell, H. Phytoremediation of mercury and organomercurials in chloroplast transgenic plants: Enhanced root uptake, translocation to shoots, and volatilization. Environ. Sci. Technol. 2007, 41, 8439–8446. [CrossRef] [PubMed]
200. Ruiz, O.N.; Hussein, H.S.; Terry, N.; Daniell, H. Phytoremediation of organomercurial compounds via chloroplast genetic engineering. Plant Physiol. 2003, 132, 1344–1352. [CrossRef]
201. Ruiz, O.N.; Daniell, H. Genetic engineering to enhance mercury phytoremediation. Curr. Opin. Biotechnol. 2009, 20, 213–219. [CrossRef] [PubMed]
202. Silver, S.; Phung, L.T. A bacterial view of the periodic table: Genes and proteins for toxic inorganic ions. J. Ind. Microbiol. Biotechnol. 2005, 32, 587–605. [CrossRef] [PubMed]
203. Singh, N.D.; Li, M.; Lee, S.-B.; Schnell, D.; Daniell, H. Arabidopsis Tic40 expression in tobacco chloroplasts results in massive proliferation of the inner envelope membrane and upregulation of associated proteins. Plant Cell 2008, 20, 3405–3417. [CrossRef]
204. Neuhierl, B.; Thanbichler, M.; Lottspeich, F.; Bock, A. A family of S-methylmethionine-dependent thiol/selenol methyltransferases: Role in selenium tolerance and evolutionary relation. J. Biol. Chem. 1999, 274, 5407–5414. [CrossRef]
205. Bañuelos, G.; Leduc, D.L.; Pilon-Smits, E.A.; Terry, N. Transgenic Indian mustard overexpressing selenocysteine lyase or selenocysteine methyltransferase exhibit enhanced potential for selenium phytoremediation under field conditions. Environ. Sci. Technol. 2007, 41, 599–605. [CrossRef]
206. LeDuc, D.L.; Tarun, A.S.; Montes-Bayon, M.; Meija, J.; Malit, M.F.;Wu, C.P.; Abdel-Samie, M.; Chiang, C.-Y.; Tagmount, A.; de Souza, M. Overexpression of selenocysteine methyltransferase in Arabidopsis and Indian mustard increases selenium tolerance and accumulation. Plant Physiol. 2004, 135, 377–383. [CrossRef]
207. Rai, A.; Singh, R.; Shirke, P.A.; Tripathi, R.D.; Trivedi, P.K.; Chakrabarty, D. Expression of rice CYP450-like gene (Os08g01480) in Arabidopsis modulates regulatory network leading to heavy metal and other abiotic stress tolerance. PLoS ONE 2015, 10, e0138574. [CrossRef]
208. Deng, L.; Wang, H.; Deng, N. Photoreduction of chromium (VI) in the presence of algae, Chlorella vulgaris. J. Hazard. Mater. 2006, 138, 288–292. [CrossRef]
209. Chaudhary, R.; Nawaz, K.; Khan, A.K.; Hano, C.; Abbasi, B.H.; Anjum, S. An overview of the algae-mediated biosynthesis of nanoparticles and their biomedical applications. Biomolecules 2020, 10, 1498. [CrossRef] [PubMed]
210. 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. [CrossRef] [PubMed]
211. Hamida, R.S.; Ali, M.A.; Redhwan, A.; Bin-Meferij, M.M. Cyanobacteria—A promising platform in green nanotechnology: A review on nanoparticles fabrication and their prospective applications. Int. J. Nanomed. 2020, 15, 6033. [CrossRef]
212. Jiang, Y.; Zhu, Y.; Hu, Z.; Lei, A.; Wang, J. Towards elucidation of the toxic mechanism of copper on the model green alga Chlamydomonas reinhardtii. Ecotoxicology 2016, 25, 1417–1425. [CrossRef] [PubMed]
213. Zhu, Q.-L.; Guo, S.-N.;Wen, F.; Zhang, X.-L.;Wang, C.-C.; Si, L.-F.; Zheng, J.-L.; Liu, J. Transcriptional and physiological responses of Dunaliella salina to cadmium reveals time-dependent turnover of ribosome, photosystem, and ROS-scavenging pathways. Aquat. Toxicol. 2019, 207, 153–162. [CrossRef] [PubMed]
214. Elleuch, J.; Hmani, R.; Drira, M.; Michaud, P.; Fendri, I.; Abdelkafi, S. Potential of three local marine microalgae from Tunisian coasts for cadmium, lead and chromium removals. Sci. Total Environ. 2021, 799, 149464. [CrossRef] [PubMed]
215. Athanasiadis, K. Effects of Heavy Metal Stress on Chlamydomonas reinhardtii: A Comprehensive Study. Master’s Thesis, University of Antwerp, Antwerp, Belgium, September 2017.
216. Ireland, H.E.; Harding, S.J.; Bonwick, G.A.; Jones, M.; Smith, C.J.;Williams, J.H. Evaluation of heat shock protein 70 as a biomarker of environmental stress in Fucus serratus and Lemna minor. Biomarkers 2004, 9, 139–155. [CrossRef]
217. Neumann, D.; Lichtenberger, O.; Günther, D.; Tschiersch, K.; Nover, L. Heat-shock proteins induce heavy-metal tolerance in higher plants. Planta 1994, 194, 360–367. [CrossRef]
218. Lewis, S.; Donkin, M.; Depledge, M. Hsp70 expression in Enteromorpha intestinalis (Chlorophyta) exposed to environmental stressors. Aquat. Toxicol. 2001, 51, 277–291. [CrossRef]
219. Vierling, E. The roles of heat shock proteins in plants. Annu. Rev. Plant Biol. 1991, 42, 579–620. [CrossRef]
220. Sathasivam, R.; Ki, J.-S. Heat shock protein genes in the green alga Tetraselmis suecica and their role against redox and non-redox active metals. Eur. J. Protistol. 2019, 69, 37–51. [CrossRef] [PubMed]
221. Kimura, T.; Itoh, N.; Andrews, G.K. Mechanisms of heavy metal sensing by metal response element-binding transcription factor-1. J. Health Sci. 2009, 55, 484–494. [CrossRef]
222. Lichtlen, P.; Schaffner, W. Putting its fingers on stressful situations: The heavy metal-regulatory transcription factor MTF-1. Bioessays 2001, 23, 1010–1017. [CrossRef]
223. Langmade, S.J.; Ravindra, R.; Daniels, P.J.; Andrews, G.K. The transcription factor MTF-1 mediates metal regulation of the mouse ZnT1 gene. J. Biol. Chem. 2000, 275, 34803–34809. [CrossRef] [PubMed]
224. Günther, V.; Lindert, U.; Schaffner, W. The taste of heavy metals: Gene regulation by MTF-1. Biochim. Biophys. Acta Mol. Cell Res. 2012, 1823, 1416–1425. [CrossRef]
225. Zhang, B.; Georgiev, O.; Hagmann, M.; Günes, C.A.; Cramer, M.; Faller, P.; Vasák, M.; Schaffner, W. Activity of metal-responsive transcription factor 1 by toxic heavy metals and H2O2 in vitro is modulated by metallothionein. Mol. Cell. Biol. 2003, 23, 8471–8485. [CrossRef]
226. Song, C.; Yan, Y.; Rosado, A.; Zhang, Z.; Castellarin, S.D. ABA alleviates uptake and accumulation of zinc in grapevine (Vitis vinifera L.) by inducing expression of ZIP and detoxification-related genes. Front. Plant Sci. 2019, 10, 872. [CrossRef]
227. Ai, T.N.; Naing, A.H.; Yun, B.-W.; Lim, S.H.; Kim, C.K. Overexpression of RsMYB1 enhances anthocyanin accumulation and heavy metal stress tolerance in transgenic petunia. Front. Plant Sci. 2018, 9, 1388. [CrossRef]
228. Sheng, Y.; Yan, X.; Huang, Y.; Han, Y.; Zhang, C.; Ren, Y.; Fan, T.; Xiao, F.; Liu, Y.; Cao, S. The WRKY transcription factor, WRKY13, activates PDR8 expression to positively regulate cadmium tolerance in Arabidopsis. Plant Cell Environ. 2019, 42, 891–903. [CrossRef]
229. Merchant, S.S.; Schmollinger, S.; Strenkert, D.; Moseley, J.L.; Blaby-Haas, C.E. From economy to luxury: Copper homeostasis in Chlamydomonas and other algae. Biochim. Biophys. Acta M Mol. Cell Res. 2020, 1867, 118822. [CrossRef]
230. Sapara, K.K.; Khedia, J.; Agarwal, P.; Gangapur, D.R.; Agarwal, P.K. SbMYB15 transcription factor mitigates cadmium and nickel stress in transgenic tobacco by limiting uptake and modulating antioxidative defence system. Funct. Plant Biol. 2019, 46, 702–714. [CrossRef] [PubMed]
231. Zhang, P.;Wang, R.; Ju, Q.; Li,W.; Tran, L.-S.P.; Xu, J. The R2R3-MYB transcription factor MYB49 regulates cadmium accumulation. Plant Physiol. 2019, 180, 529–542. [CrossRef] [PubMed]
232. Yadav, B.S.; Mani, A. Analysis of bHLH coding genes of Cicer arietinum during heavy metal stress using biological network. Physiol. Mol. Biol. Plants 2019, 25, 113–121. [CrossRef] [PubMed]
233. Xu, Z.; Liu, X.; He, X.; Xu, L.; Huang, Y.; Shao, H.; Zhang, D.; Tang, B.; Ma, H. The soybean basic helix-loop-helix transcription factor ORG3-like enhances cadmium tolerance via increased iron and reduced cadmium uptake and transport from roots to shoots. Front. Plant Sci. 2017, 8, 1098. [CrossRef]
234. Romanenko, K.; Kosakovskaya, I.; Romanenko, P. Phytohormones of microalgae: Biological role and involvement in the regulation of physiological processes. Int. J. Algae 2016, 18, 179–201. [CrossRef]
235. Piotrowska-Niczyporuk, A.; Bajguz, A.; Kotowska, U.; Zambrzycka-Szelewa, E.; Sienkiewicz, A. Auxins and cytokinins regulate phytohormone homeostasis and thiol-mediated detoxification in the green alga Acutodesmus obliquus exposed to lead stress. Sci. Rep. 2020, 10, 10193. [CrossRef]
236. Nguyen, T.Q.; Sesin, V.; Kisiala, A.; Emery, R.N. The role of phytohormones in enhancing metal remediation capacity of algae. Bull. Environ. Contam. Toxicol. 2020, 105, 671–678. [CrossRef]
237. Piotrowska-Niczyporuk, A.; Bajguz, A.; Zambrzycka, E.; Godlewska-Z˙ yłkiewicz, B. Phytohormones as regulators of heavy metal biosorption and toxicity in green alga Chlorella vulgaris (Chlorophyceae). Plant Physiol. Biochem. 2012, 52, 52–65. [CrossRef] [PubMed]
238. Piotrowska-Niczyporuk, A.; Bajguz, A.; Talarek, M.; Bralska, M.; Zambrzycka, E. The effect of lead on the growth, content of primary metabolites, and antioxidant response of green alga Acutodesmus obliquus (Chlorophyceae). Environ. Sci. Pollut. Res. 2015, 22, 19112–19123. [CrossRef]
239. Zhao, Y.; Wang, H.-P.; Han, B.; Yu, X. Coupling of abiotic stresses and phytohormones for the production of lipids and high-value by-products by microalgae: A review. Bioresour. Technol. 2019, 274, 549–556. [CrossRef] [PubMed]
240. Talarek-Karwel, M.; Bajguz, A.; Piotrowska-Niczyporuk, A. Hormonal response of Acutodesmus obliquus exposed to combined treatment with 24-epibrassinolide and lead. J. Appl. Phycol. 2020, 32, 2903–2914. [CrossRef]
241. Bajguz, A.; Piotrowska-Niczyporuk, A. Interactive effect of brassinosteroids and cytokinins on growth, chlorophyll, monosaccharide and protein content in the green alga Chlorella vulgaris (Trebouxiophyceae). Plant Physiol. Biochem. 2014, 80, 176–183. [CrossRef] [PubMed]
242. Vo, T.-T.; Lee, C.; Han, S.-I.; Kim, J.Y.; Kim, S.; Choi, Y.-E. Effect of the ethylene precursor, 1-aminocyclopropane-1-carboxylic acid on different growth stages of Haematococcus pluvialis. Bioresour. Technol. 2016, 220, 85–93. [CrossRef]
243. Raman, V.; Ravi, S. Effect of salicylic acid and methyl jasmonate on antioxidant systems of Haematococcus pluvialis. Acta Physiol. Plant. 2011, 33, 1043–1049. [CrossRef]
244. Kim, S.-H.; Lim, S.R.; Hong, S.-J.; Cho, B.-K.; Lee, H.; Lee, C.-G.; Choi, H.-K. Effect of ethephon as an ethylene-releasing compound on the metabolic profile of Chlorella vulgaris. J. Agric. Food Chem. 2016, 64, 4807–4816. [CrossRef]
245. Thomas, J.C.; Perron, M.; LaRosa, P.C.; Smigocki, A.C. Cytokinin and the regulation of a tobacco metallothionein-like gene during copper stress. Physiol. Plant. 2005, 123, 262–271. [CrossRef]
246. Jiang, L.; Liu, C.; Cao, H.; Chen, Z.; Yang, J.; Cao, S.; Wei, Z. The role of cytokinin in selenium stress response in Arabidopsis. Plant Sci. 2019, 281, 122–132. [CrossRef]
247. Mohan, T.C.; Castrillo, G.; Navarro, C.; Zarco-Fernández, S.; Ramireddy, E.; Mateo, C.; Zamarreño, A.M.; Paz-Ares, J.; Muñoz, R.; García-Mina, J.M. Cytokinin determines thiol-mediated arsenic tolerance and accumulation. Plant Physiol. 2016, 171, 1418–1426.
248. Tian, B.J.; Wang, Y.; Zhu, Y.R.; Lü, X.Y.; Huang, K.; Shao, N.; Beck, C.F. Synthesis of the photorespiratory key enzyme serine: Glyoxylate aminotransferase in C. reinhardtii is modulated by the light regime and cytokinin. Physiol. Plant. 2006, 127, 571–582. [CrossRef]
249. Piotrowska, A.; Czerpak, R. Cellular response of light/dark-grown green alga Chlorella vulgaris Beijerinck (Chlorophyceae) to exogenous adenine- and phenylurea-type cytokinins. Acta Physiol. Plant. 2009, 31, 573–585. [CrossRef]
250. Souza, J.M.; Yokoya, N.S. Effects of cytokinins on physiological and biochemical responses of the agar-producing red alga Gracilaria caudata (Gracilariales, Rhodophyta). J. Appl. Phycol. 2016, 28, 3491–3499. [CrossRef]
251. Udayan, A.; Kathiresan, S.; Arumugam, M. Kinetin and gibberellic acid (GA3) act synergistically to produce high value polyunsaturated fatty acids in Nannochloropsis oceanica CASA CC201. Algal Res. 2018, 32, 182–192. [CrossRef]
252. Du, H.; Ahmed, F.; Lin, B.; Li, Z.; Huang, Y.; Sun, G.; Ding, H.; Wang, C.; Meng, C.; Gao, Z. The effects of plant growth regulators on cell growth, protein, carotenoid, PUFAs and lipid production of Chlorella pyrenoidosa ZF strain. Energies 2017, 10, 1696. [CrossRef]
253. Yu, X.-J.; Sun, J.; Sun, Y.-Q.; Zheng, J.-Y.; Wang, Z. Metabolomics analysis of phytohormone gibberellin improving lipid and DHA accumulation in Aurantiochytrium sp. Biochem. Eng. J. 2016, 112, 258–268. [CrossRef]
254. Sulochana, S.B.; Arumugam, M. Influence of abscisic acid on growth, biomass and lipid yield of Scenedesmus quadricauda under nitrogen starved condition. Bioresour. Technol. 2016, 213, 198–203. [CrossRef] [PubMed]
255. Noble, A.; Kisiala, A.; Galer, A.; Clysdale, D.; Emery, R.N. Euglena gracilis (Euglenophyceae) produces abscisic acid and cytokinins and responds to their exogenous application singly and in combination with other growth regulators. Eur. J. Phycol. 2014, 49, 244–254. [CrossRef]
256. Contreras-Pool, P.Y.; Peraza-Echeverria, S.; Ku-González, Á.F.; Herrera-Valencia, V.A. The phytohormone abscisic acid increases triacylglycerol content in the green microalga Chlorella saccharophila (Chlorophyta). Algae 2016, 31, 267–276. [CrossRef]
257. Bajguz, A.; Piotrowska-Niczyporuk, A. Synergistic effect of auxins and brassinosteroids on the growth and regulation of metabolite content in the green alga Chlorella vulgaris (Trebouxiophyceae). Plant Physiol. Biochem. 2013, 71, 290–297. [CrossRef]
258. Lu, Y.; Xu, J. Phytohormones in microalgae: A new opportunity for microalgal biotechnology? Trends Plant Sci. 2015, 20, 273–282. [CrossRef]
259. Ding, Y.; Ding, L.; Xia, Y.; Wang, F.; Zhu, C. Emerging roles of microRNAs in plant heavy metal tolerance and homeostasis. J. Agric. Food Chem. 2020, 68, 1958–1965. [CrossRef]
260. Noman, A.; Aqeel, M. miRNA-based heavy metal homeostasis and plant growth. Environ. Sci. Pollut. Res. 2017, 24, 10068–10082. [CrossRef] [PubMed]
261. Lou, S.; Zhu, X.; Zeng, Z.;Wang, H.; Jia, B.; Li, H.; Hu, Z. Identification of microRNAs response to high light and salinity that is involved in beta-carotene accumulation in microalga Dunaliella salina. Algal Res. 2020, 48, 101925. [CrossRef]
262. Azaman, S.N.A.; Satharasinghe, D.A.; Tan, S.W.; Nagao, N.; Yusoff, F.M.; Yeap, S.K. Identification and analysis of microRNAs in Chlorella sorokiniana using high-throughput sequencing. Genes 2020, 11, 1131. [CrossRef] [PubMed]
263. Wang, X.; Miao, X.; Chen, G.; Cui, Y.; Sun, F.; Fan, J.; Gao, Z.; Meng, C. Identification of microRNAs involved in astaxanthin accumulation responding to high light and high sodium acetate (NaAC) stresses in Haematococcus pluvialis. Algal Res. 2021, 54, 102179. [CrossRef]
264. Srivastava, S.; Srivastava, A.K.; Suprasanna, P.; D’souza, S. Identification and profiling of arsenic stress-induced microRNAs in Brassica juncea. J. Exp. Bot. 2013, 64, 303–315. [CrossRef]
265. Remans, T.; Opdenakker, K.; Guisez, Y.; Carleer, R.; Schat, H.; Vangronsveld, J.; Cuypers, A. Exposure of Arabidopsis thaliana to excess Zn reveals a Zn-specific oxidative stress signature. Environ. Exp. Bot. 2012, 84, 61–71. [CrossRef]
266. Huang, S.Q.; Xiang, A.L.; Che, L.L.; Chen, S.; Li, H.; Song, J.B.; Yang, Z.M. A set of miRNAs from Brassica napus in response to sulphate deficiency and cadmium stress. Plant Biotechnol. J. 2010, 8, 887–899. [CrossRef]
267. Liu, Q.; Zhang, H. Molecular identification and analysis of arsenite stress-responsive miRNAs in rice. J. Agric. Food Chem. 2012, 60, 6524–6536. [CrossRef]
268. Zhou, Z.S.; Song, J.B.; Yang, Z.M. Genome-wide identification of Brassica napus microRNAs and their targets in response to cadmium. J. Exp. Bot. 2012, 63, 4597–4613. [CrossRef]
269. Pilon, M. The copper microRNAs. New Phytol. 2017, 213, 1030–1035. [CrossRef]
270. Ding, Y.-F.; Zhu, C. The role of microRNAs in copper and cadmium homeostasis. Biochem. Biophys. Res. Commun. 2009, 386, 6–10. [CrossRef] [PubMed]
271. Mendoza-Soto, A.B.; Sánchez, F.; Hernández, G. MicroRNAs as regulators in plant metal toxicity response. Front. Plant Sci. 2012, 3, 105. [CrossRef]
272. Paul, S.; Datta, S.K.; Datta, K. miRNA regulation of nutrient homeostasis in plants. Front. Plant Sci. 2015, 6, 232. [CrossRef] [PubMed]
273. Jagadeeswaran, G.; Li, Y.F.; Sunkar, R. Redox signaling mediates the expression of a sulfate-deprivation-inducible micro RNA 395 in Arabidopsis. Plant J. 2014, 77, 85–96. [CrossRef] [PubMed]
274. Ding, Y.; Chen, Z.; Zhu, C. Microarray-based analysis of cadmium-responsive microRNAs in rice (Oryza sativa). J. Exp. Bot. 2011, 62, 3563–3573. [CrossRef] [PubMed]
275. Zhang, L.W.; Song, J.B.; Shu, X.X.; Zhang, Y.; Yang, Z.M. miR395 is involved in detoxification of cadmium in Brassica napus. J. Hazard. Mater. 2013, 250, 204–211. [CrossRef]
276. Zeng, Q.-Y.; Yang, C.-Y.; Ma, Q.-B.; Li, X.-P.; Dong, W.-W.; Nian, H. Identification of wild soybean miRNAs and their target genes responsive to aluminum stress. BMC Plant Biol. 2012, 12, 1–16. [CrossRef]
277. Chen, L.; Wang, T.; Zhao, M.; Tian, Q.; Zhang, W.-H. Identification of aluminum-responsive microRNAs in Medicago truncatula by genome-wide high-throughput sequencing. Planta 2012, 235, 375–386. [CrossRef]
278. Halder, S. Bioremediation of heavy metals through fresh water microalgae: A review. Sch. Acad. J. Biosci. 2014, 2, 825–830.
279. Wang, Y.; Selvamani, V.; Yoo, I.-K.; Kim, T.W.; Hong, S.H. A novel strategy for the microbial removal of heavy metals: Cell-surface display of peptides. Biotechnol. Bioprocess Eng. 2021, 26, 1–9. [CrossRef]
280. Yang, T.; Chen, M.-L.; Wang, J.-H. Genetic and chemical modification of cells for selective separation and analysis of heavy metals of biological or environmental significance. Trends Anal. Chem. 2015, 66, 90–102. [CrossRef]
281. Kuroda, K.; Ueda, M. Molecular design of the microbial cell surface toward the recovery of metal ions. Curr. Opin. Biotechnol. 2011, 22, 427–433. [CrossRef] [PubMed]
282. Kuroda, K.; Shibasaki, S.; Ueda, M.; Tanaka, A. Cell surface-engineered yeast displaying a histidine oligopeptide (hexa-His) has enhanced adsorption of and tolerance to heavy metal ions. Appl. Microbiol. Biotechnol. 2001, 57, 697–701. [CrossRef] [PubMed]
283. Kuroda, K.; Ueda, M.; Shibasaki, S.; Tanaka, A. Cell surface-engineered yeast with ability to bind, and self-aggregate in response to copper ion. Appl. Microbiol. Biotechnol. 2002, 59, 259–264.
284. Kuroda, K.; Ueda, M. Bioadsorption of cadmium ion by cell surface-engineered yeasts displaying metallothionein and hexa-His. Appl. Microbiol. Biotechnol. 2003, 63, 182–186. [CrossRef] [PubMed]
285. Kuroda, K.; Ueda, M. Effective display of metallothionein tandem repeats on the bioadsorption of cadmium ion. Appl. Microbiol. Biotechnol. 2006, 70, 458–463. [CrossRef] [PubMed]
286. Kotrba, P.; Ruml, T. Surface display of metal fixation motifs of bacterial P1-type ATPases specifically promotes biosorption of Pb2+ by Saccharomyces cerevisiae. Appl. Environ. Microbiol. 2010, 76, 2615–2622. [CrossRef] [PubMed]
287. Bae, W.; Wu, C.H.; Kostal, J.; Mulchandani, A.; Chen, W. Enhanced mercury biosorption by bacterial cells with surface-displayed MerR. Appl. Environ. Microbiol. 2003, 69, 3176–3180. [CrossRef] [PubMed]
288. Sousa, C.; Kotrba, P.; Ruml, T.; Cebolla, A.; de Lorenzo, V. Metalloadsorption by Escherichia coli cells displaying yeast and mammalian metallothioneins anchored to the outer membrane protein LamB. J. Bacteriol. 1998, 180, 2280–2284. [CrossRef]
289. Krishnaswamy, R.; Wilson, D.B. Construction and characterization of an Escherichia coli strain genetically engineered for Ni(II) bioaccumulation. Appl. Environ. Microbiol. 2000, 66, 5383–5386. [CrossRef] [PubMed]
290. Shen, N.; Birungi, Z.S.; Chirwa, E. Selective biosorption of precious metals by cell-surface engineered microalgae. Chem. Eng. Trans. 2017, 61, 25–30.
291. Kuroda, K.; Ebisutani, K.; Iida, K.; Nishitani, T.; Ueda, M. Enhanced adsorption and recovery of uranyl ions by NikR mutant-displaying yeast. Biomolecules 2014, 4, 390–401. [CrossRef] [PubMed]
292. Nishitani, T.; Shimada, M.; Kuroda, K.; Ueda, M. Molecular design of yeast cell surface for adsorption and recovery of molybdenum, one of rare metals. Appl. Microbiol. Biotechnol. 2010, 86, 641–648. [CrossRef] [PubMed]
293. Kuroda, K.; Nishitani, T.; Ueda, M. Specific adsorption of tungstate by cell surface display of the newly designed ModE mutant. Appl. Microbiol. Biotechnol. 2012, 96, 153–159. [CrossRef] [PubMed]
294. Fakhrullin, R.F.; Lvov, Y.M.; Choi, I.S. Future of cell surface engineering. In Cell Surface Engineering, 1st ed.; Royal Society of Chemistry: London, UK, 2014; pp. 240–245.