Next Article in Journal
Effect of Aphid Foraging on the Intensity of Photosynthesis and Transpiration of Selected Crop Plants in Its Early Stages of Growing
Previous Article in Journal
Sole and Combined Application of Phosphorus and Glucose and Its Influence on Greenhouse Gas Emissions and Microbial Biomass in Paddy Soils
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Silicon Nanoparticles on Tomato Plants Exposed to Two Forms of Inorganic Arsenic

by
Magín González-Moscoso
1,
Nadia Martínez-Villegas
2,
Gregorio Cadenas-Pliego
3 and
Antonio Juárez-Maldonado
4,*
1
Programa de Ciencias en Agricultura Protegida, Universidad Autónoma Agraria Antonio Narro, Saltillo 25315, Mexico
2
IPICyT, Instituto Potosino de Investigación Científica y Tecnológica, San Luis Potosí 78216, Mexico
3
Departamento de Materiales Avanzados, Centro de Investigación en Química Aplicada, Saltillo 25294, Mexico
4
Departamento de Botánica, Universidad Autónoma Agraria Antonio Narro, Saltillo 25315, Mexico
*
Author to whom correspondence should be addressed.
Agronomy 2022, 12(10), 2366; https://doi.org/10.3390/agronomy12102366
Submission received: 23 August 2022 / Revised: 27 September 2022 / Accepted: 27 September 2022 / Published: 30 September 2022

Abstract

:
In the environment arsenic (As) can be found mainly as arsenite (AsIII) and arsenate (AsV), which are highly toxic and threaten food security. Currently, there is great attention on the effects of silicon dioxide nanoparticles (SiO2 NPs) on plant development, and their ability to restrict As uptake. The results show that the two forms of As negatively impacted aerial dry biomass and fruit yield. Silicon content is lower in roots than in leaves. It is observed that AsIII is the form that accumulates the most in the root; in addition, the SiO2 NPs reduce the translocation of AsV. The data show that AsIII induced a negative effect on the uptake of Ca, P, Mg, and Cu, while SiO2 NPs enhances the accumulation of Fe and Zn when exposed to AsIII. The two forms of As do not impact chlorophyll content but increases when interacting with SiO2 NPs. Antioxidant enzymes APX, CAT, and SOD are higher in roots than in leaves. Phenols, flavonoids, and glutathione increased when SiO2 NPs interacted with AsIII in roots. H2O2 increases in roots and leaves by exposure to AsV and AsIII, and its interactions with SiO2 NPs, while in the fruit, H2O2 production decreases. As for the total antioxidant capacity ABTS is observed to increase by AsIII + SiO2 NPs only in roots. The bioactive compounds of the tomato fruits are modified by the treatments and the addition of SiO2 NPs alone increase in lycopene content. Therefore, our results reveal the negative impacts of AsIII, and that SiO2 NPs can at least partially mitigate As toxicity and reduce AsV translocation in tomatoes.

1. Introduction

Arsenic (As) is a highly toxic chemical element that has been related to different negative effects on human health, such as bladder, kidneys, lung, and skin cancer, cardiovascular disease, and diabetes [1]. In addition, it is omnipresent in the lithosphere and can contaminate all water sources [2]. Additionally, in turn, As can be transferred in the food chain and distributed to entire populations of plants and animals, representing an enormous risk to human health [3].
The main forms of inorganic As in nature are found in two states of chemical oxidation, pentavalent arsenate (AsV) and trivalent arsenite (AsIII) [4]. The oxidation state directly impacts toxicity, since AsIII can generally be 100 times more harmful than AsV in most plant species [5]. The toxicity of As is because it alters the physiological, biochemical, metabolomic, and genomic processes of plants, resulting in a considerable decrease in their productivity [6]. Particularly, in northeast Mexico, As is found in high concentrations in water irrigated for agricultural purposes (8.684 mg L−1), and As concentrations exceed the 0.4 mg L−1 established by the regulations for irrigation water in Mexico [7], which can induce potential negative effects on crop production and human health [8].
Tomato (Solanum lycopersicum L.) is consumed all over the world and is one of the most important crops, around 182.3 million tons of tomato fruits are produced. This crop is sensitive to stress induced by As negatively affecting plant growth and metabolism [9,10]. Interestingly, the concentration of As in the root and aerial parts of the tomato crop is dependent on the concentration in the irrigation water, thus being able to accumulate in the fruit and therefore be a risk for human consumption [11,12].
Silicon (Si) is found in soils naturally, with about 28% of the soils ranging from 0.52% to 47% since most of the rocks and minerals contain Si [12,13]. Si accumulates in plant tissue as hydrated amorphous silica polymer, in concentrations ranging between 0.1 and 10% of dry weight [14]. This element has the capacity to alleviate the harmful impact of heavy metals on plants [15]. It has the ability to reduce the deterioration of the cellular ultrastructure of roots and leaves of plants stressed by toxic metals [16].
Nanotechnology in agriculture is mainly focused on the ecological production of nanomaterials (NMs) for agricultural use, nanofertilizers, nanoencapsulates, nanofilms, and nanoparticles (NPs) [17]. NPs and NMs in low concentrations can favor plant growth; therefore, they can be considered biostimulants [18]. In addition, NPs have the potential as a protective agent for the alleviation of As toxicity in plants [19]. However, these nanometric materials can have different formulations and therefore different mechanisms of action, resulting in different responses in plants [20].
SiO2 NPs have been shown to enhance antioxidant enzymatic activity and photosynthetic activity in Aurundianria pigmaea plants under metal stress [21]. Moreover, improve the development of Triticum aestivum L. plants and reduce the accumulation of cadmium in grains [22]. Even SiO2 NPs increase phenolic compounds and essential oils of Satureja hortensis L. under cadmium stress [23]. In addition, SiO2 NPs induce K+ uptake, modulate Na+ content, and decrease the negative impact on the cell wall of Musa acuminata “Gran Nain” plants in abiotic stress conditions [24]. Even so, they help induce root endodermal silicification and regulate cellular water balance [25]. As for the As-SiO2 NPs interaction, a decrease in the absorption of As in Oryza sativa has been reported, and it can also stabilize cell integrity [26]. Interestingly, SiO2 NPs can reduce the translocation of AsV to the shoots of tomato plants and maintain the photosynthetic balance [27].
Therefore, the present research aimed to evaluate the impact of SiO2 NPs, AsIII, and AsV on tomato plant growth and fruit quality, as well as to determine the concentrations of Si, As, and nutrients in the plant tissue to know the translocation of these elements in the tissues.

2. Materials and Methods

2.1. Crop Development

The experiment was established in a polycarbonate greenhouse of Autonomous Agrarian University Antonio Narro. Tomato seed indeterminate growth and type saladete of the hybrid var. “El Cid F1” was used in this research. The seeds were germinated and subsequently transplanted into black bags with holes, filled with a mixture of perlite-peat moss as growth substrate. To make the correct use of fertilizers more efficient to provide adequate plant nutrition, we use the Steiner solution [28]. The pH of the solution was adjusted to 6.5 and the electrical conductivity (EC) reached 2.1 dS m−1. The concentration applied was 25% for the vegetative stage, 50% for flowering, 75% for fruit set, and 100% for fruit harvest.
A drip irrigation system was used, with four irrigations per day at different times. The amount of water contained in the Steiner solution applied was different for each phenological stage, applying about 2.5 L per plant per day in the stages of highest consumption.

2.2. Treatments

The following two forms of arsenic were supplemented: arsenate (Na2HAsO4∙7H2O) and arsenite (NaAsO2), through the irrigation system containing the Steiner nutrient solution, at a constant amount of 3.2 mg L−1 of As, to simulate contaminated water. SiO2 NPs (1000 mg L−1) were applied directly to the growth substrate, and 10 mL of NPs solution was applied per plant every 21 days (6 applications in total). NPs solution was vortex before application to disperse NPs. SiO2 NPs were purchased at SkySpring Nanomaterials (Houston, TX, USA) with the following characteristics: spherical shape, size (10–20 nm), and surface area of 160 m2 g−1.
The As concentration was chosen from a previous study where different concentrations of As in irrigation water were tested (0.2, 0.4, 0.8, 1.6, and 3.2 mg L−1 of AsV) [27]. In addition, the chosen dose is considered contaminated irrigation water for agricultural use according to the Mexican legislation as follows: NOM-001-ECOL-1996 [29].
The concentration of NPs was chosen from a previous study with application of SiO2 NPs at concentrations between 250 and 1000 mg L−1 in tomato plants [27]. In addition, concentration selection in the above study was based on a study with applications of 100, 500, and 2000 mg L−1 of SiO2 NPs in cotton plants [30].
The experimental design was completely random, with the following six treatments: (1) Control; (2) 1000 mg L−1 SiO2 NPs; (3) 3.2 mg L−1 de AsV; (4) 3.2 mg L−1 de AsIII; (5) 3.2 mg L−1 de AsV + 1000 mg L−1 SiO2 NPs; (6) 3.2 mg L−1 de AsIII + 1000 mg L−1 SiO2 NPs.

2.3. Agronomic Variables and Plant Material Sampling

At five months after transplantation the plants were evaluated the variables yield, to determine the vigor of the tomato plants. In addition, root and leaf samples were collected and the dry biomass of both shoots and roots was determined.
The samples dried in drying oven for 72 h at 80 °C, then the dried tissue was ground with a mortar and pestle until you obtain a fine powder, it was used for analysis of minerals nutrients, and As. For biochemical analysis, five samples of roots and leaves were collected for each treatment (randomly). Roots were washed with water until all residues were removed and air dried inside the greenhouse. Both the leaves and the root were stored in a freezer at −20 °C. The samples were lyophilized at −80 °C for 72 h in a lyophilizer model D401 (Yamato Scientific Co., Ltd., Santa Clara, CA, USA). Completely lyophilized tissue samples were ground in a mortar to powder.

2.4. Arsenic, Silicon and Macro-Micronutrient

The concentrations of As, Si, Ca, K, P, Mg, S, Zn, Fe, and Cu, in roots, leaves, and fruit were determined. Subsequently, acid digestion was performed using concentrated HNO3, and 0.2 g of the dry sample was taken. The final solution was filtered on Whatman filter paper (No. 42). The determination was made on an ICP-plasma emission spectrophotometer (Model 7400, Termo Jarrel Ash).

2.5. Biochemical Analysis

All biochemical analysis readings were performed on a spectrophotometer (UV-Vis-model UNICO UV2150, Dayton, NJ, USA).

2.5.1. Chlorophyll Content

Chlorophylls in leaves were determined by the technique described in [31], using 0.1 g of fine lyophilized tissue. For this, a mixture of hexane-acetone solution (ratio 3:2) was made and 20 mL was added to the lyophilized tissue. Subsequently, it was centrifuged, and an aliquot of the supernatant was collected for reading in the spectrophotometer at two absorbances (between 645 and 663 nm). The absorbances used in Equations (1)–(3), and the values were expressed as mg 100 g−1 dry weight.
Chlorophyll a = 0.999 × Abs(663) − 0.0989 × Abs(645)
Chlorophyll b = 0.328 × Abs(663) + 1.77 × Abs(645)
Total Chlorophyll = Chlorophyll a + Chlorophyll b

2.5.2. Antioxidant Enzymes

For antioxidant enzyme activity determinations, 20 mg of polyvinylpyrrolidone and 200 mg of lyophilized tissue (leaves and roots) were weighed. Subsequently, a 0.1 M phosphate buffer was prepared at a pH of 7, and 1.5 mL of this solution was added, and the mixture was then centrifuged for 10 min at 16,128× g at 4 °C. Finally, the supernatant was filtered through a nylon membrane and used in the determinations of antioxidant enzymes, ABTS, and glutathione.
Glutathione peroxidase activity (GPX) was determined using glutathione as substrate according to the technique of [32]. Catalase enzyme activity (CAT) was based on the quantification of the rate of H2O2 reduction according to the technique described by [33]. Ascorbate peroxidase enzyme activity (APX) was based on the quantification of the rate of ascorbate oxidation according to the technique described by [34], and the SOD Cayman 706002® kit was used to determine the superoxide dismutase enzyme activity (SOD).

2.5.3. Non-Enzymatic Compounds

Vitamin C was performed by colorimetry using 2,6 dichlorophenol according to [35]. The determination of glutathione (GSH) was performed colorimetrically by reaction with 5,5-dithio-bis-2 nitrobenzoic acid, as described in [32]. Total flavonoids were determined with the technique described by [36]. Total phenols were quantified using the Folin-Ciocalteu assay, according to the technique described [37]. Lyophilized tissue was used for all determinations.

2.5.4. H2O2 Concentration and ABTS Radical

Hydrogen peroxide (H2O2) content was determined by the technique reported in [38]. Although some simple modifications were made as described in [39].
The ABTS radical was used to determine antioxidant capacity using the technique of [40]. For this, the lipophilic compounds were extracted with hexane:acetone solution, and the hydrophilic compounds using a phosphate buffer solution. For both H2O2 and ABTS we use lyophilized tissue.

2.6. Fruit Quality and Antioxidants

At the time of harvest, the tomato fruits were selected without damage, homogeneous size, and intense red color [41]. The quality of the tomato fruit is determined by the following variables: fruit firmness, total soluble solids, and hydrogen potential, as described in [42]. While Vitamin C, flavonoids, phenols, and glutathione are described in Section 2.5.3.
For the determination of the lycopene and β-carotene the procedure was followed as described in Section 2.5.1. according to [31]. In addition, four absorbances (453, 505, 645, 663) are used to calculate the concentration of these antioxidants as shown in Equations (4) and (5).
Lycopene = −0.0458 × Abs(663) + 0.204 × Abs(645) + 0.372 × Abs(505) − 0.0806 × Abs(453)
β-carotene = 0.216 × Abs(663) − 1.22 × Abs(645) − 0.304 × Abs(505) + 0.452 × Abs(453)

2.7. Statistical Analysis

A one-way analysis of variance was performed for all the variables evaluated, and to compare the means the Fisher’s Least Significant Difference test was applied (p ≤ 0.05), and data normality was evaluated by Shapiro-Wilks method. All procedures were performed with the statistical program Infostat (V2018) (www.infostat.com.ar, accessed on 26 September 2022).

3. Results

3.1. Effect of SiO2 NPs and Arsenic on Tomato Plants

The application of AsV, AsIII, and SiO2 NPs had significant effects on plant vigor variables of tomato plants (Figure 1). The dry shoot biomass decreased by 8.53%, 11.57%, 9.49%, and 12.75% when AsV, AsIII, and their interactions with SiO2 NPs were applied, respectively (Figure 1A). The dry root biomass did not show any statistical differences between the treatments evaluated (Figure 1B). The fruit yield was increased by 14.55% by the single application of SiO2 NPs compared to the control. As for the AsV, it did not show statistical differences with respect to the control; however, when it interacted with SiO2 NPs, it reduced the fruit yield by 13.60%. The application of AsIII presented the lowest fruit yield, being 17.66% less than the control, but the interaction with SiO2 NPs only reduced it by 13.36% (Figure 1C).

3.2. Silicon and Arsenic Concentration, and Translocation Factor

Si concentration did not show statistical differences in root tissue (Figure 2A). In the leaves, the treatment with AsIII decreased the Si concentration by 54.67% with respect to the control, while the rest of the treatments were statistically equal (Figure 2B). Si concentration in fruits increased 52.9% and 61.39% due to AsV exposure and AsIII + SiO2 NPs interaction, respectively (Figure 2C).
The concentration of As in roots showed a greater accumulation when exposed to AsIII and its interaction with the SiO2 NPs (Figure 2D). While the concentration of As in leaves showed an increase in the treatment of AsIII + SiO2 NPs, while the rest of the treatments were statistically equal (Figure 2E). On the other hand, As was not detected in ripe tomato fruits.
The treatment with AsV showed a higher translocation factor, but when the SiO2 NPs were applied, 37.3% of As was reduced to the aerial part. While AsIII did not show a statistical difference in the translocation factor when the SiO2 NPs were applied (Figure 2F).

3.3. Concentration of Macro and Micronutrients in Tissues

AsIII decreased 43.7% and 38.31% of the concentration of Ca in leaves and fruits, respectively; however, in roots there were no differences (Figure 3A). The concentration of K showed differences in the three plant organs. The K in leaves decreased by 38.48% and 42.22% in the SiO2 NPs and AsIII treatments, respectively. The concentration of K in roots was lower in all treatments compared to the control. While in fruits, the AsIII + SiO2 NPs treatment increased it by 43.04% (Figure 3B). The concentration of P in leaves decreased 50.43% with the application of AsIII, while the rest of the treatments showed no statistical differences. In roots and fruits, the concentration of P did not show statistical differences (Figure 3C). The concentration of Mg in the leaves increased 46.66% with exposure to AsV; however, when exposed to AsIII, it decreased by 47.33%. In roots and fruits, Mg did not have statistical differences between treatments compared to the control (Figure 3D). The concentration of S in leaves did not show statistical differences between treatments evaluated, while in roots a decrease of 38.37–33% was observed in the AsV + SiO2 NPs and AsIII + SiO2 NPs treatments, respectively; in fruits, S was not detected (Figure 3E).
The concentration of Fe in leaves increased by 79.64% and 155.6% due to the application of AsV and AsIII, respectively, while in roots there were no differences between treatments. The concentration of Fe in fruits increased by 105.34% by the application of AsV alone, while the AsV + SiO2 NPs and AsIII + SiO2 NPs treatments increased the accumulation of Fe by 55.43% and 127.99%, respectively (Figure 3F). The concentration of Zn in leaves did not present statistical differences in the treatments evaluated; while in roots, all treatments decreased compared to the control. In fruits, the SiO2 NPs, AsV and AsIII + SiO2 NPs treatments increased Zn by 151.6%, 96.07% and 96.04% respectively (Figure 3G).
The concentration of Cu in leaves decreased 45.60% due to the application of SiO2 NPs and 214.46% due to exposure to AsIII. The application of SiO2 NPs alone decreased the Cu concentration in roots by 25.95%, and 27.98% in fruits (Figure 3H).

3.4. Impact on Chlorophyll Content

Chlorophylls (a, b, and total) in tomato plants exposed to As and SiO2 NPs showed statistical differences between treatments evaluated (Figure 4). Chlorophyll a content increased by 13.02% and 12.52% in treatments that were exposed to AsV and AsIII in interaction with SiO2 NPs compared to As alone (Figure 4A). Chlorophyll b increased by 25.17% when AsV + SiO2 NPs were applied, while AsIII + SiO2 NPs increased by 19.17%. The rest of the treatments did not show statistical differences with respect to the control (Figure 4B). SiO2 NPs, AsV, and AsIII when applied alone did not impact the total chlorophyll content; however, when plants were exposed to AsV and AsIII in interaction with SiO2 NPs, there was an increase of 18.60% and 19.26% respectively (Figure 4C).

3.5. Antioxidant Enzymes

Antioxidant enzymatic activities of both roots and leaves were modified when exposed to As and SiO2 NPs (Figure 5). APX activity in the leaves increased 48.9% due to the effect of AsV, while AsIII had a negative effect on APX activity, decreasing up to 26.98%. However, the addition of SiO2 NPs increased APX activity by 42.90%, reversing the negative impact of AsIII. In the root, an increase was observed due to the application of As alone, and in interaction with the SiO2 NPs, the interaction of AsIII + SiO2 NPs increased by 47.6%; however, all treatments were statistically equal to the control (Figure 5A).
GPX enzymatic activity was different in roots than in leaves. In leaves, there were no statistical differences between treatments. In the root, the GPX activity increased by 112.80% by the application of SiO2 NPs, while in interaction with AsV and AsIII it decreased by 79.04% and 70.30%, respectively (Figure 5B).
The CAT activity did not show statistical differences by application of the treatments. In the CAT activity in roots, AsV induced an increase of 191%, and when it interacted with SiO2 NPs, it increased by 160.38%. The presence of AsIII seems to have a greater effect on CAT activity. It was increased by up to 491%, while the AsIII + SiO2 NPs interaction increased by 417.7% (Figure 5C).
The SOD enzymatic activity in leaves did not present statistical differences between treatments. In the root, the SiO2 NPs treatment decreased SOD activity by 34.67%, while the AsV + SiO2 NPs interaction reduced it by 34.19%; the remaining treatments were statistically the same as the control (Figure 5D).

3.6. Non-Enzymatic Antioxidants

The vitamin C content is generally higher in the leaves than in the roots, with statistical differences only in the leaves where vitamin C increased in all treatments with respect to control in a range of 63.92–100.63% (Figure 6A).
The concentration of phenols in the leaves did not present statistical differences between the treatments. While phenols increased in roots (39.16% and 30.14%) with the application of AsV and in interaction with SiO2 NPs, respectively. While AsIII increased the phenol content by 58.16%, and the AsIII + SiO2 NPs interaction induced an increase of 62.66% (Figure 6B).
Flavonoids in roots and leaves behave differently; in the leaves, a higher concentration of flavonoids was observed. The application of SiO2 NPs increased the flavonoid content by 26.1%, AsIII increased by 11%, and the AsV + SiO2 NPs and AsIII + SiO2 NPs interactions increased by 41.35% and 11.37%, respectively. In the root, the SiO2 NPs increased the flavonoid content by 27.62%, while the AsIII + SiO2 NPs interaction induced an increase of 11.37% (Figure 6C).
The GSH content in the leaves increased by 36.22% and 37.72%due to exposure to AsV and AsIII, respectively; the other treatments were statistically equal to the control. In the roots, AsIII induced an increase of GSH of 94.68%, and in interaction with SiO2 NPs it increased by 77.65%. The other treatments were statistically equal to the control (Figure 6D).

3.7. Hydrogen Peroxide and Antioxidant Capacity

The level of H2O2 in the leaves increased with all treatments in a range from 15.2 to 25.65%. In the root, H2O2 increased by 31.15% and 50% due to exposure to AsV and AsIII, respectively. While the interaction of AsV and AsIII with SiO2 NPs increased the H2O2 level by 67.39% and 43.47%, respectively. In the fruits, all the treatments decreased the level of H2O2, in a range of 44.9–54.49% (Figure 7A).
ABTS antioxidant capacity in roots had statistical differences between treatments. In hydrophilic compounds, the interaction of AsIII with SiO2 NPs increased by 21.24%. In the lipophilic compounds, AsV decreased the antioxidant capacity by 4.13%. The total antioxidant capacity was higher (6.6%) in the AsIII + SiO2 NPs treatment, in relation to control (Figure 7B).
The antioxidant capacity in leaves did not present a statistical difference between treatments in hydrophilic compounds. In the lipophilic compounds, the AsIII treatment decreased the antioxidant capacity by 6.36%. Total antioxidant capacity was increased by 3.8% with AsIII treatment (Figure 7C).
The antioxidant capacity in fruits did not show any statistical differences in lipophilic compounds. While the antioxidant capacity of hydrophilic compounds was the only difference between the two forms of arsenic, being higher with AsV (Figure 7D).

3.8. Effect of As and SiO2 NPs on the Quality of Tomato Fruits

Fruit firmness decreased by 28.93% and 21.13% in the treatments with AsV and AsIII application, respectively, while the AsIII + SiO2 NPs interaction decreased by 18.99% in relation to the control (Figure 8A).
Total soluble solids increased when plants were exposed to AsV (8.33%) and AsIII (6.57%); however, when AsV interacted with SiO2 NPs, they decreased by 11.40% (Figure 8B).
The pH of ripe fruits was not modified by the application of SiO2 NPs, however, the exposure to the two forms of As and its interaction with the SiO2 NPs decreased the pH in a range of 3.45–4.99% (Figure 8C).
The main antioxidant lycopene in ripe tomato fruit increased by 17.17% by the application of SiO2 NPs, while AsIII induced an increase of 28.12% (Figure 8D).
β-carotene increased with AsV by 42.44% (Figure 8E). Vitamin C increased with all applied treatments in a range of 36.8–49.63% in relation to the control (Figure 8F).
With the exception of the AsV + SiO2 NPs treatment, the rest decreased the glutathione content in the fruits in a range of 13.70–23.43% (Figure 8G).
Flavonoids were decreased by exposure to AsV alone or in combination with SiO2 NPs by 13.70 and 14.74%, respectively (Figure 8H).
Phenol content increased with the AsV + SiO2 NPs treatment by 19.12%, while the AsIII + SiO2 NPs interaction induced an increase of 16.38%, although it was not different from the control (Figure 8I).

4. Discussion

4.1. Impact of As and SiO2 NPs on Biomass and Fruit Yield

In this study, both AsV and AsIII negatively impacted shoot dry biomass, and AsIII reduced fruit yield. The literature reports a negative effect on plant development by exposure to arsenic, and our study showed that reported trend. For example, negative effects on biomass due to the presence of As in the soil have been reported in tomato, rice, and parsley plants [9,43]. Moreover, AsIII causes a reduction of the dry biomass of rice plants by up to 50% when they are grown in a nutrient solution with a concentration of 4.0 mg L−1 of this metalloid [43]. Although it has been reported that AsIII is much more toxic than AsV, a serious affectation of growth was observed in the Atriplex atacamensis plant, an effect not observed in AsV treatment [44]. Therefore, the exposition of plants to As causes physiological alterations in plants and interrupts the accumulation of biomass due to its high toxicity [45].
The negative impact is due to arsenic contamination in any of its two forms (AsV and AsIII) can negatively interfere with plant metabolism, increasing the formation of reactive oxygen species (ROS) [46]. Furthermore, AsV disrupts cellular metabolism because it replaces phosphorus in respiration processes, instead of generating adenosine triphosphate, it generates adenosine diphosphate-arsenate [47]. While the interaction of AsIII with the sulfhydryl groups of enzymes and the formation of ROS, severely alters plant metabolism, DNA structure, and lipid-protein metabolism [48].
In our study, SiO2 NPs alone increased fruit yield. When the NPs come in contact with the root or leaves, they are absorbed and can be transported, generating a stimulating effect on plants subjected to abiotic stress [49]. This may be due to the promotion of enzymatic activity induced by the NPs, which intervene in the metabolism and acquisition of nutrients, and improve the opening of stomata, which results in an increase in gas exchange and CO2 assimilation [50]. They also improve hormone content in plants, such as indoleacetic acid, gibberellins, and cytokinins, and can positively regulate gene transcription [51,52].
The application of Si NPs to plants subjected to heavy metal stress conditions can improve the development of plants [53,54]. Although in our study, SiO2 NPs could not reverse the negative effect on shoot biomass and yield of tomato plants in the presence of As, and also did not cause increased toxicity when interacting with As. In this regard, the effect of Si NPs on plant stress is related to the physical and chemical properties of the Si NPs, the soil, and the type of stress [55].

4.2. Concentration of Si, As and Nutrients

In this study, the application of AsIII in the irrigation water did not impact root Si uptake but decreased the content of silicon in the leaves of tomato plants. This impact can be attributed to AsIII uptake by roots, which is mediated by Si transporters [56]. For example, (Lsi1, Lsi2, Lsi3, and Lsi6), and also the anatomy of the different roots, show a variable absorption of Si between plants [57]. It has also been shown that Si accumulates mainly in plant shoots and has a lower content in the roots [58]. While a higher applied amount of Si linearly increases the content of this element in the different organs [59].
Interestingly, in our study, the application of SiO2 NPs to AsIII stressed plants induced a higher Si content in leaves. This is due to the form of Si (ionic or NPs) translating into a different concentration, being higher when it is applied in the form of NPs [60]. In this regard, Wang et al. [61] reported that the SiO2 NPs in rice seedlings induced a greater accumulation of Si when the seedlings developed in the presence of Cd and As. The impact was dependent on the concentration of NPs (between 150, 500 and 2000 mg kg−1) and the conditions of irrigation.
In our study, both forms of As induced As accumulation in the roots and leaves of tomato plants, but in the tomato fruits, it was not detected. The entry of As into tomato plants is due to its similarity with P. Therefore, As can enter through aquaglyceroporins and P nutrient uptake pathways, e.g., phosphate permeases [62]. In addition, the increase in As in the water used to irrigate agricultural plants results in a greater accumulation of As in the roots and shoots [63]. Various studies have shown that tomato plants accumulate most of the As in the roots and a very low concentration is translocated to the aerial parts, which means that the consumption of the human population presents a lower risk [8,64,65,66]. Interestingly, tomato plants reduce As transport to the aerial parts and increase As accumulation in the roots, as a strategy to tolerate this metalloid [67].
In our research, the application of SiO2 NPs reduces the translocation of AsV to the aerial parts but has no impact on the translocation of AsIII. It has been shown that the addition of SiO2 NPs to the roots of tomato plants irrigated with AsV helps to restrict the translocation of As to the aerial parts [27]. Even the application of chitosan-Si NPs promotes arsenic fixation in the leaf cell wall and reduces the As concentration in rice grain by 61.2% [68]. In this regard, [69] informed that Si can chelate, compartmentalize, and coprecipitate toxic metals, and can regulate antioxidants. In addition, Si NPs decrease oxidative stress and increase the potential of the photosynthetic system in metal-stressed plants [22]. Even the application of SiO2 NPs can enhance gene expression of OsNIP1;1 and OsNIP3;3 and can significantly reduce the entry of As into the cells [26]. All of these processes can improve plant health and can restrict metal translocation. On the other hand, a reduction of AsIII translocation in rice plants with silicon application is commonly reported, but it has also been shown that Si application does not affect AsIII concentration in roots and does not affect DMA uptake and accumulation in roots and shoots [56]. Therefore, further studies should be conducted on tomato plants to understand more about this Si-AsIII interaction process because in our study, the application of SiO2 NPs did not influence AsIII translocation.
In this study, the uptake and accumulation of most essential nutrients in tomato plants were negatively impacted by the application of the two forms of As in irrigation water. This is because As can alter metabolic processes and can compete directly with nutrients, influencing the uptake and distribution of nutrients in plants [70]. Reduced Ca, P, Mg, and N uptake have been reported in different tissues of perennial ryegrass plants subjected to an As stress condition [71]. In Cajanus cajan L. plants, lower nutrient uptake was observed when they were exposed to AsIII compared to AsV [72].
Furthermore, findings of current research revealed that the addition of SiO2 NPs in tomato plants subjected to arsenic stress conditions on the absorption of mineral nutrients showed the following three impacts: (1) no effect, (2) negative effect, and (3) positive effect at least on Ca and Cu in fruits. In this context, it has been shown that SiO2 NPs increased the content of nutrients such as Mg, P, and S in Lilium plants, improving the quality of their flowers and shelf life [73]. In Cucumis sativus plants, Si NPs increased K uptake in different plant organs [74]. In tomato plants, SiO2 NPs have a positive impact on the uptake and concentration of Fe in leaves, and Cu and Zn in roots even the uptake of macronutrients in the roots and leaves also increases even when tomato plants are subjected to AsV stress conditions [75]. These observed effects are due to the fact that SiO2 NPs can modify nutrient uptake due to the roots can excrete organic acids and therefore can improve the absorption of mineral nutrients [76].
Moreover, the negative impact of high concentrations of SiO2 NPs (500 and 2000 mg L−1) on the content of mineral nutrients, such as Cu and Mg, has been reported in cotton plants [30]. The addition of SiO2 NPs in our study reduced S and Zn uptake in roots, but the rest of the nutrients were not impacted. Probably there was a regulation in the homeostasis of essential nutrients to tolerate the toxic effect of As.

4.3. Impact of As and SiO2 NPs on Plant Photosynthetic Pigments

In this study, the two forms of As did not affect chlorophyll content. A negative effect is commonly reported due to the fact that the As is toxic to organisms in the ecosystem, including plants. The degree of toxicity of As species follows the following order: (AsIII) > (AsV) > (MMA: monomethylarsonous acid) > (DMA: dimethylarsinic acid) [77]. Nevertheless, our results are not the only ones where a null impact of As on plants is reported. For example, a high tolerance to AsV has been reported in some plants such as Taray (Tamaris gallica), without showing a significant variation in chlorophyll synthesis [78]. Even the electron transport rate of the photosystems (PSII and PSI) of rice seedlings exposed to AsV does not show significant effects [79]. AsIII in Vitis vinifera plants first reduces photosynthesis and, over time, stimulates it [80].
The processes involved in not generating a negative impact on chlorophylls or plant development may be due to AsV being rapidly reduced to AsIII in plants. The efflux of AsIII to the external environment is a form to minimize the effect of As toxicity on plants [81]. In addition, AsIII forms complexes with phytochelatins (PC) within root cells and subsequently accumulates AsIII-PC complexes in vacuoles via OsABCC1 transporters as a detoxification strategy [82].
In our research, the addition of SiO2 NPs to plants subjected to high As concentration induced an increase in chlorophyll content. AsIII-Fe3O4 NPs interaction has been reported to improve chlorophyll and carotenoid content in Brassica juncea plants in relation to exposure to As applicated alone [83]. Under chromium exposure, Si NPs increase chlorophyll levels in Pisum sativum plants [53]. Even in banana plants (Musa acuminata ‘Grand Nain’) under abiotic stress, the addition of SiO2 NPs improves the chlorophyll content [24].
The positive impact of the application of the NPs on photosynthetic pigments may be due to NPs having photocatalytic properties that allow them to induce a rapid oxidation-reduction reaction, which leads to charge transfer with the light-harvesting complexes that are linked to the photosystems (PSI and PSII) [84]. In this regard, the addition of Si NPs shows a maximized enhancement of PSII and of the electron transport rate in Cymbopogon flexuosus (Steud.), and the leaves appear more proficient at capturing and quenching energized photons [85]. Even the chlorophylls in the photosynthetic reaction center can bind to NPs, and due to the plasmon resonance of NPs, they can produce more excited electrons, thus forming a new hybrid system [25]. In addition, Si NPs initiate a cascade of morphophysiological adjustments that improve plant physiology by regulating the expression of many photosynthetic genes and proteins together with photosystem I (PSI) and PSII assemblies [86]. Therefore, it is possible that SiO2 NPs may have modulated the expression level of photosynthesis-related genes in tomato plants.

4.4. Impact of As and SiO2 NPs on Antioxidant System

In our study, APX, CAT, and SOD enzymatic activities in roots were higher compared to activity in leaves. Similarly, a study with Centella asiatica plants subjected to Pb stress conditions showed higher activities of GPX, APX, and CAT in the roots than in the leaves [87]. The difference in enzyme activities between the leaf and the root may be due to the fact that the contaminant comes into natural contact with the radical system, and this system activates a rapid response as protection against oxidative stress [88].
Furthermore, in this study, enzymatic activity shows different responses due to exposure to the two forms of As in both roots and leaves. For example, APX enzymatic activity in leaves was the only one that increased by As exposure, while in roots, APX and CAT activity showed an increase. In this context, Duquesnoy et al. [89] reported that in Zea mays plants exposed to AsV and AsIII, CAT activity increased in both leaves and roots. While SOD activity was reduced in Hydrilla verticillata by exposure to high doses of AsV and AsIII [90]. Although in our study, the two forms of As did not influence the SOD activity in the two plant organs evaluated. As generates oxidative stresses in plants in the following two ways: (1) binding with a sulfhydryl group deactivates antioxidants, and (2) producing ROS in the oxidation-reduction process of their chemical forms [91]. The plant rapidly activates its antioxidant defense system with which it can counteract the oxidative stress generated by this metalloid [46].
In our study, SiO2 NPs applied to As-stressed tomato plants showed an increase in APX and CAT activity, while GPX and SOD reduced their activity. In this context, Tripathi et al. [92] showed a significant enhancement of APX activity in Zea mays plants by the addition of Si NPs that had been inhibited by As. In addition, the application of ZnO NPs increased CAT activity in Glycine max plants supplemented with AsIII [6]. While the AsV-Fe2O3 NPs interaction in Vigna radiata plant roots decreased the activity of this enzyme compared to the addition of only As [93]. Interestingly, the high concentration of As in the irrigation water and its interaction with high concentrations of Si NPs may induce additional stress to tomato plants, which may be reflected in decreases in SOD and GPX activity [27]. Regarding the reduction of SOD activity, it may be due to the fact that this enzyme actively participates in the elimination of free radicals produced by stress in plants and is considered the first line of defense [94]. While GPX activity reduction may be due to an inhibition of glutathione production in the leaves, as shown in Figure 6D, since this compound is part of the substrate of this enzyme.
Furthermore, the effect of NPs on the antioxidant system of plants can be diverse. The interactions of the surface charges of the NPs cause the non-specific interaction of the cell membrane receptors and cell wall, which can lead to the following two responses: toxicity or biostimulation [95]. When biostimulation occurs, plants can accumulate higher concentrations of antioxidant compounds due to exposure to NPs. This increase will be reflected in the fruits as a redox state that NPs induce in response to stress [96].
As for non-enzymatic antioxidants such as flavonoids, glutathione, carotenoids, phenols, and vitamin C, act naturally against stress in plants [97]. In our study, non-enzymatic antioxidant compounds were lower in the roots than in the leaves. In addition, GSH, vitamin C, and flavonoids were increased by exposure to both forms of As in leaves, and phenols in roots were increased by As exposure. In this context, GSH is a relevant compound in heavy metal tolerance since it is the main substrate of non-protein thiols in most plant cells and can detoxify metalloids by phytochelatins in response to As toxicity [98]. In addition, As forms a complex with GSH and can be sequestered in the vacuole [99]. Therefore, the As deposited in the vacuole is probably inert. Likewise, it has been reported that in hyperaccumulator plants, arsenic exposure induces a higher flavonoid content [100]. As for as phenols, exposure of Ulmus laevis Pall to different chemical forms of As accumulation increases the accumulation of this important antioxidant [101]. This is because phenols act as protectants in response to high doses of toxic metals or metalloids [102]. In contrast to our results, it has been reported that in Oryza sativa seedlings the ascorbic acid content decreased by 33% and 51% with the exposition to 0.5 and 1 mM AsV respectively [103].
In this study, the addition of SiO2 NPs in the absence of both forms of arsenic increased vitamin C and flavonoid content. While when SiO2 NPs are applied to tomato plants stressed with AsV, only flavonoid content in leaves increases and glutathione content in roots decreases. The other compounds were not affected and in interaction with AsIII showed no effect. An increase in the content of phenols and flavonoids due to exposure to NPs has been reported in plants with abiotic stress [104]. Possibly, this increase is related to an upregulation of phenylpropanoid pathway genes [105]. Moreover, in lettuce seedlings with exposure to NPs, it has been observed that an increase in glutathione and ascorbic acid, possibly as a result of an alteration in the ascorbate-glutathione (AsA-GSH) pathway [106]. Even so, the application of NPs shows a positive regulation in the expression of genes related to glutathione biosynthesis, glutathione reductase, and glutathione S-transferase in Arabidopsis thaliana [107]. The positive impact of the Si NPs on the antioxidant system may be due to the fact that the Si allows a lower entry of the metal into the plants [108]. In addition, SiO2 NPs can maintain cell integrity by decreasing cellular ROS levels, increasing cell wall thickness, and increasing the proportion of As in pectin [26].
In this study, the total ABTS antioxidant capacity in roots showed an increase due to the interaction of AsIII + SiO2 NPs, in leaves, it decreased due to exposure to AsIII, and in fruits, a lower capacity of AsIII was observed compared to AsV. Antioxidant capacity is widely used as a parameter to characterize the ability to scavenge or neutralize free radicals [109]. Exposure to high doses of AsV and SiO2 NPs has been reported to increase antioxidant capacity in both hydrophilic and lipophilic compounds in tomato fruits and also H2O2 levels [39]. NPs entering the plant can induce the generation of ROS because they cause an imbalance at the molecular, biochemical, and physiological levels that produces oxidative stress [110]. However, ROS induced by NPs can act as signalers in plants and may be able to alter secondary metabolism through their transcriptional reprogramming, causing a series of positive responses [111].
In our study, H2O2 production increased in leaves in all treatments with respect to the control and decreased in fruits. In roots, the two forms of As applied alone and in interaction with SiO2 NPs increased H2O2 production. Contrary to our results, Tripathi et al. [92] reported that Si NPs can reduce H2O2 production in Oryza sativa leaves exposed to AsV. However, H2O2 is also an important signaler in biological pathways, so that slight accumulations of H2O2 can stimulate photosynthesis [112], and in our study, chlorophylls were stimulated as shown in Figure 4.
On the other hand, in fruits, H2O2 decreased. This result was similar to a previous study, in which the concentration of 1000 mg L−1 SiO2 NPs decreased the production of H2O2 in tomato fruits [39]. The decrease in H2O2 in tomato fruits is probably due to a higher concentration of some antioxidant compounds in the fruit, as shown in Figure 7, and these compounds could counteract this free radical. In addition, both the natural ripening process, arsenic stress, and environmental changes can modify the production of ROS in tomato fruit.

4.5. Impact of As and SiO2 NPs on the Quality of Ripe Tomato Fruits

The antioxidant content of plants can be modified by both the stage of fruit development and environmental restrictions. The stage of fruit maturity at harvest is an essential factor in determining antioxidant levels in tomatoes [113]. In our study, both forms of arsenic decreased the firmness, glutathione, and hydrogen potential of tomato fruits. However, both AsV and AsIII increased the content of total soluble solids, β-carotene, and vitamin C. As exposure can modify the nutraceutical quality, e.g., GSH, β-carotene, lycopene, phenols, and H2O2 in tomato fruits of different cultivars [11], as a response to oxidative stress generated by As in the plant [114]. This is because As can accumulate and be distributed to tomato fruits, and the degree of entry of As into the fruit varies from one cultivar to another [115]. Even As causes ROS accumulation in tomato fruits, accumulation of proteins related to the response to oxidative stress and other proteins related to ubiquitination [116]. Alterations in quality attributes of tomato fruit due to the effect of a heavy metal are related to the expression of genes associated with sucrose biosynthesis (SlSuSys, SlLin5, SlLin6, SlLin7) and carotenoids (SlZDS, SlCRTISO, SlbLCY) [117].
In this study, the application of SiO2 NPs in the absence of As increases lycopene and vitamin C content. The increase in the concentrations of bioactive compounds by exposure to NPs will be reflected in the fruit as a redox state that NPs induce in response to oxidative stress [96]. Moreover, the NPs improve the quality of the fruit since they can significantly alter the proteins related to ascorbate and glutathione, exhibiting a greater amount of antioxidants in the fruits [118]. Furthermore, the addition of NPs to tomato plants can increase fruit quality through transcriptional upregulation of the carotene isomerase gene (CRTISO), a gene involved in carotenoid biosynthesis [119]. While the nutritional quality of the fruits exposed to NPs can be negatively affected; however, this may depend on different factors such as type, shape, size, and applied concentration of NPs, as well as the crop [120].
Furthermore, in this study, the AsV + SiO2 NPs interaction showed a decrease in total soluble solids, hydrogen potential, and flavonoids, but increased phenols and vitamin C in ripe tomato fruits. While the AsIII + SiO2 NPs interaction decreased glutathione content, firmness, and hydrogen potential, this interaction only increased vitamin C content. Interestingly, it has been shown that the SiO2 NPs under abiotic stress conditions improve the yield and quality of mango fruits [121]. Even plants growing under different arsenic concentrations and SiO2 NPs are applied, which have a positive effect on the modification of the generation processes of antioxidants in tomato fruit [39]. This may be due to the fact that Si improves the antioxidant compounds of fruits and improves commercial qualities such as firmness, this last attribute indicates that Si may have a key role in stabilizing the cell wall, protecting it from degrading enzymes such as polygalacturonase, β-galactosidase, and pectin methyl esterase [122].
In tomato fruits, there is still a lack of knowledge about redox metabolism. Future studies could generate a better understanding because there is a lack of information on the impact of nanoparticles on the nutraceutical quality of ripe fruits, as generally only an increase or decrease is reported for different doses and types of NPs. There is even a lack of information on the impact of As on fruit quality, as there are few studies that address this problematic on the quality of fruit that is ultimately consumed by humans.

5. Conclusions

This research shows that AsIII has a greater negative impact than AsV. Both forms of As reduced shoot dry biomass production and firmness, and glutathione content in ripe fruit. However, AsIII reduced fruit yield and decreased the concentration of Si, Mg, P, K, and Cu in the leaves of tomato plants.
The application of SiO2 NPs increased the content of chlorophyll. Moreover, it improves nutrient uptake and antioxidant content when plants were exposed to AsIII and reduces AsV translocation, resulting in an adaptation of tomato plants to stress caused by this metalloid. Therefore, the use of SiO2 NPs is possible as an alternative to improve the productivity of tomato plants when the waters contain the presence of As and they are used for agricultural purposes.

Author Contributions

Conceptualization, A.J.-M. and N.M.-V.; methodology, M.G.-M. and G.C.-P.; supervision, A.J.-M.; writing—original draft preparation, M.G.-M.; writing—review and editing, M.G.-M., N.M.-V., G.C.-P. and A.J.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fleming, D.E.B.; Crook, S.L.; Evans, C.T.; Nader, M.N.; Atia, M.; Hicks, J.M.T.; Sweeney, E.; McFarlane, C.R.; Kim, J.S.; Keltie, E.; et al. Assessing arsenic in human toenail clippings using portable X-Ray fluorescence. Appl. Radiat. Isot. 2020, 167, 109491. [Google Scholar] [CrossRef] [PubMed]
  2. Atabaki, N.; Shaharuddin, N.A.; Ahmad, S.A.; Nulit, R.; Abiri, R. Assessment of water mimosa (Neptunia oleracea Lour.) morphological, physiological, and removal efficiency for phytoremediation of arsenic-polluted water. Plants 2020, 9, 1500. [Google Scholar] [CrossRef] [PubMed]
  3. He, D.; Xiong, Y.; Wang, L.; Sun, W.; Liu, R.; Yue, T. Arsenic (III) removal from a high-concentration arsenic (III) solution by forming ferric arsenite on red mud surface. Minerals 2020, 10, 583. [Google Scholar] [CrossRef]
  4. Xu, W.; Dai, W.; Yan, H.; Li, S.; Shen, H.; Chen, Y.; Xu, H.; Sun, Y.; He, Z.; Ma, M. Arabidopsis NIP3;1 plays an important role in arsenic uptake and root-to-shoot translocation under arsenite stress conditions. Mol. Plant 2015, 8, 722–733. [Google Scholar] [CrossRef]
  5. Chandrakar, V.; Naithani, S.C.; Keshavkant, S. Arsenic-induced metabolic disturbances and their mitigation mechanisms in crop plants: A review. Biologia 2016, 71, 367–377. [Google Scholar] [CrossRef]
  6. Ahmad, P.; Alyemeni, M.N.; Al-huqail, A.A.; Alqahtani, M.A.; Wijaya, L.; Ashraf, M.; Kaya, C.; Bajguz, A. Zinc oxide nanoparticles application alleviates arsenic (As) toxicity in soybean plants by restricting the uptake of as and modulating key biochemical ascorbate-glutathione cycle and glyoxalase system. Plants 2020, 9, 825. [Google Scholar] [CrossRef]
  7. Ruíz-Huerta, E.A.; de la Garza Varela, A.; Gómez-Bernal, J.M.; Castillo, F.; Avalos-Borja, M.; SenGupta, B.; Martínez-Villegas, N. Arsenic contamination in irrigation water, agricultural soil and maize crop from an abandoned smelter site in Matehuala, Mexico. J. Hazard. Mater. 2017, 339, 330–339. [Google Scholar] [CrossRef]
  8. González-Moscoso, M.; González-García, Y.; Martínez-Villegas, N.V.; Cadenas-Pliego, G.; Juárez-Maldonado, A. Nitric oxide modified growth, nutrient uptake and the antioxidant defense system in tomato seedlings stressed with arsenic. Theor. Exp. Plant Physiol. 2021, 33, 205–223. [Google Scholar] [CrossRef]
  9. Madeira, A.C.; de Varennes, A.; Abreu, M.M.; Esteves, C.; Magalhães, M.C.F. Tomato and parsley growth, arsenic uptake and translocation in a contaminated amended soil. J. Geochem. Explor. 2012, 123, 114–121. [Google Scholar] [CrossRef]
  10. Singh, V.K.; Upadhyay, R.S. Effects of arsenic on reactive oxygen species and antioxidant defense system in tomato plants. Toxicol. Environ. Chem. 2014, 96, 1374–1383. [Google Scholar] [CrossRef]
  11. Marmiroli, M.; Mussi, F.; Imperiale, D.; Lencioni, G.; Marmiroli, N. Abiotic stress response to As and As+Si, composite reprogramming of fruit metabolites in tomato cultivars. Front. Plant Sci. 2017, 8, 2201. [Google Scholar] [CrossRef] [PubMed]
  12. Tubana, B.S.; Babu, T.; Datnoff, L.E. A review of silicon in soils and plants and its role in us agriculture: History and future perspectives. Soil Sci. 2016, 181, 393–411. [Google Scholar] [CrossRef]
  13. Deng, Q.; Yu, T.; Zeng, Z.; Ashraf, U.; Shi, Q.; Huang, S.; Lian, T.; Chen, J.; Muzaffar, W.; Shen, W. Silicon application modulates the growth, rhizosphere soil characteristics, and bacterial community structure in sugarcane. Front. Plant Sci. 2021, 12, 710139. [Google Scholar] [CrossRef] [PubMed]
  14. Savvas, D.; Ntatsi, G. Biostimulant activity of silicon in horticulture. Sci. Hortic. (Amst.) 2015, 196, 66–81. [Google Scholar] [CrossRef]
  15. Baggio, G.; Dupas, E.; Galindo, F.S.; Megda, M.M.; Pereira, N.C.M.; Luchetta, M.O.; Tritapepe, C.A.; da Silva, M.R.; Jalal, A.; Teixeira Filho, M.C.M. Silicon application induced alleviation of aluminum toxicity in xaraés palisadegrass. Agronomy 2021, 11, 1938. [Google Scholar] [CrossRef]
  16. Ali, S.; Farooq, M.A.; Yasmeen, T.; Hussain, S.; Arif, M.S.; Abbas, F.; Bharwana, S.A.; Zhang, G. The influence of silicon on barley growth, photosynthesis and ultra-structure under chromium stress. Ecotoxicol. Environ. Saf. 2013, 89, 66–72. [Google Scholar] [CrossRef]
  17. Acharya, A.; Pal, P.K. Agriculture nanotechnology: Translating research outcome to field applications by influencing environmental sustainability. NanoImpact 2020, 19, 100232. [Google Scholar] [CrossRef]
  18. Juárez-Maldonado, A.; Ortega-Ortiz, H.; González-Morales, S.; Morelos-Moreno, Á.; Cabrera-de la Fuente, M.; Sandoval-Rangel, A.; Cadenas-Pliego, G.; Benavides-Mendoza, A. Nanoparticles and nanomaterials as plant biostimulants. Int. J. Mol. Sci. 2019, 20, 162. [Google Scholar] [CrossRef]
  19. Kandhol, N.; Aggarwal, B.; Bansal, R.; Parveen, N.; Singh, V.P.; Chauhan, D.K.; Sonah, H.; Sahi, S.; Grillo, R.; Peralta-Videa, J.; et al. Nanoparticles as a potential protective agent for arsenic toxicity alleviation in plants. Environ. Pollut. 2022, 300, 118887. [Google Scholar] [CrossRef]
  20. Sharma, A.; Vishwakarma, K.; Singh, N.K.; Prakash, V.; Ramawat, N.; Prasad, R.; Sahi, S.; Singh, V.P.; Tripathi, D.K.; Sharma, S. Synergistic action of silicon nanoparticles and indole acetic acid in alleviation of chromium (CrVI) toxicity in Oryza sativa seedlings. J. Biotechnol. 2021; in press. [Google Scholar] [CrossRef]
  21. Emamverdian, A.; Ding, Y.; Mokhberdoran, F.; Ahmad, Z.; Xie, Y. Determination of heavy metal tolerance threshold in a bamboo species (Arundinaria pygmaea) as treated with silicon dioxide nanoparticles. Glob. Ecol. Conserv. 2020, 24, e01306. [Google Scholar] [CrossRef]
  22. Ali, S.; Rizwan, M.; Hussain, A.; Zia ur Rehman, M.; Ali, B.; Yousaf, B.; Wijaya, L.; Alyemeni, M.N.; Ahmad, P. Silicon nanoparticles enhanced the growth and reduced the cadmium accumulation in grains of wheat (Triticum Aestivum L.). Plant Physiol. Biochem. 2019, 140, 1–8. [Google Scholar] [CrossRef] [PubMed]
  23. Farajzadeh Memari-Tabrizi, E.; Yousefpour-Dokhanieh, A.; Babashpour-Asl, M. Foliar-applied silicon nanoparticles mitigate cadmium stress through physio-chemical changes to improve growth, antioxidant capacity, and essential oil profile of summer savory (Satureja hortensis L.). Plant Physiol. Biochem. 2021, 165, 71–79. [Google Scholar] [CrossRef] [PubMed]
  24. Mahmoud, L.M.; Dutt, M.; Shalan, A.M.; El-Kady, M.E.; El-Boray, M.S.; Shabana, Y.M.; Grosser, J.W. Silicon nanoparticles mitigate oxidative stress of in vitro-derived banana (Musa acuminata ‘Grand Nain’) under simulated water deficit or salinity stress. S. Afr. J. Bot. 2020, 132, 155–163. [Google Scholar] [CrossRef]
  25. Siddiqui, M.H.; Al-Whaibi, M.H.; Mohammad, F. Rol of nanoparticles in plants. In Nanotechnology and Plant Sciences: Nanoparticles and Their Impact on Plants; Springer: Cham, Switzerland, 2015; pp. 19–35. ISBN 9783319145020. [Google Scholar]
  26. Cui, J.; Li, Y.; Jin, Q.; Li, F. Silica nanoparticles inhibit arsenic uptake into rice suspension cells: Via improving pectin synthesis and the mechanical force of the cell wall. Environ. Sci. Nano 2019, 7, 162–171. [Google Scholar] [CrossRef]
  27. González-Moscoso, M.; Juárez-Maldonado, A.; Cadenas-Pliego, G.; Meza-Figueroa, D.; SenGupta, B.; Martínez-Villegas, N. Silicon nanoparticles decrease arsenic translocation and mitigate phytotoxicity in tomato plants. Environ. Sci. Pollut. Res. 2022. [Google Scholar] [CrossRef]
  28. Steiner, A.A. A Universal method for preparing nutrient solutions of a certain desired composition. Plant Soil 1961, 15, 134–154. [Google Scholar] [CrossRef]
  29. DOF Norma Oficial Mexicana NOM-001-ECOL-1996, Que establece los límites máximos permisibles de contaminantes en las descargas de aguas residuales en aguas y bienes nacionales. D. Of. Fed. 1997, 1–14. Available online: https://dof.gob.mx/nota_detalle.php?codigo=4863829&fecha=06/01/1997#gsc.tab=0 (accessed on 7 April 2022).
  30. Le, V.N.; Rui, Y.; Gui, X.; Li, X.; Liu, S.; Han, Y. Uptake, Transport, distribution and bio-effects of SiO2 nanoparticles in Bt-transgenic cotton. J. Nanobiotechnol. 2014, 12, 50. [Google Scholar] [CrossRef]
  31. Nagata, M.; Yamashita, I. Simple method for simultaneous determination of chlorophyll and carotenoids in tomato fruit. Nippon. Shokuhin Kogyo Gakkaishi 1992, 39, 925–928. [Google Scholar] [CrossRef]
  32. Xue, T.; Hartikainen, H.; Piironen, V. Antioxidative and Growth-promoting effect of selenium on senescing lettuce. Plant Soil 2001, 237, 55–61. [Google Scholar] [CrossRef]
  33. Dhindsa, R.S.; Plumb-dhindsa, P.; Thorpe, T.A. Leaf senescence: Correlated with increased levels of membrane permeability and lipid peroxidation, and decreased levels of superoxide dismutase and catalase. J. Exp. Bot. 1981, 32, 93–101. [Google Scholar] [CrossRef]
  34. Nakano, Y.; Asada, K. Purification of ascorbate peroxidase in spinach chloroplasts; its inactivation in ascorbate-depleted medium and reactivation by monodehydroascorbate radical. Plant Cell Physiol. 1987, 28, 131–140. [Google Scholar] [CrossRef]
  35. Hung, C.Y.; Yen, G.C. Antioxidant activity of phenolic compounds isolated from Mesona procumbens Hemsl. J. Agric. Food Chem. 2002, 50, 2993–2997. [Google Scholar] [CrossRef] [PubMed]
  36. Arvouet-Grand, A.; Vennat, B.; Pourrat, A.; Legret, P. Standardization of propolis extract and identification of principal constituents. J. Pharm. Belg. 1994, 49, 462–468. [Google Scholar]
  37. Cumplido-Nájera, C.F.; González-Morales, S.; Ortega-Ortíz, H.; Cadenas-Pliego, G.; Benavides-Mendoza, A.; Juárez-Maldonado, A. The application of copper nanoparticles and potassium silicate stimulate the tolerance to Clavibacter michiganensis in tomato plants. Sci. Hortic. (Amst.) 2019, 245, 82–89. [Google Scholar] [CrossRef]
  38. Velikova, V.; Yordanov, I.; Edreva, A. Oxidative stress and some antioxidant systems in acid rain-treated bean plants. Plant Sci. 2000, 151, 59–66. [Google Scholar] [CrossRef]
  39. González-Moscoso, M.; Martínez-Villegas, N.V.; Cadenas-Pliego, G.; Benavides-Mendoza, A.; Del Carmen Rivera-Cruz, M.; González-Morales, S.; Juárez-Maldonado, A. Impact of silicon nanoparticles on the antioxidant compounds of tomato fruits stressed by arsenic. Foods 2019, 8, 612. [Google Scholar] [CrossRef]
  40. Re, R.; Pellegrini, N.; Proteggente, A.; Pannala, A. Antioxidant Activity applying an improved ABTS radical cation decolorization assay. Free Radic. Biol. 1999, 26, 1231–1237. [Google Scholar] [CrossRef]
  41. USDA. United States Standards for Grades of Fresh Tomatoes. United States Department of Agriculture. 1977; pp. 1–13. Available online: https://www.ams.usda.gov/grades-standards/tomato-grades-and-standards (accessed on 7 April 2022).
  42. López-Vargas, E.; Ortega-Ortíz, H.; Cadenas-Pliego, G.; de Alba Romenus, K.; Cabrera de la Fuente, M.; Benavides-Mendoza, A.; Juárez-Maldonado, A. Foliar application of copper nanoparticles increases the fruit quality and the content of bioactive compounds in tomatoes. Appl. Sci. 2018, 8, 1020. [Google Scholar] [CrossRef]
  43. Wang, H.B.; He, H.B.; Yang, G.D.; Ye, C.Y.; Niu, B.H.; Lin, W.X. Effects of two species of inorganic arsenic on the nutrient physiology of rice seedlings. Acta Physiol. Plant. 2010, 32, 245–251. [Google Scholar] [CrossRef]
  44. Vromman, D.; Martínez, J.P.; Kumar, M.; Šlejkovec, Z.; Lutts, S. Comparative effects of arsenite (As(III)) and arsenate (As(V)) on whole plants and cell lines of the arsenic-resistant halophyte plant species Atriplex atacamensis. Environ. Sci. Pollut. Res. 2018, 25, 34473–34486. [Google Scholar] [CrossRef] [PubMed]
  45. Garg, N.; Singla, P. Arsenic toxicity in crop plants: Physiological effects and tolerance mechanisms. Environ. Chem. Lett. 2011, 9, 303–321. [Google Scholar] [CrossRef]
  46. Gusman, G.S.; Oliveira, J.A.; Farnese, F.S.; Cambraia, J. Mineral nutrition and enzymatic adaptation induced by arsenate and arsenite exposure in lettuce plants. Plant Physiol. Biochem. 2013, 71, 307–314. [Google Scholar] [CrossRef] [PubMed]
  47. Meharg, A.; Wood, M.; Ripton, A. Integrated tolerance mechanisnfis: Constitutive and adaptive plant responses to elevated metal concentrations in the environment. Plant Cell Environ. 1994, 17, 989–993. [Google Scholar] [CrossRef]
  48. Finnegan, P.M.; Chen, W. Arsenic Toxicity: The effects on plant metabolism. Front. Physiol. 2012, 3, 182. [Google Scholar] [CrossRef]
  49. Pérez-Labrada, F.; López-Vargas, E.R.; Ortega-Ortiz, H.; Cadenas-Pliego, G.; Benavides-Mendoza, A.; Juárez-Maldonado, A. Responses of tomato plants under saline stress to foliar application of copper nanoparticles. Plants 2019, 8, 151. [Google Scholar] [CrossRef]
  50. Landa, P. Positive effects of metallic nanoparticles on plants: Overview of involved mechanisms. Plant Physiol. Biochem. 2021, 161, 12–24. [Google Scholar] [CrossRef]
  51. El-Batal, A.I.; Gharib, F.A.E.-L.; Ghazi, S.M.; Hegazi, A.Z.; Hafz, A.G.M.A. Physiological responses of two varieties of common bean (Phaseolus vulgaris L.) to foliar application of silver nanoparticles. Nanomater. Nanotechnol. 2016, 6, 13. [Google Scholar] [CrossRef]
  52. Mahakham, W.; Sarmah, A.K.; Maensiri, S.; Theerakulpisut, P. Nanopriming technology for enhancing germination and starch metabolism of aged rice seeds using phytosynthesized silver nanoparticles. Sci. Rep. 2017, 7, 8263. [Google Scholar] [CrossRef]
  53. Tripathi, D.K.; Singh, V.P.; Prasad, S.M.; Chauhan, D.K.; Dubey, N.K. Silicon nanoparticles (SiNp) alleviate chromium (VI) phytotoxicity in Pisum sativum (L.) seedlings. Plant Physiol. Biochem. 2015, 96, 189–198. [Google Scholar] [CrossRef]
  54. Cui, J.; Liu, T.; Li, F.; Yi, J.; Liu, C.; Yu, H. Silica nanoparticles alleviate cadmium toxicity in rice cells: Mechanisms and size effects. Environ. Pollut. 2017, 228, 363–369. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, L.; Ning, C.; Pan, T.; Cai, K. Role of silica nanoparticles in abiotic and biotic stress tolerance in plants: A review. Int. J. Mol. Sci. 2022, 23, 1947. [Google Scholar] [CrossRef] [PubMed]
  56. Gao, Z.; Chen, H.; Zhang, X.; Xiao, Z.; Fan, X.; Yin, C.; Tang, X.; Han, F.X.; Liang, Y. Silicon enhances abundances of reducing microbes in rhizoplane and decreases arsenite uptake by rice (Oryza sativa L.). Environ. Pollut. 2022, 306, 119405. [Google Scholar] [CrossRef] [PubMed]
  57. Ahire, M.L.; Mundada, P.S.; Nikam, T.D.; Bapat, V.A.; Penna, S. Multifaceted roles of silicon in mitigating environmental stresses in plants. Plant Physiol. Biochem. 2021, 169, 291–310. [Google Scholar] [CrossRef]
  58. Mehrabanjoubani, P.; Abdolzadeh, A.; Sadeghipour, H.R.; Aghdasi, M. Impacts of silicon nutrition on growth and nutrient status of rice plants grown under varying zinc regimes. Theor. Exp. Plant Physiol. 2015, 27, 19–29. [Google Scholar] [CrossRef]
  59. Nwugo, C.C.; Huerta, A.J. Effects of silicon nutrition on cadmium uptake, growth and photosynthesis of rice plants exposed to low-level cadmium. Plant Soil 2008, 311, 73–86. [Google Scholar] [CrossRef]
  60. Karimian, N.; Nazari, F.; Samadi, S. Morphological and biochemical properties, leaf nutrient content, and vase life of tuberose (Polianthes tuberosa L.) affected by root or foliar applications of silicon (Si) and silicon nanoparticles (SiNPs). J. Plant Growth Regul. 2021, 40, 2221–2235. [Google Scholar] [CrossRef]
  61. Wang, X.; Jiang, J.; Dou, F.; Sun, W.; Ma, X. Simultaneous mitigation of arsenic and cadmium accumulation in rice (Oryza sativa L.) seedlings by silicon oxide nanoparticles under different water management schemes. Paddy Water Environ. 2021, 19, 569–584. [Google Scholar] [CrossRef]
  62. Garbinski, L.D.; Rosen, B.P.; Chen, J. Pathways of arsenic uptake and efflux. Environ. Int. 2019, 126, 585–597. [Google Scholar] [CrossRef]
  63. Sandil, S.; Óvári, M.; Dobosy, P.; Vetési, V.; Endrédi, A.; Takács, A.; Füzy, A.; Záray, G. Effect of arsenic-contaminated irrigation water on growth and elemental composition of tomato and cabbage cultivated in three different soils, and related health risk assessment. Environ. Res. 2021, 197, 111098. [Google Scholar] [CrossRef]
  64. Ghorbani, A.; Pishkar, L.; Roodbari, N.; Pehlivan, N.; Wu, C. Nitric oxide could allay arsenic phytotoxicity in tomato (Solanum lycopersicum L.) by modulating photosynthetic pigments, phytochelatin metabolism, molecular redox status and arsenic sequestration. Plant Physiol. Biochem. 2021, 167, 337–348. [Google Scholar] [CrossRef] [PubMed]
  65. Stazi, S.R.; Cassaniti, C.; Marabottini, R.; Giuffrida, F.; Leonardi, C. Arsenic uptake and partitioning in grafted tomato plants. Hortic. Environ. Biotechnol. 2016, 57, 241–247. [Google Scholar] [CrossRef]
  66. Tlustoš, P.; Száková, J.; Pavlíková, D.; Balík, J. The response of tomato (Lycopersicon esculentum) to different concentrations of inorganic and organic compounds of arsenic. Biologia 2006, 61, 91–96. [Google Scholar] [CrossRef]
  67. Carbonell-Barrachina, A.A.; Burló, F.; Burgos-Hernández, A.; López, E.; Mataix, J. The influence of arsenite concentration on arsenic accumulation in tomato and bean plants. Sci. Hortic. (Amst.) 1997, 71, 167–176. [Google Scholar] [CrossRef]
  68. Jia-Yi, Y.; Meng-Qiang, S.; Zhi-Liang, C.; Yu-Tang, X.; Hang, W.; Jian-Qiang, Z.; Ling, H.; Qi, Z. Effect of foliage applied chitosan-based silicon nanoparticles on arsenic uptake and translocation in rice (Oryza sativa L.). J. Hazard. Mater. 2022, 433, 128781. [Google Scholar] [CrossRef] [PubMed]
  69. Bhat, J.A.; Shivaraj, S.M.; Singh, P.; Navadagi, D.B.; Tripathi, D.K.; Dash, P.K.; Solanke, A.U.; Sonah, H.; Deshmukh, R. Role of silicon in mitigation of heavy metal stresses in crop plants. Plants 2019, 8, 71. [Google Scholar] [CrossRef] [PubMed]
  70. Praveen, A.; Mehrotra, S.; Singh, N.; Pandey, V. Nutrient constraints in arsenic phytoremediation. Russ. J. Plant Physiol. 2018, 65, 15–22. [Google Scholar] [CrossRef]
  71. Li, J.; Yang, Z.; Song, G.; Zhang, X. Correlations of arsenic and nutrient elements in different tissues of perennial ryegrass under arsenic stress. J. Soil Sci. Plant Nutr. 2021, 21, 1542–1551. [Google Scholar] [CrossRef]
  72. Garg, N.; Kashyap, L. Silicon and rhizophagus irregularis: Potential candidates for ameliorating negative impacts of arsenate and arsenite stress on growth, nutrient acquisition and productivity in Cajanus cajan (L.) Millsp. genotypes. Environ. Sci. Pollut. Res. 2017, 24, 18520–18535. [Google Scholar] [CrossRef]
  73. Sanchez-Navarro, J.F.; Gonzalez-García, Y.; Benavides-mendoza, A.; Morales-Días, A.B.; Gonzalez-Morales, S.; Cadenas-pliego, G.; Garcia-Guillermo, M. del S.; Juárez-Maldonado, A. Silicon nanoparticles improve the shelf life and antioxidant status of lilium. Plants 2021, 10, 2338. [Google Scholar] [CrossRef] [PubMed]
  74. Alsaeedi, A.; El-Ramady, H.; Alshaal, T.; El-Garawany, M.; Elhawat, N.; Al-Otaibi, A. Silica nanoparticles boost growth and productivity of cucumber under water deficit and salinity stresses by balancing nutrients uptake. Plant Physiol. Biochem. 2019, 139, 1–10. [Google Scholar] [CrossRef]
  75. González-Moscoso, M.; Martínez-Villegas, N.V.; Meza-Figueroa, D.; Rivera-Cruz, M. del C.; Cadenas-Pliego, G.; Juárez-Maldonado, A. SiO2 nanoparticles improve nutrient uptake in tomato plants developed in the presence of arsenic. Rev. Bio Cienc. 2021, 8, e1084. [Google Scholar] [CrossRef]
  76. de Sousa, A.; Saleh, A.M.; Habeeb, T.H.; Hassan, Y.M.; Zrieq, R.; Wadaan, M.A.M.; Hozzein, W.N.; Selim, S.; Matos, M.; AbdElgawad, H. Silicon dioxide nanoparticles ameliorate the phytotoxic hazards of aluminum in maize grown on acidic soil. Sci. Total Environ. 2019, 693, 133636. [Google Scholar] [CrossRef]
  77. González de las Torres, A.I.; Giráldez, I.; Martínez, F.; Palencia, P.; Corns, W.T.; Sánchez-Rodas, D. Arsenic accumulation and speciation in strawberry plants exposed to inorganic arsenic enriched irrigation. Food Chem. 2020, 315, 126215. [Google Scholar] [CrossRef] [PubMed]
  78. Sghaier, D.B.; Duarte, B.; Bankaji, I.; Caçador, I.; Sleimi, N. Growth, chlorophyll fluorescence and mineral nutrition in the halophyte Tamarix gallica cultivated in combined stress conditions: Arsenic and NaCl. J. Photochem. Photobiol. B Biol. 2015, 149, 204–214. [Google Scholar] [CrossRef] [PubMed]
  79. Song, R.; Zhao, C.Y.; Liu, J.; Zhang, J.; Du, Y.X.; Li, J.Z.; Sun, H.Z.; Zhao, H.B.; Zhao, Q.Z. Effect of sulphate nutrition on arsenic translocation and photosynthesis of rice seedlings. Acta Physiol. Plant. 2013, 35, 3237–3243. [Google Scholar] [CrossRef]
  80. Songy, A.; Vallet, J.; Gantet, M.; Boos, A.; Ronot, P.; Tarnus, C.; Clément, C.; Larignon, P.; Goddard, M.L.; Fontaine, F. Sodium arsenite effect on Vitis vinifera L. Physiology. J. Plant Physiol. 2019, 238, 72–79. [Google Scholar] [CrossRef]
  81. Abedi, T.; Mojiri, A. Arsenic uptake and accumulation mechanisms in rice species. Plants 2020, 9, 129. [Google Scholar] [CrossRef]
  82. Sun, S.K.; Chen, Y.; Che, J.; Konishi, N.; Tang, Z.; Miller, A.J.; Ma, J.F.; Zhao, F.J. Decreasing arsenic accumulation in rice by overexpressing OsNIP1;1 and OsNIP3;3 through disrupting arsenite radial transport in roots. New Phytol. 2018, 219, 641–653. [Google Scholar] [CrossRef]
  83. Praveen, A.; Khan, E.; Ngiimei, S.; Perwez, M.; Sardar, M.; Gupta, M. Iron oxide nanoparticles as nano-adsorbents: A possible way to reduce arsenic phytotoxicity in indian mustard plant (Brassica juncea L.). J. Plant Growth Regul. 2018, 37, 612–624. [Google Scholar] [CrossRef]
  84. Kataria, S.; Jain, M.; Rastogi, A.; Živčák, M.; Brestic, M.; Liu, S.; Tripathi, D.K. Role of nanoparticles on photosynthesis: Avenues and applications. In Nanomaterials in Plants, Algae and Microorganisms: Concepts and Controversies: Volume 2; Elsevier: Amsterdam, The Netherlands, 2018; Volume 6, pp. 103–127. ISBN 9780128114889. [Google Scholar]
  85. Mukarram, M.; Khan, M.M.A.; Corpas, F.J. Silicon nanoparticles elicit an increase in lemongrass (Cymbopogon flexuosus (steud.) wats) agronomic parameters with a higher essential oil yield. J. Hazard. Mater. 2021, 412, 125254. [Google Scholar] [CrossRef] [PubMed]
  86. Mukarram, M.; Petrik, P.; Mushtaq, Z.; Khan, M.M.A.; Gulfishan, M.; Lux, A. Silicon nanoparticles in higher plants: Uptake, action, stress tolerance, and crosstalk with phytohormones, antioxidants, and other signalling molecules. Environ. Pollut. 2022, 310, 119855. [Google Scholar] [CrossRef] [PubMed]
  87. Yap, C.K.; Tan, W.S.; Wong, K.W.; Ong, G.H.; Cheng, W.H.; Nulit, R.; Ibrahim, M.H.; Chew, W.; Berandah Edward, F.; Okamura, H.; et al. Antioxidant enzyme activities as biomarkers of Cu and Pb stress in Centella asiatica. Stresses 2021, 1, 253–265. [Google Scholar] [CrossRef]
  88. Souza, L.A.; Piotto, F.A.; Dourado, M.N.; Schmidt, D.; Franco, M.R.; Boaretto, L.F.; Tezotto, T.; Ferreira, R.R.; Azevedo, R.A. Physiological and biochemical responses of Dolichos Lablab L. to cadmium support its potential as a cadmium phytoremediator. J. Soils Sediments 2015, 17, 1413–1426. [Google Scholar] [CrossRef]
  89. Duquesnoy, I.; Champeau, G.M.; Evray, G.; Ledoigt, G.; Piquet-Pissaloux, A. Enzymatic adaptations to arsenic-induced oxidative stress in Zea mays and genotoxic effect of arsenic in root tips of Vicia faba and Zea mays. Comptes Rendus Biol. 2010, 333, 814–824. [Google Scholar] [CrossRef]
  90. Srivastava, S.; Mishra, S.; Tripathi, R.D.; Dwivedi, S.; Trivedi, P.K.; Tandon, P.K. Phytochelatins and antioxidant systems respond differentially during arsenite and arsenate stress in Hydrilla verticillata (L.f.) Royle. Environ. Sci. Technol. 2007, 41, 2930–2936. [Google Scholar] [CrossRef]
  91. Sharma, I. Arsenic induced oxidative stress in plants. Biologia 2012, 67, 447–453. [Google Scholar] [CrossRef]
  92. Tripathi, D.K.; Singh, S.; Singh, V.P.; Prasad, S.M.; Chauhan, D.K.; Dubey, N.K. Silicon nanoparticles more efficiently alleviate arsenate toxicity than silicon in maize cultiver and hybrid differing in arsenate tolerance. Front. Environ. Sci. 2016, 4, 46. [Google Scholar] [CrossRef]
  93. Shabnam, N.; Kim, M.; Kim, H. Iron (III) oxide nanoparticles alleviate arsenic induced stunting in Vigna radiata. Ecotoxicol. Environ. Saf. 2019, 183, 109496. [Google Scholar] [CrossRef]
  94. Gill, S.S.; Tuteja, N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiol. Biochem. 2010, 48, 909–930. [Google Scholar] [CrossRef] [PubMed]
  95. Benavides-Mendoza, A.; Gonzalez-Moscoso, M.; Ojeda-Barrios, D.L.; Fuentes-lara, L.O. Biostimulation and toxicity: Two levels of action of nanomaterials in plants. In Nanotechnology in Plant Growth Promotion and Protection: Recent Advances and Impacts; John Wiley and Sons Ltd.: Hoboken, NJ, USA, 2021; pp. 283–303. [Google Scholar]
  96. García-López, J.I.; Niño-Medina, G.; Olivares-Sáenz, E.; Lira-Saldivar, R.H.; Barriga-Castro, E.D.; Vázquez-Alvarado, R.; Rodríguez-Salinas, P.A.; Zavala-García, F. Foliar application of zinc oxide nanoparticles and zinc sulfate boosts the content of bioactive compounds in habanero peppers. Plants 2019, 8, 254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kasote, D.M.; Katyare, S.S.; Hegde, M.V.; Bae, H. Significance of antioxidant potential of plants and its relevance to therapeutic applications. Int. J. Biol. Sci. 2015, 11, 982–991. [Google Scholar] [CrossRef]
  98. Requejo, R.; Tena, M. Influence of glutathione chemical effectors in the response of maize to arsenic exposure. J. Plant Physiol. 2012, 169, 649–656. [Google Scholar] [CrossRef] [PubMed]
  99. Li, L.; Zeng, X.; Williams, P.N.; Gao, X.; Zhang, L.; Zhang, J.; Shan, H.; Su, S. Arsenic resistance in fungi conferred by extracellular bonding and vacuole-septa compartmentalization. J. Hazard. Mater. 2021, 401, 123370. [Google Scholar] [CrossRef] [PubMed]
  100. Wang, H.B.; Wong, M.H.; Lan, C.Y.; Qin, Y.R.; Shu, W.S.; Qiu, R.L.; Ye, Z.H. Effect of arsenic on flavonoid contents in pteris species. Biochem. Syst. Ecol. 2010, 38, 529–537. [Google Scholar] [CrossRef]
  101. Drzewiecka, K.; Gąsecka, M.; Rutkowski, P.; Magdziak, Z.; Goliński, P.; Mleczek, M. Arsenic forms and their combinations induce differences in phenolic accumulation in Ulmus laevis Pall. J. Plant Physiol. 2018, 220, 34–42. [Google Scholar] [CrossRef]
  102. Woźniak, A.; Drzewiecka, K.; Kȩsy, J.; Marczak, Ł.; Narozna, D.; Grobela, M.; Motała, R.; Bocianowski, J.; Morkunas, I. The influence of lead on generation of signalling molecules and accumulation of flavonoids in pea seedlings in response to pea aphid infestation. Molecules 2017, 22, 1404. [Google Scholar] [CrossRef]
  103. Rahman, A.; Mostofa, M.G.; Alam, M.M.; Nahar, K.; Hasanuzzaman, M.; Fujita, M. Calcium mitigates arsenic toxicity in rice seedlings by reducing arsenic uptake and modulating the antioxidant defense and glyoxalase systems and stress markers. Biomed Res. Int. 2015, 2015, 340812. [Google Scholar] [CrossRef]
  104. Moradbeygi, H.; Jamei, R.; Heidari, R.; Darvishzadeh, R. Investigating the enzymatic and non-enzymatic antioxidant defense by applying iron oxide nanoparticles in Dracocephalum moldavica L. plant under salinity stress. Sci. Hortic. (Amst.) 2020, 272, 109537. [Google Scholar] [CrossRef]
  105. Nourozi, E.; Hosseini, B.; Maleki, R.; Abdollahi Mandoulakani, B. Iron oxide nanoparticles: A novel elicitor to enhance anticancer flavonoid production and gene expression in Dracocephalum kotschyi hairy-root cultures. J. Sci. Food Agric. 2019, 99, 6418–6430. [Google Scholar] [CrossRef] [PubMed]
  106. Jurkow, R.; Agnieszka, S.; Pokluda, R.; Smole, S.; Kalisz, A. Biochemical Response of Oakleaf Lettuce Seedlings to Different Concentrations of Some Metal(oid) Oxide Nanoparticles. Agronomy 2020, 10, 997. [Google Scholar] [CrossRef]
  107. Nair, P.M.G.; Chung, I.M. Assessment of silver nanoparticle-induced physiological and molecular changes in Arabidopsis thaliana. Environ. Sci. Pollut. Res. 2014, 21, 8858–8869. [Google Scholar] [CrossRef] [PubMed]
  108. Shi, Z.; Yang, S.; Han, D.; Zhou, Z.; Li, X.; Liu, Y.; Zhang, B. Silicon alleviates cadmium toxicity in wheat seedlings (Triticum aestivum L.) by reducing cadmium ion uptake and enhancing antioxidative capacity. Environ. Sci. Pollut. Res. 2018, 25, 7638–7646. [Google Scholar] [CrossRef] [PubMed]
  109. Dawidowicz, A.L.; Olszowy, M. Antioxidant properties of BHT estimated by ABTS assay in systems differing in pH or metal ion or water concentration. Eur. Food Res. Technol. 2011, 232, 837–842. [Google Scholar] [CrossRef]
  110. Siddiqi, K.S.; Husen, A. Engineered gold nanoparticles and plant adaptation potential. Nanoscale Res. Lett. 2016, 11, 400. [Google Scholar] [CrossRef]
  111. Marslin, G.; Sheeba, C.J.; Franklin, G. Nanoparticles alter secondary metabolism in plants via ROS burst. Front. Plant Sci. 2017, 8, 832. [Google Scholar] [CrossRef]
  112. Muthuramalingam, M.; Matros, A.; Scheibe, R.; Mock, H.P.; Dietz, K.J. The hydrogen peroxide-sensitive proteome of the chloroplast in vitro and in vivo. Front. Plant Sci. 2013, 4, 54. [Google Scholar] [CrossRef]
  113. Nour, V.; Ionica, M.E.; Trandafir, I. Bioactive compounds, antioxidant activity and color of hydroponic tomato fruits at different stages of ripening. Not. Bot. Horti Agrobot. Cluj-Napoca 2015, 43, 404–412. [Google Scholar] [CrossRef]
  114. Marmiroli, M.; Mussi, F.; Imperiale, D.; Marmiroli, N. Target proteins reprogrammed by As and As+Si treatments in Solanum lycopersicum L. Fruit. BMC Plant Biol. 2017, 17, 210. [Google Scholar] [CrossRef]
  115. Marmiroli, M.; Pigoni, V.; Savo-Sardaro, M.L.; Marmiroli, N. The effect of silicon on the uptake and translocation of arsenic in tomato (Solanum lycopersicum L.). Environ. Exp. Bot. 2014, 99, 9–17. [Google Scholar] [CrossRef]
  116. Marmiroli, M.; Mussi, F.; Imperiale, D.; Marmiroli, N. Arsenic and Silicon in tomato fruit: An “omics” analysis for food safety. In Proceedings of the IPC2019 CHANGSHA The 16th International Phytotechnology Conference—Phytotechnologies for Food Safety & Environmental Health, Changsha, China, 23–27 September 2019; p. 178. [Google Scholar]
  117. Christou, A.; Georgiadou, E.C.; Zissimos, A.M.; Christoforou, I.C.; Christofi, C.; Neocleous, D.; Dalias, P.; Ioannou, A.; Fotopoulos, V. Uptake of hexavalent chromium by tomato (Solanum lycopersicum L.) plants and mediated effects on their physiology and productivity, along with fruit quality and safety. Environ. Exp. Bot. 2021, 189, 104564. [Google Scholar] [CrossRef]
  118. Dai, Y.; Li, T.; Wang, Z.; Xing, B. Physiological and proteomic analyses reveal the effect of CeO2 Nanoparticles on strawberry reproductive system and fruit quality. Sci. Total Environ. 2022, 814, 152494. [Google Scholar] [CrossRef]
  119. Neysanian, M.; Iranbakhsh, A.; Ahmadvand, R.; Ardebili, Z.O.; Ebadi, M. Comparative efficacy of selenate and selenium nanoparticles for improving growth, productivity, fruit quality, and postharvest longevity through modifying nutrition, metabolism, and gene expression in tomato; potential benefits and risk assessment. PLoS ONE 2020, 15, e0244207. [Google Scholar] [CrossRef]
  120. Ogunkunle, C.O.; Adegboye, E.F.; Okoro, H.K.; Vishwakarma, V.; Alagarsamy, K.; Fatoba, P.O. Effect of nanosized anatase TiO2 on germination, stress defense enzymes, and fruit nutritional quality of Abelmoschus esculentus (L.) Moench (Okra). Arab. J. Geosci. 2020, 13, 120. [Google Scholar] [CrossRef]
  121. Elsheery, N.I.; Helaly, M.N.; El-Hoseiny, H.M.; Alam-Eldein, S.M. Zinc oxide and silicone nanoparticles to improve the resistance mechanism and annual productivity of salt-stressed mango trees. Agronomy 2020, 10, 558. [Google Scholar] [CrossRef]
  122. Karagiannis, E.; Michailidis, M.; Skodra, C.; Molassiotis, A.; Tanou, G. Silicon influenced ripening metabolism and improved fruit quality traits in apples. Plant Physiol. Biochem. 2021, 166, 270–277. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Effect of two forms of inorganic As (AsV-AsIII) and SiO2 NPs on the growth and yield of tomato plants. (A) Dry shoot biomass, (B) dry root biomass, and (C) fruit yield. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Different letters per bar indicate significant differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Figure 1. Effect of two forms of inorganic As (AsV-AsIII) and SiO2 NPs on the growth and yield of tomato plants. (A) Dry shoot biomass, (B) dry root biomass, and (C) fruit yield. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Different letters per bar indicate significant differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Agronomy 12 02366 g001
Figure 2. Silicon concentration in (A) roots, (B) leaves, and (C) fruits. Arsenic concentration in (D) roots, (E) leaves, and (F) arsenic translocation factor. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. ND: no detected. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Figure 2. Silicon concentration in (A) roots, (B) leaves, and (C) fruits. Arsenic concentration in (D) roots, (E) leaves, and (F) arsenic translocation factor. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. ND: no detected. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Agronomy 12 02366 g002
Figure 3. Macronutrients (AE) and micronutrients (FH) in leaves, roots, and fruits of tomato plants exposed to AsV and AsIII in irrigation water and SiO2 NPs via substrate. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. ND: no detected. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Figure 3. Macronutrients (AE) and micronutrients (FH) in leaves, roots, and fruits of tomato plants exposed to AsV and AsIII in irrigation water and SiO2 NPs via substrate. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. ND: no detected. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Agronomy 12 02366 g003
Figure 4. Chlorophylls content in leaves of tomato plants. (A) chlorophyll a, (B) chlorophyll b, and (C) total chlorophyll increase when plants are exposed to the combination of AsV-AsIII with SiO2 NPs. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Figure 4. Chlorophylls content in leaves of tomato plants. (A) chlorophyll a, (B) chlorophyll b, and (C) total chlorophyll increase when plants are exposed to the combination of AsV-AsIII with SiO2 NPs. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Agronomy 12 02366 g004
Figure 5. Antioxidant enzyme activity in the roots and leaves of tomato plants. (A) APX: ascorbate peroxidase, (B) GPX: glutathione peroxidase, (C) CAT: catalase, and (D) SOD: superoxide dismutase. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Figure 5. Antioxidant enzyme activity in the roots and leaves of tomato plants. (A) APX: ascorbate peroxidase, (B) GPX: glutathione peroxidase, (C) CAT: catalase, and (D) SOD: superoxide dismutase. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Agronomy 12 02366 g005
Figure 6. Non-enzymatic antioxidant compounds in tomato roots and leaves exposed to two forms of inorganic As and SiO2 NPs. (A) Vitamin C, (B) phenols, (C) flavonoids, and (D) glutathione. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case different letters per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Figure 6. Non-enzymatic antioxidant compounds in tomato roots and leaves exposed to two forms of inorganic As and SiO2 NPs. (A) Vitamin C, (B) phenols, (C) flavonoids, and (D) glutathione. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case different letters per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Agronomy 12 02366 g006
Figure 7. Hydrogen peroxide (H2O2) and ABTS antioxidant capacity in roots, leaves, and fruits of tomato plants. (A) H2O2 levels, (B) antioxidant capacity ABTS in roots, (C) antioxidant capacity ABTS in leaves, and (D) antioxidant capacity ABTS in fruits. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Figure 7. Hydrogen peroxide (H2O2) and ABTS antioxidant capacity in roots, leaves, and fruits of tomato plants. (A) H2O2 levels, (B) antioxidant capacity ABTS in roots, (C) antioxidant capacity ABTS in leaves, and (D) antioxidant capacity ABTS in fruits. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Agronomy 12 02366 g007
Figure 8. Quality of tomato fruits exposed to two forms of inorganic As and SiO2 NPs, (A) Firmness, (B) total soluble solids, (C) hydrogen potential, (D) lycopene content, (E) β-carotene content, (F) vitamin C, (G) glutathione, (H) flavonoids, and (I) phenols. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Figure 8. Quality of tomato fruits exposed to two forms of inorganic As and SiO2 NPs, (A) Firmness, (B) total soluble solids, (C) hydrogen potential, (D) lycopene content, (E) β-carotene content, (F) vitamin C, (G) glutathione, (H) flavonoids, and (I) phenols. Plants treated with 3.2 mg L−1 of AsV (Na2HAsO4∙7H2O) and AsIII (NaAsO2) via irrigation through the nutrient solution and 1000 mg L−1 of SiO2 NPs via substrate. Lower-case letters different per bar indicate statistical differences according to the Fisher’s least significant difference test (p ≤ 0.05). n = 5 ± standard error.
Agronomy 12 02366 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

González-Moscoso, M.; Martínez-Villegas, N.; Cadenas-Pliego, G.; Juárez-Maldonado, A. Effect of Silicon Nanoparticles on Tomato Plants Exposed to Two Forms of Inorganic Arsenic. Agronomy 2022, 12, 2366. https://doi.org/10.3390/agronomy12102366

AMA Style

González-Moscoso M, Martínez-Villegas N, Cadenas-Pliego G, Juárez-Maldonado A. Effect of Silicon Nanoparticles on Tomato Plants Exposed to Two Forms of Inorganic Arsenic. Agronomy. 2022; 12(10):2366. https://doi.org/10.3390/agronomy12102366

Chicago/Turabian Style

González-Moscoso, Magín, Nadia Martínez-Villegas, Gregorio Cadenas-Pliego, and Antonio Juárez-Maldonado. 2022. "Effect of Silicon Nanoparticles on Tomato Plants Exposed to Two Forms of Inorganic Arsenic" Agronomy 12, no. 10: 2366. https://doi.org/10.3390/agronomy12102366

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