Next Article in Journal
Identification of Black Rot Resistance in a Wild Brassica Species and Its Potential Transferability to Cauliflower
Previous Article in Journal
Role of Biocontrol Agents in Management of Corm Rot of Saffron Caused by Fusarium oxysporum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ionic Selenium and Nanoselenium as Biofortifiers and Stimulators of Plant Metabolism

by
Víctor García Márquez
,
Álvaro Morelos Moreno
,
Adalberto Benavides Mendoza
and
Julia Medrano Macías
*
Horticulture Departament, Agrarian Autonomous University Antonio Narro, Calzada Antonio Narro 1923, Saltillo 25315, Mexico
*
Author to whom correspondence should be addressed.
Agronomy 2020, 10(9), 1399; https://doi.org/10.3390/agronomy10091399
Submission received: 5 August 2020 / Revised: 9 September 2020 / Accepted: 11 September 2020 / Published: 16 September 2020
(This article belongs to the Section Horticultural and Floricultural Crops)

Abstract

:
Selenium (Se) is an essential element in mammals; however, there is frequently an insufficient intake due to several factors. Different techniques have been used to deal with this problem, such as plant biofortification with Se in its ionic forms and, more recently, at the nanoscale. Additionally, despite the fact that Se is not considered an essential element in plants, it has been shown to stimulate (through still unknown mechanisms) plant metabolism, causing an increase in the synthesis of molecules with reducing power, including enzymes such as glutathione peroxidase, catalase and ascorbate peroxidase as well as non-enzymatic antioxidants such as phenolic compounds, glucosinolates, vitamins and chlorophylls. A positive correlation has also been shown with other essential elements, achieving an increase in tolerance to environmental adversities. This article describes the advances made in the field of the biofortification of horticultural crops with ionic Se and nanoselenium (nSe) from 2009 to 2019. The aspects covered include various concentrations used, the findings made regarding the impact these chemical forms have on plant metabolism, and indications of its participation in the synthesis of primary and secondary metabolites that increase stress tolerance.

Graphical Abstract

1. Introduction

Selenium (Se) is an essential trace element in humans and other vertebrates. It fulfills various functions, especially those related to the synthesis of selenoproteins, and also includes antioxidant and anti-inflammatory properties, immune system and antiviral effects, cancer prevention, brain development and cognitive aspects [1]. The minimum daily requirement for Se in humans was established by the Recommended Daily Allowances (RDA) and the World Health Organization (WHO), and the optimal range is between 50 and 60 µg per day to reduce health risks. Currently, it is known that Se intake is less than the minimum daily requirement for most populations due to the shortage of this element in crop food sources, mainly due to the low concentration of Se in the soil. The soil Se concentration has been established as being between 0.01 and 2 mg kg−1, with 0.4 mg kg−1 being the global mean [2]. To deal with this problem, various Se biofortification strategies for different crops have been used, such as exogenous incorporation in ionic forms [3,4] and more recently in the form of nanoparticles [5,6], and encouraging results have been obtained to mitigate such deficiency.
During the exploration of Se biofortification, beneficial effects on plant metabolism have been found, including an increase in molecules with reducing power leading to an increase in stress tolerance [7], an increase in plant growth and fruits with greater nutraceutical value, among others [8]. The objective of this review is to analyze, compare and conclude results obtained in the last decade (2009–2019) regarding ionic and nanoparticulated Se biofortification as well as metabolic stimulations, mainly in horticultural plants.

2. Selenium in Human Health

Selenium (Se) was discovered in 1817 by Berzelius and was considered a highly toxic and polluting element. It was not until after the 1950s that functions essential to human and animal health were attributed to Se [9], including antioxidant activity [10,11], hormonal regulation such as in the thyroid [12,13], anticancer effects [14,15], cardiovascular protection [16,17] and antiviral properties [18,19]. Nowadays, various review articles have shown and updated the latest findings on the role of Se in organic functioning [1,20,21,22,23], where such functions are carried out through the synthesis of 25 known selenoproteins derived from selenocysteine (SeCys) [24,25]. The function of most of these proteins remains unknown, but a group of these selenoproteins was shown to have oxidoreduction catalytic activity, where SeCys is located in the active site because it is more reactive than cysteine (Cys) under physiological conditions; thus, SeCys can exist as a nucleophile without electrostatic interactions and can enhance the catalytic effectiveness [26]. The enzymes related to the oxidoreduction catalytic activity include five glutathione peroxidases, which catalyze the reduction of peroxide in the presence of glutathione; three thioredoxin reductases, which reduce thioredoxin or other proteins in the presence of NADPH; three deiodinases in the thyroid, which promote the reductive de-iodization of thyroid hormones; methionine-R-sulfoxide reductase, which reduces oxidized methionine residues within proteins; selenophosphate synthase, which catalyzes the synthesis of selenophosphate and is ATP dependent. On the other hand, there is a group of selenoproteins that do not have catalytic functions, among which are K, a simple protein with a transmembrane domain; O, the largest selenoprotein; H, related to glutathione gene expression (GSH); I, a membrane protein; T, containing a predicted redox motif. The W and V proteins are related at the C-terminal, M and Sep 15 are distant homologs proteins, and the functions of the S, R and N proteins remain even more unknown [27,28].
The daily intake of Se is established based on the concentration of selenoprotein P (SepP) in the blood serum, since it has been shown to be the most effective marker. A study in one province of China indicated that Se deficiency induced SepP saturation by consuming 49 µg of Se per day in people weighing 58 kg [29], which corresponds to approximately 1 µg of Se per kg of body weight. Therefore, the reference values were stipulated as follows: daily intake of 70 µg of Se for men with an average weight of 70 kg, and 60 µg of Se for adult women with an average weight of 60 kg, but this could increase up to 75 µg during pregnancy. Table 1 shows the Se intake specifications according to age and condition [30]. Additionally, it has been shown that the human body can tolerate large amounts of this element; however, amounts near 400 µg of Se per day show signs of toxicity [31].
It is estimated that between 500 million and 1 billion people around the world have a deficient intake of Se, mainly due to the scarcity of this element in the soil since plants are the main source of Se in the diet [32]. The world average of Se in the soil is 0.4 mg kg−1; however, the availability of this nutrient depends on the physicochemical properties of the soil, such as the redox state, pH, content of organic matter and microbiota [33,34].
To deal with this problem, one solution has been to biofortify Se in staple crops for human consumption, and evaluations have included using different Se chemical forms, concentrations, doses and application times; application directly to soil, foliar spraying, hydroponics and direct application to the seed and the nude root by imbibition. Table 2 and Table 3 show some prominent results obtained in the last 10 years on the topic of biofortification with ionic Se and nanoparticle Se (nSe).

3. Results and Discussion

3.1. Biofortification with Ionic Se and nSe in Crops

3.1.1. Ionic Selenium

Se biofortification in crops for human consumption has yielded encouraging results, including from the nutraceutical perspective to the increase in metabolites that enhance tolerance to adverse factors. Interesting results have been found with foliar application to crops, such as cereals, where both Na2SeO3 and Na2SeO4 work in a range between 30 and 300 g ha−1. In this range, it is possible to find an increase in the accumulation of Se in grains and stimulation of metabolism. On the other hand, to obtain the same result in vegetables, it is necessary to apply the aforementioned chemical species between 1 and 20 mg L−1 once or twice per cycle.
It has been found that the application of Se in the form of sodium selenite (Na2SeO3) between 0.86 and 5 mg L−1, as well as sodium selenate (Na2SeO4) in a range of 0.5 to 18 mg L−1, in hydroponic cultures increases the accumulation of this element and the antioxidants in the edible parts of species such as lettuce (Latuca sativa), tomato (Solanum lycopersicon L.), strawberry (Fragaria x ananassa) and basil (Ocinum basilicum).
Similar encouraging results were shown, regarding accumulation and stimulation linked to metabolism, for Se application directly to the soil with both sodium selenite (Na2SeO3) or sodium selenate in cereals from 10 to 100 g ha−1 and vegetables in the range of 20 to 940 mg L−1.
The 24 h seed imbibition technique using Na2SeO4 at concentrations ranging from 0.1 to 8.54 mg L−1 has been found to be both easy to perform and convenient due to its accumulation and stimulation of the metabolism.

3.1.2. Selenium Nanoparticles (nSe)

Recently, nanotechnology has revolutionized a wide range of areas such as pharmaceuticals, energy, communications engineering, medicine, the environment, chemical industry and plant nutrition, among others, due to the low area/surface ratio, states of aggregation and stability of nanoparticles. The synthesis and use of nSe as a nutrient and biofortifier has proven to be an interesting strategy because, additionally, it has been found to possess greater chemical stability, biocompatibility, rapid absorption and less toxicity compared to the ionic forms of this element [78,79]. As can be seen in Table 3, foliar applications range from 10 to 50 mg L−1 in species such as coriander (Coriandum sativum), pomegranate (Punica granatum), peanut (Archis hypogea) and wheat (Triticum aestivum).
Different studies have been carried out by imbibing seeds or bare roots at nSe concentrations ranging from 1.18 to 50 mg L−1; results have shown no changes in growth, but increases in Se in the edible parts and antioxidant content. In addition, various works have been carried out where nanoparticles are applied directly to the substrate in ranges as wide as 10 to 100 ppm between horticultural species.
Information regarding nSe application in biofortification and improvement in plant metabolism (e.g., the appropriate concentrations in hydroponic or soil cultures) is still lacking.
In the same way, the mechanism by which these nanoparticles are absorbed is still not fully elucidated. One of the widely accepted options is that absorption happens both intra- and extracellularly through the tissues, until reaching the xylem; the way in which nanoparticles pass through the casparian strip is not yet clear, but it could be through the meristematic zone. The cell wall acts as a physical barrier; however, it contains pores with diameters between 5 and 20 nm, and nanoparticles smaller than this will enter freely [80]. It is also possible that nanoparticles greater than 20 nm enlarge the pores, inducing the formation of cavities to enter via endocytosis or even through transmembrane proteins or ionic channels [81].
As described above, to date, encouraging results have been obtained in the field of biofortification with ionic selenium in horticultural species widely consumed by humans, such as lettuce and tomato, as well as in the stimulation of redox metabolism, which leads to an increase in tolerance to adverse factors. However, a promising alternative is the use of selenium nanoparticles, where a reduction in application complexity may be achieved, and this leads to important results in the potentiation of antioxidant metabolism, the promotion of agronomic sustainability and a reduction in waste. Therefore, more research should be carried out on the plant cell level as well as interactions within the entire trophic chain and environment [82].

3.2. Stimulation of Plant Metabolism

It has been shown that the application of selenium in plants, via foliar, hydroponic cultures, directly to the soil or by imbibition, stimulates plant metabolism, resulting in improved growth, improved synthesis of molecules involved in defense [83] and an increase in stress tolerance [84]. However, the exact mechanics under which these processes occur are unclear. The main pathway of selenium’s impact will be pointed out below, when it is applied in ionic form and in nanoparticles.

3.2.1. Ionic Selenium

Several research works have been carried out to demonstrate the effect of selenium on plant metabolism, more specifically the impact on phytochemicals, and the results have been quite varied. In a relatively constant way, it has been established that low concentrations of Se can act as an antioxidant and high concentrations as a pro-oxidant [85]. However, establishing the values of “high” or “low” concentrations is difficult. According to what is shown in Table 2, each form of application tolerates different selenium concentration ranges, taking into account the following pattern: imbibition > hydroponics > foliar > soil.
The application of Se and its impact on antioxidant metabolism in plants have been linked to both primary and secondary metabolites.

Primary Metabolites

Regarding primary metabolites, there is an increase in the activity of glutathione peroxidase (GPX), an enzyme that involves selenic acid (PSeOH) in its catalytic cycle. This reacts with glutathione, a tripeptide that functions as a coenzyme in this reaction and contains a sulfhydryl functional group (-SH), to form a selenyl-sulfide adduct. At this point, a second GSH molecule intervenes, and here selenol (PSeH) is formed in the active site where peroxide reduction takes place [86]. In addition, it has been reported that the application of selenium promotes an increase in the activity of enzymes with an antioxidant capacity, such as ascorbate peroxidase, catalase, superoxide dismutase, dehydroascorbate reductase, glutathione reductase and monodehydroascorbate reductase [87].
Although not as thoroughly studied, it has also been related to the increase in other non-catalytic antioxidant proteins, such as thioredoxin (TrxR) and P protein, the latter containing more than 10 Se atoms [88].
In a study carried out in rapeseed (Brassica napus), important effects on primary metabolism were elucidated, including a higher concentration of glucose, coupled with higher ATP production; increased superoxide dismutase activity in the mitochondria and potentiation in the pentose pathway phosphate, which supplies a large number of non-enzymatic antioxidants; as well as a reduction in the tricarboxylic acid (TCA) cycle [89].
An increase has also been found in the synthesis of sulfur amino acids (Cys and Met) and selenoamino acids such as SeCys and Semet, which are incorporated into proteins. However, there is also evidence of other non-protein amino acids being synthesized, such as γ-glutamyl methyl seleniocysteine (γ-gluMetSeCys), methyl-SeCys and methyl-Semet, mainly in hyperaccumulator families of S, such as Brassica and Allium. In addition, these metabolites have shown powerful anticancer activities [47,52]. Within this same group of plants, volatile species such as dimethyl-diselenide (DMDS) are produced, which partly controls selenium accumulation. In non-accumulating plant species (those that accumulate < 100 mg Se per kg DW), the methylated species of selenium, dimethyl selenide (DMSe), is synthesized [90].
It is known that the presence of Se promotes increases of sulfur transporters sultr 1 and 2 as well as ATP sulfurylase at the transcriptomic level, which leads to greater absorption of sulfur and, therefore, the synthesis of both primary and secondary metabolites that contain these elements [91]. However, it is important to highlight the balance between Se and S concentrations. It has been found that high Se and low S concentrations promote competition, avoiding adequate absorption of sulfur and, in this case, a reduction in the synthesis of the mentioned metabolites [92].
In a study carried out in rice (Oryza sativum), a three-fold increase in fatty acids (oleic, linoleic and linolenic) was found after the application of selenite and selenite; however, the metabolic pathway that was affected is not clear [60].

Secondary Metabolites

After the application of ionic species of selenium, both positive and negative effects have been reported on the concentration of secondary sulfur metabolites such as glutathione and glucosinolates. For the first, examples of increased concentrations were evidenced in radish (Raphanus sativum L.) [55] and plum trees (Prunus domestica) [93]. On the other hand, the reduction in GSH found in strawberry plants was associated with an increase in the concentration of its oxidized form: glutathione disulfide (GSSH) [94].
Various S hyperaccumulating plants, such as broccoli (Brassica oleracea), show increased synthesis of glucosinolates with exogenous application of Se, which participate in defense against herbivores and are synthesized mainly from methionine and phenylalanine [95].
There is a close relationship between the metabolism of sulfur and nitrogen; the proportion of these in most plant species is preserved between 1:45 and 1:30, respectively. About 80% of S and N assimilation is directed to the production of protein amino acids. After modifications in the absorption of S due to the aforementioned effect on transporters with the application of Se [96], the content of metabolites dependent on these elements, such as glucosinolates, is also affected.
It has also been found that selenium modifies the phenylpropanoid pathway, increasing the enzymatic activity of phenylalanine ammonium lyase (PAL), which is why several investigations have found a positive correlation with phenolic compounds, which act as non-enzymatic antioxidants [49]. Similarly, an increase in the synthesis of ascorbic acid has been found, which functions as a direct scavenger of reactive oxygen species and a cofactor of enzymes with antioxidant activity, such as APX [97].
In summary, selenium stimulates metabolism in two ways: (1) via antioxidants, where it participates in the catalytic cycle of enzymes such as APX, and (2) via pro-oxidants, where selenite and selenate probably mimic moderate oxidative stress and the detection of synthesized reactive species will trigger signaling to achieve the formation of all antioxidant machinery [98].

3.2.2. Selenium Nanoparticles (nSe)

Unlike readily available research on Se applications in its ionic form, research on its nanoparticulate (nSe) form is much more recent and scarcer. Although results have been favorable, even more than that observed in the ionic form, there is still a long way to go. It is generally known that nSe have a strong impact on antioxidant metabolism, which is why they have probably been successfully tested in various species to cope with different types of stress. Examples of this are reported in sorghum (Sorghum bicolor) subjected to high temperatures [65]; strawberry, tomato, coriander, basil and barley (Hordeum vulgare) under stress by salinity [57,62,76,77,78]; increase in tolerance to stress caused by pathogens such as Alternaria solani [79], Meloidogyne incognita [80] and Botrytis cinearea [99]. The reason why Se can trigger these beneficial effects could be related to a change in the redox status of the cell, causing a greater stimulation in the synthesis of non-enzymatic antioxidants such as lycopene and carotenoids. It could also be related to an increase in the activity of the main enzymes involved in the antioxidant pathway such as glutathione peroxidase, where selenium acts as a cofactor, and it is possible to obtain ionic selenium from nanoparticles [66]. A reduction in the loss of photosynthetic pigments such as chlorophyll a and b has also been noted during adverse conditions, which leads to an improvement in the photosynthetic rate, maintenance of homeostasis and improvement in growth [62]. This can be explained, at least partially, by nSe being stored in the thylakoid membranes, providing them with stability. A reduction in lipoperoxidation has also been proposed as protecting and stabilizing some enzymes that catalyze the synthesis of these photosynthetic pigments [6].

4. Conclusions

This review addressed the doses and forms of application of selenium ionic (Na2SeO3 and Na2SeO4) and nanoparticulate in the main crops, for biofortification purposes as well as the implications in the synthesis of molecules with reducing power that entails an increase in nutraceutical quality and stress tolerance.

5. Outlook

When the available information regarding ionic Se and nanoselenium (nSe) applications in different crops is analyzed in parallel, it can be seen that there is a need to address the missing information in order to more precisely understand the absorption, assimilation and metabolism of both selenium forms.
It is necessary to test the Se nanoparticle application more extensively, especially in crops widely consumed by humans, such as vegetables. The recommended form of application based on a cost–benefit analysis is the foliar route at concentrations less than 10 mg L−1 and monitoring the genomic and metabolomic behaviors around this process.

Author Contributions

Conceptualization, J.M.M. and A.B.M.; methodology, V.G.M., Á.M.M., A.B.M. and J.M.M.; formal analysis, J.M.M. and Á.M.M.; investigation, V.G.M., Á.M.M., Á.B.M. and J.M.M.; data curation, J.M.M., A.B.M. and Á.M.M.; writing—original draft preparation, J.M.M., V.G.M., A.B.M. and J.M.M.; writing—review and editing, Á.M.M., A.B.M., and J.M.M.; supervision, J.M.M. and A.B.M.; project administration, A.B.M., Á.M.M. and J.M.M., funding acquisition, A.B.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work was carried out with the support of the National Council of Science and Technology (CONACyT) through the granting of the Master’s Scholarship No. 714021.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roman, M.; Jitaru, P.; Barbante, C. Selenium biochemistry and its role for human health. Metallomics 2014, 6, 25–54. [Google Scholar] [CrossRef] [PubMed]
  2. Winkel, L.H.E.; Vriens, B.; Jones, G.D.; Schneider, L.S.; Pilon-Smits, E.; Bañuelos, G.S. Selenium Cycling Across Soil-Plant-Atmosphere Interfaces: A Critical Review. Nutrients 2015, 7, 4199–4239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Malagoli, M.; Schiavon, M.; Dall’Acqua, S.; Pilon-Smits, E.A.H. Effects of selenium biofortification on crop nutritional quality. Front. Plant Sci. 2015, 6, 280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Zhu, Y.G.; Pilon-Smits, E.A.H.; Zhao, F.J.; Williams, P.N.; Meharg, A.A. Selenium in higher plants: Understanding mechanisms for biofortification and phytoremediation. Trends Plant Sci. 2009, 14, 436–442. [Google Scholar] [CrossRef]
  5. El-Ramady, H.; Abdalla, N.; Alshaal, T.; El-Henawy, A.; Elmahrouk, M.; Bayoumi, Y.; Shalaby, T.; Amer, M.; Shehata, S.; Fári, M.; et al. Plant Nano-Nutrition: Perspectives and Challenges; Springer: Cham, Switzerland, 2018; ISBN 9783319701660. [Google Scholar]
  6. Hussein, H.A.A.; Darwesh, O.M.; Mekki, B.B. Environmentally friendly nano-selenium to improve antioxidant system and growth of groundnut cultivars under sandy soil conditions. Biocatal. Agric. Biotechnol. 2019, 18, 101080. [Google Scholar] [CrossRef]
  7. Paciolla, C.; de Leonardis, S.; Dipierro, S. Effects of selenite and selenate on the antioxidant systems in Senecio scandens L. Plant Biosyst. 2011, 145, 253–259. [Google Scholar] [CrossRef]
  8. Mohamed Zeid, I.; El Lateef Gharib, F.A.; Mohamed Ghazi, S.; Zakaria Ahmed, E. Promotive Effect of Ascorbic Acid, Gallic Acid, Selenium and Nano-Selenium on Seed Germination, Seedling Growth and Some Hydrolytic Enzymes Activity of Cowpea (Vigna unguiculata) Seedling. J. Plant Physiol. Pathol. 2019, 7, 1–8. [Google Scholar] [CrossRef]
  9. Schwars, K.; Folts, C. Factor 3 activity of selenium compounds. J. Biol. Chem. 1958, 233, 245–251. [Google Scholar] [CrossRef]
  10. Steinbrenner, H.; Speckmann, B.; Klotz, L.O. Selenoproteins: Antioxidant selenoenzymes and beyond. Arch. Biochem. Biophys. 2016, 595, 113–119. [Google Scholar] [CrossRef] [PubMed]
  11. Zoidis, E.; Seremelis, I.; Kontopoulos, N.; Danezis, G.P. Selenium-dependent antioxidant enzymes: Actions and properties of selenoproteins. Antioxidants 2018, 7, 66. [Google Scholar] [CrossRef] [Green Version]
  12. Ambrosio, R.; De Stefano, M.A.; Di Girolamo, D.; Salvatore, D. Thyroid hormone signaling and deiodinase actions in muscle stem/progenitor cells. Mol. Cell. Endocrinol. 2017, 459, 79–83. [Google Scholar] [CrossRef] [PubMed]
  13. Rayman, M.P.; Duntas, L.H. Selenium Deficiency and Thyroid Disease. Thyroid Dis. 2019, 4, 109–126. [Google Scholar] [CrossRef]
  14. Kuršvietienė, L.; Mongirdienė, A.; Bernatonienė, J.; Šulinskienė, J.; Stanevičienė, I. Selenium anticancer properties and impact on cellular redox status. Antioxidants 2020, 9, 80. [Google Scholar] [CrossRef] [Green Version]
  15. Spallholz, J.E. Selenomethionine and Methioninase: Selenium Free Radical Anticancer Activity. Methods Mol. Biol. 2019, 1866, 199–210. [Google Scholar] [CrossRef] [PubMed]
  16. Alehagen, U.; Johansson, P.; Björnstedt, M.; Rosén, A.; Post, C.; Aaseth, J. Relatively high mortality risk in elderly Swedish subjects with low selenium status. Eur. J. Clin. Nutr. 2016, 70, 91–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Zhang, X.; Liu, C.; Guo, J.; Song, Y. Selenium status and cardiovascular diseases: Meta-analysis of prospective observational studies and randomized controlled trials. Eur. J. Clin. Nutr. 2016, 70, 162–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Li, Y.; Lin, Z.; Guo, M.; Zhao, M.; Xia, Y.; Wang, C.; Xu, T.; Zhu, B. Inhibition of H1N1 influenza virus-induced apoptosis by functionalized selenium nanoparticles with amantadine through ROS-mediated AKT signaling pathways. Int. J. Nanomed. 2018, 13, 2005–2016. [Google Scholar] [CrossRef] [Green Version]
  19. Zhang, J.; Will, E.T.; Bennett, K.; Saad, R.; Rayman, M.P. Association between regional selenium status and reported outcome of COVID-19 cases in China. Am. J. Clin. Nutr. 2020, 111, 1297–1299. [Google Scholar] [CrossRef]
  20. Esmaeili, S.; Fazelifard, R.S.; Ahmadzadeh, S.; Shokouhi, M. The influence of Selenium on human health. KAUMS J. 2013, 16, 779–780. [Google Scholar]
  21. Köhrle, J. Selenium and the thyroid. Curr. Opin. Endocrinol. Diabetes Obes. 2015, 22, 392–401. [Google Scholar] [CrossRef]
  22. Rayman, M.P. Selenium intake, status, and health: A complex relationship. Hormones 2020, 19, 9–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Vinceti, M.; Filippini, T.; Wise, L.A. Environmental Selenium and Human Health: An Update. Curr. Environ. Health Rep. 2018, 5, 464–485. [Google Scholar] [CrossRef] [PubMed]
  24. Kryukov, G.V.; Castellano, S.; Novoselov, S.V.; Lobanov, A.V.; Zehtab, O.; Guigó, R.; Gladyshev, V.N. Characterization of mammalian selenoproteomes. Science 2003, 300, 1439–1443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Santesmasses, D.; Mariotti, M.; Gladyshev, V.N. Bioinformatics of Selenoproteins. Antioxid. Redox Signal. 2020, 2. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Roh, Y.J.; Han, S.; Park, I.; Lee, H.M.; Ok, Y.S. Role of Selenoproteins in Redox Regulation of Signaling and the Antioxidant System: A Review. Antioxidants 2020, 9, 383. [Google Scholar] [CrossRef] [PubMed]
  27. Hatfield, D.L.; Tsuji, P.A.; Carlson, B.A.; Gladyshev, V.N. Selenium and selenocysteine: Roles in cancer, health and development. Trends Biochem. Sci. 2014, 39, 112–120. [Google Scholar] [CrossRef] [Green Version]
  28. Lobanov, A.V.; Hatfield, D.L.; Gladyshev, V.N. Eukaryotic selenoproteins and selenoproteomes Alexey. Biochim. Biophys. Acta-Bioenerg. 2009, 1790, 1424–1428. [Google Scholar] [CrossRef] [Green Version]
  29. Xia, Y.; Hill, K.E.; Li, P.; Xu, J.; Zhou, D.; Motley, A.K.; Wang, L.; Byrne, D.W.; Burk, R.F. Optimization of selenoprotein P and other plasma selenium biomarkers for the assessment of the selenium nutritional requirement: A placebo-controlled, double-blind study of selenomethionine supplementation in selenium-deficient Chinese subjects. Am. J. Clin. Nutr. 2010, 92, 525–531. [Google Scholar] [CrossRef] [Green Version]
  30. Kipp, A.; Brigelius-Flohé, R.; Schomburg, L.; Bechthold, A.; Leschink-Bonnet, E.; Heseker, H. revised reference values for selenium intake. J. Trace. Elem. Med. Biol. 2015, 195–199. [Google Scholar] [CrossRef] [Green Version]
  31. Haytowitz, D.; Lemar, L.E.; Pehrsson, P.R.; Exler, J.; Patterson, K.K.; Thomas, R.G.; Duvall, M. USDA Review of USDA National Nutrient Database for Standard Reference, Release 24 and Dietary Supplement Ingredient Database, Release 2. J. Agric. Food Inf. 2012, 13, 358–359. [Google Scholar] [CrossRef]
  32. Kieliszek, M. Selenium–fascinating microelement, properties and sources in food. Molecules 2019, 24, 1298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Dinh, Q.; Wang, M.; Tran, T.; Zhou, F.; Wang, D.; Zhai, H.; Peng, Q.; Xue, M.; Du, Z.; Bañuelos, G.; et al. Bioavailability of selenium in soil-plant system and a regulatory approach. Crit. Rev. Environ. Sci. Technol. 2019, 49, 443–517. [Google Scholar] [CrossRef]
  34. Li, Z.; Liang, D.; Peng, Q.; Cui, Z.; Huang, J.; Lin, Z. Interaction between selenium and soil organic matter and its impact on soil selenium bioavailability: A review. Geoderma 2017, 295, 69–79. [Google Scholar] [CrossRef]
  35. Ríos, J.J.; Blasco, B.; Cervilla, L.M.; Rosales, M.A.; Sanchez-Rodriguez, E.; Romero, L.; Ruiz, J.M. Production and detoxification of H2O2in lettuce plants exposed to selenium. Ann. Appl. Biol. 2009, 154, 107–116. [Google Scholar] [CrossRef]
  36. Malorgio, F.; Diaz, K.E.; Ferrante, A.; Mensuali-Sodi, A.; Pezzarossa, B. Effects of selenium addition on minimally processed leafy vegetables grown in a floating system. J. Sci. Food Agric. 2009, 89, 2243–2251. [Google Scholar] [CrossRef]
  37. Ramos, S.J.; Faquin, V.; Guilherme, L.R.G.; Castro, E.M.; Ávila, F.W.; Carvalho, G.S.; Bastos, C.E.A.; Oliveira, C. Selenium biofortification and antioxidant activity in lettuce plants fed with selenate and selenite. Plant Soil Environ. 2010, 56, 584–588. [Google Scholar] [CrossRef] [Green Version]
  38. Leija-Martínez, P.; Benavides-Mendoza, A.; Cabrera-De La Fuente, M.; Robledo-Olivo, A.; Ortega-Ortíz, H.; Sandoval-Rangel, A.; González-Morales, S. Lettuce biofortification with selenium in chitosan-polyacrylic acid complexes. Agronomy 2018, 8, 275. [Google Scholar] [CrossRef] [Green Version]
  39. Becvort-Azcurra, A.; Fuentes-Lara, L.; Benavides-Mendoza, A.; Ramírez, H.; Robledo-Torres, V.; Rodríguez-Mendoza, M.D.L.N. Aplicación de selenio en tomate: Crecimiento productividad y estado antioxidante del fruto. Terra Latinoam. 2012, 30, 291–301. [Google Scholar]
  40. Businelli, D.; D’Amato, R.; Onofri, A.; Tedeschini, E.; Tei, F. Se-enrichment of cucumber (Cucumis sativus L.), lettuce (Lactuca sativa L.) and tomato (Solanum lycopersicum L. Karst) through fortification in pre-transplanting. Sci. Hortic. 2015, 197, 697–704. [Google Scholar] [CrossRef]
  41. Pezzarossa, B.; Rosellini, I.; Malorgio, F.; Borghesi, E.; Tonutti, P. Effects of selenium enrichment of tomato plants on ripe fruit metabolism and composition. Acta Hortic. 2013, 1012, 247–252. [Google Scholar] [CrossRef]
  42. Zhu, Z.; Chen, Y.; Zhang, X.; Li, M. Effect of foliar treatment of sodium selenate on postharvest decay and quality of tomato fruits. Sci. Hortic. 2016, 198, 304–310. [Google Scholar] [CrossRef]
  43. Zhu, Z.; Chen, Y.; Shi, G.; Zhang, X. Selenium delays tomato fruit ripening by inhibiting ethylene biosynthesis and enhancing the antioxidant defense system. Food Chem. 2017, 219, 179–184. [Google Scholar] [CrossRef] [PubMed]
  44. Puccinelli, M.; Malorgio, F.; Terry, L.A.; Tosetti, R.; Rosellini, I.; Pezzarossa, B. Effect of selenium enrichment on metabolism of tomato (Solanum lycopersicum) fruit during postharvest ripening. J. Sci. Food Agric. 2019, 99, 2463–2472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Schiavon, M.; Dall’Acqua, S.; Mietto, A.; Pilon-Smits, E.A.H.; Sambo, P.; Masi, A.; Malagoli, M. Selenium fertilization alters the chemical composition and antioxidant constituents of tomato (Solanum lycopersicon L.). J. Agric. Food Chem. 2013, 61, 10542–10554. [Google Scholar] [CrossRef] [PubMed]
  46. Castillo-Godina, R.G.; Foroughbakhch-Pournavab, R.; Benavides-Mendoza, A. Effect of selenium on elemental concentration and antioxidant enzymatic activity of tomato plants. J. Agric. Sci. Technol. 2016, 18, 233–244. [Google Scholar]
  47. Brummell, D.A.; Watson, L.M.; Pathirana, R.; Joyce, N.I.; West, P.J.; Hunter, D.A.; McKenzie, M.J. Biofortification of tomato (Solanum lycopersicum) fruit with the anticancer compound methylselenocysteine using a selenocysteine methyltransferase from a selenium hyperaccumulator. J. Agric. Food Chem. 2011, 59, 10987–10994. [Google Scholar] [CrossRef]
  48. De los Santos-Vázquez, M.E.; Benavides-Mendoza, A.; Ruiz-Torres, N.A.; Cabrera-de la Fuente, M.; Morelos-Moreno, Á. Sodium selenite treatment of vegetable seeds and seedlings and the effect on antioxidant status. Emirates J. Food Agric. 2016, 28, 589–593. [Google Scholar] [CrossRef] [Green Version]
  49. Mimmo, T.; Tiziani, R.; Valentinuzzi, F.; Lucini, L.; Nicoletto, C.; Sambo, P.; Scampicchio, M.; Pii, Y.; Cesco, S. Selenium biofortification in fragaria × ananassa: Implications on strawberry fruits quality, content of bioactive health beneficial compounds and metabolomic profile. Front. Plant Sci. 2017, 8, 1887. [Google Scholar] [CrossRef]
  50. Narváez-Ortiz, W.A.; Becvort-Azcurra, A.A.; Fuentes-Lara, L.O.; Benavides-Mendoza, A.; Valenzuela-García, J.R.; González-Fuentes, J.A. Mineral composition and antioxidant status of tomato with application of selenium. Agronomy 2018, 8, 185. [Google Scholar] [CrossRef] [Green Version]
  51. Poblaciones, M.J.; Rodrigo, S.M.; Santamaría, O. Evaluation of the potential of peas (Pisum sativum L.) to be used in selenium biofortification programs under mediterranean conditions. Biol. Trace Elem. Res. 2013, 151, 132–137. [Google Scholar] [CrossRef]
  52. Ávila, F.W.; Yang, Y.; Faquin, V.; Ramos, S.J.; Guilherme, L.R.G.; Thannhauser, T.W.; Li, L. Impact of selenium supply on Se-methylselenocysteine and glucosinolate accumulation in selenium-biofortified Brassica sprouts. Food Chem. 2014, 165, 578–586. [Google Scholar] [CrossRef] [PubMed]
  53. Bañuelos, G.S.; Arroyo, I.; Pickering, I.J.; Yang, S.I.; Freeman, J.L. Selenium biofortification of broccoli and carrots grown in soil amended with Se-enriched hyperaccumulator Stanleya pinnata. Food Chem. 2015, 166, 603–608. [Google Scholar] [CrossRef] [PubMed]
  54. de Oliveira, V.C.; Faquin, V.; Guimarães, K.C.; Andrade, F.R.; Pereira, J.; Guilherme, L.R.G. Agronomic biofortification of carrot with selenium. Cienc. Agrotecnol. 2018, 42, 138–147. [Google Scholar] [CrossRef]
  55. Schiavon, M.; Berto, C.; Malagoli, M.; Trentin, A.; Sambo, P.; Dall’Acqua, S.; Pilon-Smits, E.A.H. Selenium biofortification in radish enhances nutritional quality via accumulation of methyl-selenocysteine and promotion of transcripts and metabolites related to glucosinolates, phenolics amino acids. Front. Plant Sci. 2016, 7, 1371. [Google Scholar] [CrossRef] [Green Version]
  56. Puccinelli, M.; Malorgio, F.; Rosellini, I.; Pezzarossa, B. Uptake and partitioning of selenium in basil (Ocimum basilicum L.) plants grown in hydroponics. Sci. Hortic. 2017, 225, 271–276. [Google Scholar] [CrossRef]
  57. Puccinelli, M.; Pezzarossa, B.; Rosellini, I.; Malorgio, F. Selenium Enrichment Enhances the Quality and Shelf Life of Basil Leaves. Plants 2020, 9, 801. [Google Scholar] [CrossRef]
  58. Bañuelos, G.S.; Stushnoff, C.; Walse, S.S.; Zuber, T.; Yang, S.I.; Pickering, I.J.; Freeman, J.L. Biofortified, selenium enriched, fruit and cladode from three Opuntia Cactus pear cultivars grown on agricultural drainage sediment for use in nutraceutical foods. Food Chem. 2012, 135, 9–16. [Google Scholar] [CrossRef]
  59. Premarathna, L.; McLaughlin, M.J.; Kirby, J.K.; Hettiarachchi, G.M.; Stacey, S.; Chittleborough, D.J. Selenate-enriched urea granules are a highly effective fertilizer for selenium biofortification of paddy rice grain. J. Agric. Food Chem. 2012, 60, 6037–6044. [Google Scholar] [CrossRef] [Green Version]
  60. Lidon, F.C.; Oliveira, K.; Ribeiro, M.M.; Pelica, J.; Pataco, I.; Ramalho, J.C.; Leitão, A.E.; Almeida, A.S.; Campos, P.S.; Ribeiro-Barros, A.I.; et al. Selenium biofortification of rice grains and implications on macronutrients quality. J. Cereal Sci. 2018, 81, 22–29. [Google Scholar] [CrossRef]
  61. Rodrigo, S.; Santamaría, O.; López-Bellido, F.J.; Poblaciones, M.J. Agronomic selenium biofortification of two-rowed barley under Mediterranean conditions. Plant Soil Environ. 2013, 59, 115–120. [Google Scholar] [CrossRef] [Green Version]
  62. Broadley, M.R.; Alcock, J.; Alford, J.; Cartwright, P.; Foot, I.; Fairweather-Tait, S.J.; Hart, D.J.; Hurst, R.; Knott, P.; McGrath, S.P.; et al. Selenium biofortification of high-yielding winter wheat (Triticum aestivum L.) by liquid or granular Se fertilisation. Plant Soil 2010, 332, 5–18. [Google Scholar] [CrossRef]
  63. Poblaciones, M.J.; Rodrigo, S.; Santamaría, O.; Chen, Y.; McGrath, S.P. Agronomic selenium biofortification in Triticum durum under Mediterranean conditions: From grain to cooked pasta. Food Chem. 2014, 146, 378–384. [Google Scholar] [CrossRef]
  64. Ramkissoon, C.; Degryse, F.; da Silva, R.C.; Baird, R.; Young, S.D.; Bailey, E.H.; McLaughlin, M.J. Improving the efficacy of selenium fertilizers for wheat biofortification. Sci. Rep. 2019, 9, 1–9. [Google Scholar] [CrossRef] [PubMed]
  65. Lara, T.S.; de Lima Lessa, J.H.; de Souza, K.R.D.; Corguinha, A.P.B.; Martins, F.A.D.; Lopes, G.; Guilherme, L.R.G. Selenium biofortification of wheat grain via foliar application and its effect on plant metabolism. J. Food Compos. Anal. 2019, 81, 10–18. [Google Scholar] [CrossRef]
  66. Hussein, H.A.A.; Darwesh, O.M.; Mekki, B.B.; El-Hallouty, S.M. Evaluation of cytotoxicity, biochemical profile and yield components of groundnut plants treated with nano-selenium. Biotechnol. Rep. 2019, 24, e00377. [Google Scholar] [CrossRef]
  67. Morales-Espinoza, M.C.; Cadenas-Pliego, G.; Pérez-Alvarez, M.; Hernández-Fuentes, A.D.; De La Fuente, M.C.; Benavides-Mendoza, A.; Valdés-Reyna, J.; Juárez-Maldonado, A. Se nanoparticles induce changes in the growth, antioxidant responses, and fruit quality of tomato developed under nacl stress. Molecules 2019, 24, 3030. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Hu, T.; Li, H.; Li, J.; Zhao, G.; Wu, W.; Liu, L.; Guo, Y. Absorption and Bio-Transformation of Selenium Nanoparticles by Wheat Seedlings (Triticum aestivum L.). Front. Plant Sci. 2018, 9, 597. [Google Scholar] [CrossRef] [Green Version]
  69. Nazerieh, H.; Ardebili, Z.O.; Iranbakhsh, A. Potential benefits and toxicity of nanoselenium and nitric oxide in peppermint. Acta Agric. Slov. 2018, 111, 357–368. [Google Scholar] [CrossRef] [Green Version]
  70. Li, D.; Zhou, C.; Zhang, J.; An, Q.; Wu, Y.; Li, J.; Pan, C. Nano-selenium foliar applications enhance the nutrient quality of pepper by activating the capsaicinoid synthetic pathway Nano-selenium foliar applications enhance the nutrient quality of pepper by activating the capsaicinoid synthetic pathway. J. Agric. Food Chem. 2020. [Google Scholar] [CrossRef]
  71. Gudkov, S.V.; Shafeev, G.A.; Glinushkin, A.P.; Shkirin, A.V.; Barmina, E.V.; Rakov, I.I.; Simakin, A.V.; Kislov, A.V.; Astashev, M.E.; Vodeneev, V.A.; et al. Production and Use of Selenium Nanoparticles as Fertilizers. ACS Omega 2020, 5, 17767–17774. [Google Scholar] [CrossRef]
  72. El-Kinany, R.; Brengi, S.; Nassar, A.; El-Batal, A. Enhancement of Plant Growth, Chemical Composition and Secondary Metabolites of Essential Oil of Salt-Stressed Coriander (Coriandrum Sativum L.) Plants Using Selenium, Nano-Selenium, and Glycine Betaine. Sci. J. Flowers Ornam. Plants 2019, 6, 151–173. [Google Scholar] [CrossRef] [Green Version]
  73. Zahedi, S.M.; Abdelrahman, M.; Hosseini, M.S.; Hoveizeh, N.F.; Tran, L.S.P. Alleviation of the effect of salinity on growth and yield of strawberry by foliar spray of selenium-nanoparticles. Environ. Pollut. 2019, 253, 246–258. [Google Scholar] [CrossRef] [PubMed]
  74. Amina, Z.; Samar, O. Nano Selenium: Reduction of severe hazards of Atrazine and promotion of changes in growth and gene expression patterns on Vicia faba seedlings. Afr. J. Biotechnol. 2019, 18, 502–510. [Google Scholar] [CrossRef]
  75. Ahmed, H.S.; Ahmed, M.F.; Shoala, T.; Salah, M. Impact of Single or Fractionated Radiation and Selenium Nano-particles on Acid Lime (Citrus aurantifolia L.) Seed Germination Ability and Seedlings Growth. Adv. Agric. Environ. Sci. Open Access 2018, 1, 91–100. [Google Scholar] [CrossRef]
  76. Djanaguiraman, M.; Belliraj, N.; Bossmann, S.H.; Prasad, P.V.V. High-Temperature Stress Alleviation by Selenium Nanoparticle Treatment in Grain Sorghum. ACS Omega 2018, 3, 2479–2491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Tocai, M.; Laslo, V.; Vicas, S. Antioxidant Capacity and Total Phenols Content Changes on Cress (LepidiumSativum) Sprouts after Exogenous Supply with Nano Selenium. Nat. Resour. Sustain. Dev. 2018, 8, 131–137. [Google Scholar] [CrossRef]
  78. Li, Y.; Zhu, N.; Liang, X.; Zheng, L.; Zhang, C.; Li, Y.F.; Zhang, Z.; Gao, Y.; Zhao, J. A comparative study on the accumulation, translocation and transformation of selenite, selenate, and SeNPs in a hydroponic-plant system. Ecotoxicol. Environ. Saf. 2020, 189, 109955. [Google Scholar] [CrossRef]
  79. Skalickova, S.; Milosavljevic, V.; Cihalova, K.; Horky, P.; Richtera, L.; Adam, V. Selenium nanoparticles as a nutritional supplement. Nutrition 2017, 33, 83–90. [Google Scholar] [CrossRef]
  80. Zhu, Y.; Huang, Y.; Hu, Y.; Liu, Y.; Christie, P. Interactions between selenium and iodine uptake by spinach (Spinacia oleracea L.) in solution culture. Plant Soil 2004, 261, 99–105. [Google Scholar] [CrossRef] [Green Version]
  81. Nair, R.; Varghese, S.H.; Nair, B.G.; Maekawa, T.; Yoshida, Y.; Kumar, D.S. Nanoparticulate material delivery to plants. Plant Sci. 2010, 179, 154–163. [Google Scholar] [CrossRef]
  82. Juárez-maldonado, A.; González-morales, S.; Fuente, C.; Medrano-macías, J.; Benavides-mendoza, A. Nanometals as Promoters of Nutraceutical Quality in Crop Plants. In Impact of Nanoscience in the Food Industry; Academic Press: Cambridge, MA, USA, 2018; pp. 277–310. ISBN 9780128114414. [Google Scholar]
  83. Wrobel, K.; Guerrero Esperanza, M.; Yanez Barrientos, E.; Corrales Escobosa, A.R.; Wrobel, K. Different approaches in metabolomic analysis of plants exposed to selenium: A comprehensive review. Acta Physiol. Plant 2020, 42, 125. [Google Scholar] [CrossRef]
  84. D’Amato, R.; Regni, L.; Falcinelli, B.; Mattioli, S.; Benincasa, P.; Dal Bosco, A.; Pacheco, P.; Proietti, P.; Troni, E.; Santi, C.; et al. Current Knowledge on Selenium Biofortification to Improve the Nutraceutical Profile of Food: A Comprehensive Review. J. Agric. Food Chem. 2020, 68, 4075–4097. [Google Scholar] [CrossRef]
  85. Hartikainen, H.; Xue, T.; Piironen, V. Selenium as an anti-oxidant and pro-oxidant in ryegrass. Plant Soil 2000, 225, 193–200. [Google Scholar] [CrossRef]
  86. Battin, E.E.; Brumaghim, J.L. Antioxidant activity of sulfur and selenium: A review of reactive oxygen species scavenging, glutathione peroxidase, and metal-binding antioxidant mechanisms. Cell Biochem. Biophys. 2009, 55, 1–23. [Google Scholar] [CrossRef] [PubMed]
  87. Schiavon, M.; Lima, L.W.; Jiang, Y.; Hawkesfors, M. Effects of selenium on plant metabolism and implications for crops and consumers. In Selenium in Plants. Plant Ecophysiology; Pilon-Smits, E., Winkel, L., Lin, Z.Q., Eds.; Springer International Publishing: Cham, Switzerland, 2017; Volume 11. [Google Scholar]
  88. Pyrzynska, K.; Sentkowska, A. Selenium in plant foods: Speciation analysis, bioavailability, and factors affecting composition. Crit. Rev. Food Sci. Nutr. 2020, 8398. [Google Scholar] [CrossRef]
  89. Dimkovikj, A.; Van Hoewyk, D. Selenite activates the alternative oxidase pathway and alters primary metabolism in Brassica napus roots: Evidence of a mitochondrial stress response. BMC Plant Biol. 2014, 14, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Gupta, M.; Gupta, S. An overview of selenium uptake, metabolism, and toxicity in plants. Front. Plant Sci. 2017, 7, 2074. [Google Scholar] [CrossRef] [Green Version]
  91. Schiavon, M.; Pilon, M.; Malagoli, M.; Pilon-Smits, E.A.H. Exploring the importance of sulfate transporters and ATP sulphurylases for selenium hyperaccumulation—A comparison of Stanleys pinnata and Brassica juncea (Brassicaceae). Front. Plant Sci. 2015, 6, 2. [Google Scholar] [CrossRef] [Green Version]
  92. Cheng, B.; Lian, H.F.; Liu, Y.Y.; Yu, X.H.; Sun, Y.L.; Sun, X.D.; Shi, Q.H.; Liu, S.Q. Effects of selenium and sulfur on antioxidants and physiological parameters of garlic plants during senescence. J. Integr. Agric. 2016, 15, 566–572. [Google Scholar] [CrossRef] [Green Version]
  93. Sun, X.; Han, G.; Ye, S.; Luo, Y.; Zhou, X. Effects of Selenium on Serotonin Synthesis and the Glutathione Redox Cycle in Plum Leaves. J. Soil Sci. Plant Nutr. 2020. [Google Scholar] [CrossRef]
  94. Huang, C.; Qin, N.; Sun, L.; Yu, M.; Hu, W.; Qi, Z. Selenium improves physiological parameters and alleviates oxidative stress in strawberry seedlings under low-temperature stress. Int. J. Mol. Sci. 2018, 19, 1913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Tian, M.; Yang, Y.; Ávila, F.W.; Fish, T.; Yuan, H.; Hui, M.; Pan, S.; Thannhauser, T.W.; Li, L. Effects of Selenium Supplementation on Glucosinolate Biosynthesis in Broccoli. J. Agric. Food Chem. 2018, 66, 8036–8044. [Google Scholar] [CrossRef]
  96. Kopriva, S.; Suter, M.; Von Ballmoos, P.; Hesse, H.; Krähenbühl, U.; Rennenberg, H.; Brunold, C. Interaction of sulfate assimilation with carbon and nitrogen metabolism in Lemna minor. Plant Physiol. 2002, 130, 1406–1413. [Google Scholar] [CrossRef] [Green Version]
  97. Tavakoli, S.; Enteshari, S.; Yousefifard, M. Investigation of the effect of selenium on growth, antioxidant capacity and secondary metabolites in Melissa officinalis. Iran. J. Plant Physiol. 2020, 10, 3125–3134. [Google Scholar]
  98. Tamaoki, M.; Mayurama-Nakashita, A. Molecular mechanism of selenium responses and resistence in plants. In Selenium in Plants. Plant Ecophysiology; Pilon-Smits, E., Winkel, L., Lin, Z.Q., Eds.; Springer: Cham, Switzerland, 2017; Volume 11, pp. 35–51. [Google Scholar]
  99. Liu, J.; Zhu, X.; Chen, X.; Liu, Y.; Gong, Y.; Yuan, G.; Liu, J.; Chen, L. Defense and inhibition integrated mesoporous nanoselenium delivery system against tomato gray mold. Environ. Sci. Nano 2020, 7, 210–227. [Google Scholar] [CrossRef]
Table 1. Recommended Se dietary allowances.
Table 1. Recommended Se dietary allowances.
AgeSe Intake (µg day−1)
Babies up to 6 months10
Babies from 7 to 12 months20
Children from 1 to 3 years20
Children from 4 to 8 years30
Children from 9 to 13 years45
Adolescents from 14 to 18 years old60–70
Adults 19–50 years60–70
Pregnant women60
Lactating women75
Adults over 51 years old70
Table 2. Biofortification of staple and horticultural crops with Se ionic species.
Table 2. Biofortification of staple and horticultural crops with Se ionic species.
Crop, VegetableChemical Specie and DoseApplication FormHighlighted FindingsReference
LettuceNa2SeO3 0.86 mg L−1 and Na2SeO4 3.7 mg L−1Nutrient solutionThe two forms of Se increased the biomass production, ascorbate peroxidase (APX) and glutathione peroxidase (GPX) activities, and reduced the lipoperoxidation. Na2SeO4 was more efficient than Na2SeO3 to accumulate 15 and 2.5 mg Se per kg (DW), respectively.[35]
LettuceNa2SeO4 0.5 mg L−1Nutrient solutionAccumulation of 0.53 mg Se per kg (DW), increase in biomass production, reduction in ethylene concentration.[36]
LettuceNa2SeO3 0.7 mg L−1 and Na2SeO4 1.5 mg L−1Nutrient solutionThe two forms of Se increased the SOD and catalase (CAT) activities. The aerial part biomass increased 3.5% with Na2SeO3 and 5.6% with Na2SeO4.[37]
LettuceSeO2 5 mg L−1 + quitosánNutrient solutionAccumulation of 24 mg Se per kg in leaves (DW).[38]
TomatoNa2SeO3 10 and 20 mg L−1Foliar, every 20 daysIncrease by 3.48 times the Se accumulation in the fruit (DW), and greater total antioxidant capacity. [39]
Tomato, cucumber and lettuceNa2SeO4 10 to 20 mg kg−1Applied directly to peatAccumulation in editable parts were 29.3–48.0 µg kg−1 for cucumber, 22.7–53.4 µg kg−1 for lettuce leaves and 15.2–19.9 µg kg−1. Higher shelf life in lettuce and increased content of vitamin A in tomato[40]
TomatoNa2SeO4 1 mg L−1 HydroponicsDelay in the onset of fruit ripening, lower content of ß-carotene. Se accumulation of 58 µg/100 g fresh fruit[41]
TomatoNa2SeO4 1 mg L−1 FoliarIncreased enzymatic and non-enzymatic antioxidants. Control of gray mold rot[42]
TomatoNa2SeO4 1 mg L−1 FoliarDelayed fruit ripening due to reduced ethylene production and respiration rate[43]
TomatoNa2SeO4 1 and 1.5 L−1 HydroponicsDelayed postharvest ripening due to physiological process such as respiration, ethylene synthesis. Reduced weight loss post harvest[44]
TomatoNa2SeO4 0.9, 1.8, 4.7, 9.4 and 18.8 mg L−1HydroponicsIncreased synthesis of phenolic compounds in leaves. Reduction in Mo, Fe, Mn, Cu in roots[45]
Na2SeO4 2 y 20 mg per plantFoliarThe biofortification of tomato fruit (19 and 256 Se mg kg−1, respectively) increased flavonoids.
TomatoNa2SeO3 2 and 5 mg L−1Nutrient solutionIncrease in plant growth and enzyme antioxidant activity, and average accumulation of 35.8 µg Se per g of fruit (DW).[46]
Transgenic tomato and wild tomatoNa2SeO4 940 mg L−1To soil, from the fruit filling twice a week for 10 weeksThe Se concentration increased from 0.7 to 137 mg per kg of fruit (DW) in transgenic tomato and from 0.3 to 60 mg per kg of fruit (DW) in wild tomato. The transgenic tomato was more efficient than the wild tomato to the overexpression of methyl-selenocysteine transferase (Me-Sec) with 10.9% and 1.5%, respectively.[47]
Seeds and seedlings of melon, lettuce and tomatoNa2SeO3 0.1 and 1 mg L−1, 2 mg L−1, and 5 mg L−1Seed imbibition, nutrient solution and foliar, respectivelyPositive effect on the antioxidant status and vitamin C concentration in the seedlings. [48]
StrawberryNa2SeO4 1.9 and 19 mg L−1Nutrient solutionExcessive accumulation of Se in fruits (46 µg per g DW) with 1.9 mg L−1, and adequate accumulation (60 µg per 150 g FW) with 19 mg L−1. The lower dose increased the biomass and nutraceutical quality, while the higher concentration did not modify the ionomics or growth, and increased the growth regulators, amino acids (phe and arg) alkaloids and isoflavones.[49]
StrawberryNa2SeO3 2 mg L−1Nutrient solutionAccumulation of 31.2 mg Se per kg in fruits (DW) and increase in antioxidant potential.[50]
PeasNa2SeO3 10 g ha−1 and Na2SeO4 10 g ha−1Foliar at the beginning of flowering on sunny daysNa2SeO4 was more efficient than Na2SeO3 to accumulate 148 and 19 µg Se per kg of pea (DW) for each g of Se applied, respectively.[51]
Brassica sproutsNa2SeO4 8.54 mg L−1 Germination of seeds in soaked paper and 1 mL every 24 h Increased the accumulation of organic Se: selenocysteine and reduction in glucosinolates.[52]
Carrot and BroccoliPowdered of Stanleya pinnata 25–200 g (equivalent to 700 mg per plantSoilIncrease in accumulation from 0.5 to 3.5 mg Se per kg in broccoli (DW) and from 0.3 to 22.3 mg Se per kg in carrot (DW). There was no effect on growth.[53]
CarrotNa2SeO4 1 mg L−1FoliarIncreased yield, reducing root ripening and increased titrable acidity.[54]
RadishNa2SeO4 5 mg per plantFoliar, an application when the radish was redAccumulation of 1200 mg Se per kg of leaves (DW) and 120 mg Se per kg of root (DW). Increased the Cys and GSH activities by 2 and 3 times, respectively. Increased 35% the glucosinolates and Me-SeCys.[55]
BasilNa2SeO4 4, 8 and 12 mg L−1Nutrient solutionThe Se accumulation values in the leaves were 2.8, 7.9 and 16.9 µg per g (DW) in the 4, 8 and 12 mg L−1 treatments, respectively. All Se treatments did not affect the growth.[56]
Sweet basilNa2SeO4 4 mg L−1Nutrient solutionSe application increased antioxidants, reduced ethylene synthesis and increased leaf Se in the first cut.[57]
Nopal Se 0.8 mg L−1To the soil contaminated with Se and saltsAccumulation of 3.9 mg Se per kg of fruit (DW) and cladode 15.9 between 16% and 24% in the form of Me-Sec. There was no effect on antioxidant potential or vitamin C concentration.[58]
RiceNa2SeO3 30 g ha−1 and Na2SeO4 30 g ha−1Soil, foliar spray, with granulated urea and nutrient solution in the soilAccumulation of 0.4 to 0.6 mg Se per kg of grain (DW) when applied as Na2SeO4 with granulated urea, the predominant specie was selenomethionine.[59]
Rice Na2SeO3 and Na2SeO4 120–300 g ha−1FoliarIncreased Se in grains, total lipids mostly oleic, linoleic and palmitic acid, and sugars.[60]
BarleyNa2SeO3 40 g ha−1 and Na2SeO4 40 g ha−1Soil in tilleringAccumulation of 0.069 mg to 0.52 and 2.3 g per kg of grain (DW) with Na2SeO3 and Na2SeO4, respectively.[61]
Winter wheatNa2SeO4 100 g ha−1Drench route during early stem extensionAccumulation of 1.6 to 2.6 µg Se per g of grain (DW). Biomass and yield were not affected.[62]
WheatNa2SeO4 40 g ha−1Foliar, an application at the end of the tillering stateIncreased the accumulation of Se by 90% in the form of SeMet from 0.089 to 5.5 mg per kg of grain (DW).[63]
WheatNa2SeO4 10 g ha−1Soil Accumulation of Se in grains close to 50% of Se applied with 10 g ha−1, and between 46% and 61% of Se.[64]
WheatNa2SeO4 + surfactante 21 and 120 g ha−1Foliar, two applications at 36 and 41 days after sowing (DAS)The lower dose increased the production by 48% and biomass by 30%, while the higher dose increased the accumulation from 0.08 to 2.86 mg Se per kg of grain (DW) and the biomass by 31%.[65]
Table 3. Biofortification of staple and horticultural crops with nSe.
Table 3. Biofortification of staple and horticultural crops with nSe.
Crop, VegetablenSe DoseApplication FormHighlighted FindingsReference
PeanutsnSe 40 mg L−1Foliar at 45 and 60 DAS. Increase in the plant growth, unsaturated fatty acids and antioxidant potential. No toxicity was evidenced. [66]
TomatonSe 10 mg L−1To the substrate every 15 daysIncreased the tolerance to salinity stress, better in growth, enzymatic antioxidants (APX and SOD) and non-enzymatic (flavonoids) and fruit quality.[67]
LubianSe 1.18 mg L−1Imbibition of seeds for 2 hIncreased the length of the seedlings as well as the activity of the enzymes α, β amylase and protease, total sugars and total proteins.[8]
WheatnSe chemically and biologically synthesized 5 μMImbibition of seedling bare rootAquaporins are involved in nSe absorption, and Se chemical species such as SeMet and Se IV were accumulated.[68]
Peppermint2 and 20 mg L−1Foliar 8-leaf seedlings, 15 times every 2 daysIncrease in the size of the plant, increase in enzymatic activity of peroxidases and proline at an application concentration of 2 mg L−1. Adverse effects on growth and chlorophylls at 20 mg L−1 were obtained.[69]
Pepper20 mg L−1Foliar. Plants were sprayed four times at 10 d intervalsIncreased chlorophyll, activated phenylpropanoid and capsaicinoid pathways, enhanced photosynthesis.[70]
Radish, arugula, eggplant, cucumber, tomato and chili pepper0.001, 0.005, 0.01 and 0.025 mg kg−1Directly to soil0.01 mg kg−1 increased tolerance to high temperature and increased leaf plate surface. [71]
CoriandernSe 25 and 50 mg L−1Foliar + surfactant tween80 0.005%, twice every 15 daysThe growth of plants subjected to salinity stress was not improved, but the yield, AsA content and electrolyte loss were improved.[72]
PeanutsnSe 20 and 40 mg L−1Foliar at 30 DASDifferent varieties were tested, the growth was improved, increases in chlorophylls, carotenoids and total sugars, as well as in enzymatic antioxidants.[6]
PomegranatenSe 1 and 2 μMFoliar one week prior to flowering, in two consecutive years. The application of 2 µM increased the number and quality of fruits, leaf area, total sugars, phenolic compounds, antioxidants and anthocyanins.[73]
Broad beansnSe 10 and 20 mg L−1Imbibition of seedsThe cytotoxicity and mutagenicity of seedlings intoxicated with the herbicide atrazine were reduced. [74]
LimenSe 50 mg L−1 Imbibition of seedsThe impact of gamma radiation was reduced, and the germination, damping off, percent of albino plants and growth were improved. [75]
SorghumnSe 10 mg L−1Foliar 10 days before appearance of the pinnaclenSe reduced the stress due to high temperatures, increased the SOD, CAT, POX and GPX activities, the unsaturated phospholipids and the pollen germination, and decreased the oxidants.[76]
WatercressnSe 10, 50 and 100 mg L−1Spraying seeds twice a day for 8 daysGrowth was not affected, phenolic compounds increased at 50 mg L−1 and antioxidant capacity at 10 and 100 mg L−1. [77]

Share and Cite

MDPI and ACS Style

García Márquez, V.; Morelos Moreno, Á.; Benavides Mendoza, A.; Medrano Macías, J. Ionic Selenium and Nanoselenium as Biofortifiers and Stimulators of Plant Metabolism. Agronomy 2020, 10, 1399. https://doi.org/10.3390/agronomy10091399

AMA Style

García Márquez V, Morelos Moreno Á, Benavides Mendoza A, Medrano Macías J. Ionic Selenium and Nanoselenium as Biofortifiers and Stimulators of Plant Metabolism. Agronomy. 2020; 10(9):1399. https://doi.org/10.3390/agronomy10091399

Chicago/Turabian Style

García Márquez, Víctor, Álvaro Morelos Moreno, Adalberto Benavides Mendoza, and Julia Medrano Macías. 2020. "Ionic Selenium and Nanoselenium as Biofortifiers and Stimulators of Plant Metabolism" Agronomy 10, no. 9: 1399. https://doi.org/10.3390/agronomy10091399

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

Article Metrics

Back to TopTop