Next Article in Journal
Identification of Multiple Pancreatic and Extra-Pancreatic Pathways Underlying the Glucose-Lowering Actions of Acacia arabica Bark in Type-2 Diabetes and Isolation of Active Phytoconstituents
Next Article in Special Issue
Assessment of Nutritional and Quality Properties of Leaves and Musts in Three Local Spanish Grapevine Varieties Undergoing Controlled Climate Change Scenarios
Previous Article in Journal
Special Issue: “Plant Virus Epidemiology”
Previous Article in Special Issue
Physiological Alteration in Sunflower Plants (Helianthus annuus L.) Exposed to High CO2 and Arbuscular Mycorrhizal Fungi
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Foliar Nourishment with Nano-Selenium Dioxide Promotes Physiology, Biochemistry, Antioxidant Defenses, and Salt Tolerance in Phaseolus vulgaris

1
Botany Department, Faculty of Agriculture, Fayoum University, Fayoum 63514, Egypt
2
Botany Department, Faculty of Agriculture, Zagazig University, Zagazig 44519, Egypt
3
Department of Biology, College of Science, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
4
Botany and Microbiology Department, Faculty of Science, Zagazig University, Zagazig 44519, Egypt
*
Author to whom correspondence should be addressed.
Plants 2021, 10(6), 1189; https://doi.org/10.3390/plants10061189
Submission received: 17 May 2021 / Revised: 3 June 2021 / Accepted: 4 June 2021 / Published: 11 June 2021
(This article belongs to the Special Issue Plant Physiological Responses to Climate Change)

Abstract

:
Novel strategic green approaches are urgently needed to raise the performance of plants subjected to stress. Two field-level experimental attempts were implemented during two (2019 and 2020) growing seasons to study the possible effects of exogenous nourishment with selenium dioxide nanoparticles (Se-NPs) on growth, physio-biochemical ingredients, antioxidant defenses, and yield of Phaseolus vulgaris (L.) plant growing on a salt-affected soil (EC = 7.55–7.61 dS m−1). At 20, 30, and 40 days from seeding, three foliar sprays were applied to plants with Se-NPs at a rate of 0.5, 1.0, or 1.5 mM. The experimental design was accomplished in randomized complete plots. The data indicate noteworthy elevations in indicators related to growth and yield; pigments related to effective photosynthesis, osmoprotectant (free proline and soluble sugars), nutrient and Se contents, K+/Na+ ratio, cell integrity (water content and stability of membranes), all enzyme activities; and all features related to leaf anatomy induced by Se-NPs foliar spray. Conversely, marked lowering in markers of Na+ content-induced oxidative stress (superoxide radical and hydrogen peroxide) and their outcomes in terms of ionic leakage and malondialdehyde were reported by foliar nourishment with Se-NPS compared to spraying leaves with water as an implemented control. The best results were recorded with Se-NPs applied at 1.0 mM, which mitigated the negative effects of soil salinity (control results). Therefore, the outcomes of this successful study recommend the use of Se-NPs at a rate of 1.0 mM as a foliar spray to grow common beans on saline soils with EC up to 7.55–7.61 dS m−1.

1. Introduction

Common bean (Phaseolus vulgaris L.) is an important legume crop for human food due to its wide range of production of seeds rich in protein content [1,2,3]. It is classified as a sensitive crop to salt stress [4]. It is produced under soil salinity conditions in Latin America (approximately 5–10%) and the Middle East (approximately 20–30%) [5].
Salinity stress is a major issue of concern for workers in the agricultural sector. It negatively influences crop production directly or indirectly. Many plants are vulnerable to salinity effects and are unequal to withstand low levels of salinity [2]. Irrigation with water of poor quality and/or lack of quantity is one of the key factors that accumulate salts in the soil, which discourages plant metabolism and physical and chemical process, and, consequently, plant growth and productivity [6]. Salinity reduces the chance of plant roots absorbing enough water due to the physiological drought stimulated by salinity, which limits the plants’ growth and productivity performance by inhibiting the metabolism process [7,8]. Plant metabolism is discouraged with salinity due to osmotic stress and Na+ and Cl ions toxicity, which contribute to the inhibition of plant growth, various physio-biochemical attributes, and outputs by the overproduction of some species of reactive oxygen (ROS): O2•–, H2O2, and OH [7,8,9,10]. Physiological drought, toxic influences of undesirable ions (Na+ and Cl), and nutrient imbalance are key adverse effects that occur to plants growing on salt-affected soil [6,11]. Salt accumulations in leafy apoplasm to toxic scale lead to loss of cell turgor and cell shrinkage, leading to cell death. The physiological process most affected under salt stress is photosynthesis. This process is associated with a marked reduction in the chlorophyll content and closing of stomata, and thus the efficiency of photosynthesis [6,12,13].
Because of the aforementioned negative salt influences, it is important to use applicable approaches to enhance salinity tolerance and alleviate the deleterious influences of salinity stress [14]. Exogenous use of organic or inorganic substances is used to boost the productivity of common beans under salinity stress [15]. Selenium (Se) is a microelement essential for plant performance at low concentration, but it is toxic in high concentration for plants, and thus humans, animals, and microorganisms [16]. With a higher concentration (more than 100 mg per kg in plant tissues), Se toxicity is induced by the substitution of sulfur (S) with Se in amino acids, resulting in the production of nonfunctional proteins and enzymes. However, at lower concentrations (up to 16 mg per L), Se can boost plant tolerability, delay senescence, suppress oxidative stress, and promote aging seedling growth [17,18].
Recently, nanomaterials have become a desirable solution to many technological and environmental challenges in numerous fields [19]. Taking into account nanotechnology, agricultural sustainability, and ecological issues, it is imperative to examine the various approaches in which nanoproducts may alter plant metabolism associated with growth and development. At present, selenium nanoparticles (Se-NPs) have been introduced as highly stable nanoparticles to be used in the medical industry and as fertilizers in the agricultural and food industries [20]. Compared to selenates or selenites, Se-NPs have lower toxicity and higher bioactivity [21]. Se-NPs have been used as a foliar application recently to protect plants subjected to severe stress through improved antioxidant defense mechanisms [21], photosynthetic indices, and secondary metabolism [22]; however, there is still a long way to go to explain the effects of exogenous Se-NPs supplies on these traits in plants affected by salinity. Research on the performance, mechanism, and effect of nanoparticles such as Se-NPs in plants, as well as agricultural application, is still in the rudimentary stage.
Based on the foregoing, the present study aimed to define the fruitful functions of the exogenous application of Se-NPs in improving salt tolerance in salt-affected common bean plants by exploring the potential enhancing impacts of Se-NPs on plant performance in terms of growth and yield traits, as well as plant physiology and biochemistry.

2. Materials and Methods

2.1. Experimental Layout and Growing Conditions

The two seasons 2019 and 2020 were specified to conduct two field trials on a special farm at El-Noubaria (31.15651, 29.86418), Egypt. For each season, planting began on February 25 and May 10 was the end time for the experiments. This period (25 February to 10 May) of the 2019 and 2020 seasons did not have any rainy days. Before starting each trial, soil samples were taken from the experimental farm site for analyzing the physicochemical and fertility status using the procedures described in [23,24], and the obtained data are shown in Table 1. Based on these analyses, the site’s soil had an ECe value of 7.55–7.61 dS m1 in soil paste extract for both seasons (Table 1), which classified the experimental soil as saline, as stated in [25].
As the most salt-sensitive Phaseolus vulgaris (L.) in Egypt, Bronco was identified for this study. The Egyptian Agricultural Research Center was the source from which healthy seeds of similar size, color, and disease-free were obtained. The seed surface was sterilized by rinsing in NaClO for 10 min, washed well with distilled water (d.W), and dried in the open air for 2 h. Sowing was implemented in late February for both trial seasons at 95 kg per ha to reach the seeding rate recommended for the commercial production of this crop. The seeds were grown on 10.5 m2 plots (long = 3.0 m × width = 3.50 m) in hills (4 seeds per hill, and 15–20 cm between every two hills) in rows spaced 60 cm apart. Immediately before the first watering, two healthy, strong seedlings were kept in each hill. Prior to field planting, 280 kg ha1 (NH4)2SO4 (20% N) and 350 kg ha1 Ca2+-superphosphate (15.5% P2O5) were added to the soil.
All treatment plots were organized in the open conditions of the selected special farm and the transplants/plants were preserved under natural climatic conditions (e.g., the mean temperatures and relative humidity were 31 ± 3 °C/17 ± 3 °C for day/night (approximately 12 h for each) and 62–66%, respectively). One level of foliar d.W was accomplished as a control, and there were three levels of Se-NPs (0.5, 1.0, and 1.5 mM) with three replicates per treatment. The experimental diagrams of all treatments were organized into a completely randomized design (CRD).

2.2. Foliar Application of Nano-Selenium (Se-NPs)

The Se-NPs was purchased from Sigma Aldrich Co. (Sigma-Aldrich, 99.5% purity of nano-SiO2) with properties shown in Table 2.
When the plants reached 25 days old, they were sprayed with d.W (as a control), 0.5, 1.0, or 1.5 mM Se-NPs according to the treatments designed for this study. The spraying of d.W and the different concentrations of Se-NPs was repeated twice (i.e., the second and third sprays) when the plants were 35 and 45 days old, respectively. The concentrations of Se-NPs were selected based on our preliminary pot study, in which 11 concentrations (e.g., 0, 0.5, 1.0, 1.5, 2.0, 2.5, 3.0, 3.5, 4.0, 4.5, and 5.0) were examined. The concentrations 0.5, 1.0, and 1.5 conferred the best results (data not shown), and therefore they were selected for this main field study. Using a pressurized spray bottle, plants were foliar sprayed to run off. The volume sprayed of each solution was 25, 30, and 40 mL per plant at the three ages (25, 35, and 45 days old, respectively) with 5, 6, and 8 L of spray solution per 10.5 m2 plot considering 200 plants for the three ages, respectively. An appropriate surfactant (e.g., few drops of Tween-20) was applied to the spraying solutions.

2.3. Sampling Date and Sample Preparation for Different Determinations

After 8 weeks of planting, 9 plants were harvested indiscriminately from the middle rows in each plot (10.5 m2) of each of the four treatments in both seasons (2019 and 2020). After cleaning plant shoots using a water-filled bucket to remove any adhering dusts, the indices of plant growth, physiology, biochemistry, oxidative stress, and enzymatic activities, as well as leaf anatomy (measured only in the 2020 season), were assessed. In the merchantable green pod’s stage (10–11 weeks after planting), pods were reaped from 50 indiscriminately selected plants from the two outer rows of each trial plot to assess the indices of green yield. Data are displayed as an average of nine plants as three replicates.

2.4. Assessment of Growth and Green Yield Traits

After separating plants, shoots were subjected for taking lengths of shoots (cm), leaf number on each plant, area of leaves of each plant (cm2), and shoot dry weight (DW-Sh; g). DW-Sh was assessed after drying at 70 °C until two or three constant weights. Pods number for each plant and yield of green pods per ha (ton) were recorded.

2.5. Evaluation of Indicators of Plant Physiology and Biochemistry

Fresh tissue of a fully extended upper leaf devoid of the midribs was extracted using acetone (80%, v/v) to determine chlorophylls (a and b) and carotenoids [26]. The resulting supernatants were subjected to the monitoring of their optical densities (at wavelengths of 480, 645, and 663 nm) on a spectrophotometer apparatus.
Utilizing the same leafy material, the rate of both the transpiration (Tr) and net photosynthesis (Pn) of the photosynthetic system was detected using a portable photosynthesis system (LF6400XTR, LI-COR, USA). During 09:00–11:00 a.m., all measurements were recorded.
The methods described in [9,27,28,29,30,31,32,33] were utilized to determine each of proline, total soluble sugars, hydrogen peroxide (H2O2), superoxide (O2•−), peroxidation of membrane lipid level (evaluated as malondialdehyde level (MDA)), leakage of ions (EL), stability index of membranes (MSI), and relative content of tissue water (RWC), using a fully extended upper leaf devoid of the midribs.
Using a fully extended upper leaf devoid of the midribs, a constant number of 20 discs was specified to assess the total inorganic ions escaping from the leaves. EC1 (electrical conductivity) was recorded in the solution of the discs before heating. EC2 was recorded after heating at 45–55 °C for 0.5 h. Then, EC3 was recorded after boiling for 10 min. EL percentage was obtained using the following equation:
EL (%) = [(EC2 − EC1)/EC3] × 100
Using a fully extended upper leaf devoid of the midribs, pieces with a constant weight of 0.2 g were used to determine the MSI. EC1 was recorded after heating the solution of 0.2 g sample on 40 °C for 0.5 h. EC2 was recorded after boiling the solution of another sample for 10 min. By applying the following equation, MSI percentage was obtained:
MSI (%) = [1 − (EC1/EC2)] × 100
Using a fully extended upper leaf devoid of the midribs, a constant number of 2 cm diameter discs was specified to evaluate RWC. Immediately, fresh mass was recorded after the discs were weighed. Then, in the dark, the discs were water-saturated for an entire day. After gently removing the adhering water, the recording of the turgid mass was accomplished. After drying (at 70 °C until constant weights), the discs dry mass was taken. RWC (%) was recorded using this equation:
RWC (%) = [(fresh mass – dry mass)/(turgid mass − dry mass)] × 100
To analyze nutrient (N, P, and K+) and Na+ contents, fully extended upper leaf samples were dried at 70 °C until reaching constant weights. After grinding, a 1:3 mixture of perchloric and nitric acids, respectively, was applied to digest the dried samples. Total N was determined by applying the micro-Kjeldahl method [34]. Total phosphorus was determined by colormetrically applying the ascorbic acid method [35]. The same solution was applied to determine K+ and Na+ ion contents using atomic absorption spectrophotometry [36].

2.6. Determination of Enzyme Activities

Using a fully extended upper leaf devoid of the midribs (0.5 g), the enzymatic extract was prepared and used as the supernatant obtained from the centrifugation (12,000× g, 4 °C, 0.25 h) of the leafy homogenate to assay enzyme activities (unit: mg−1 protein). The methods in [37,38,39,40] were utilized to assay catalase, peroxidase, ascorbate peroxidase, and glutathione reductase activities, respectively.
A frozen sample (500 mg) was homogenized in a mortar and pestle fixed on ice. The solution of homogenization was HEPES buffer (10 mL, 50 mM) and 0.l mM Na2EDTA (pH 7.6). After performing centrifugation (15,000× g for 15 min at 4 °C) to obtain the resulting homogenates, the resulting extract was applied to assay the concentration of protein and the activity of superoxide dismutase (SOD; unit, mg−1 protein). Overnight, the extract was dialyzed against a dilute homogenizing solution to separate the low-molecular-mass substances that interfere in the SOD assay. The protein–dye binding procedure [41] was used to measure the concentration of soluble protein. The procedure of Yu and Rengel [42] was used for the SOD activity assay. This procedure is established on the basis of observation of photochemical reduction inhibition of NBT (nitro blue tetrazolium). To assay total SOD, a 5 mL reaction mixture of 50 mM, 0.l mM, 50 mM, 13 mM, 0.025% (w/v), 75 µM, 2 µM, 0.2 mL for HEPES (pH 7.6), EDTA, Na2CO3 (pH 10.4), methionine, Triton X-l00, NBT, riboflavin, and enzyme extract, was prepared. Using a light intensity of 350 µM m−2 s−1, the mixture was illuminated for 15 min. One SOD activity unit was specified as an enzyme quantity that causes a 50% NBT reduction inhibition (observed spectrophotometrically on 560 nm).

2.7. Leaf Anatomy

Leaf specimens were secured from the middle internode with its leaf blade. The selected specimens were chosen from plants at the flowering stage for killing and fixing for 48 h in 100 mL of F.A.A. solution containing 50 mL of C2H5OH (95%), 5 mL of glacial acetic acid, and 10 mL of formalin, in addition to 35 mL of distilled water. Then, samples were exposed to washing using C2H5OH (50%). Dehydration and clearance were then performed with normal butyl alcohol series, and the resulting samples were embedded in paraffin wax (54–56 °C m.p). A rotary microtome was functioned to cut samples for 20 μm thick cross-sections that adhered (Haupt’s adhesive). The samples were then stained [43,44]. Slide photography was performed and then read with a light microscope equipped with a digital camera (canon power shot S80) connected to computer; the photographs were taken by zoom browsers ex. program. Sections were examined to detected histological manifestations of the chosen treatments and photomicrographed. The obtained leaf anatomical features were expressed in µm (x = 200 µm).

2.8. Analysis of the Resulting Data

Differences among all the resulting data means were compared at p ≤ 0.05 by the Duncan multiple range test. Analyses were performed, statistically, by applying COSTAT software (version 6.303, Berkeley, CA, USA).

3. Results

3.1. Growth, Productivity, and Photosynthetic Efficiency Responses of Salt-Stressed Common Bean to Foliar Nourishment with Se-NPs

The data presented in Table 3 and Table 4 indicate the lowest common bean plant growth (number and area of leaves per plant and length and dry weight of shoot per plant), yield (number of pods per plant and yield of green pods per hectare), and photosynthetic efficiency (leaf pigments (chlorophylls and carotenoids) and rate of both net photosynthesis (Pn) and transpiration (Tr)) as shown in the treatment of comparison (control, foliar spray with distilled water (d.W)), which reflected the negative effect of soil salinity. However, foliar nourishment with Se-NPs positively affected plant performances, as it led to significant elevations in the indices related to plant growth, productivity, and efficiency of photosynthesis comparing with the comparison treatment. The best-applied rate for Se-NPs was 1 mM, which maximized the results of all the growth and yield indices mentioned above. The same outcome trends were gained over both seasons (2019 and 2020).

3.2. Plant Tissue Cell Integrity Response of Salt-Stressed Common Bean to Foliar Nourishment with Se-NPs

The comparison treatment, indicating salt stress and spraying leaves with d.W, presented adverse effects of salt stress such as reduced relative content of water (RWC), stability index of cell membranes (MSI), and contents of free proline, soluble sugars, and endogenous Se and raised levels of markers of oxidative stress (O2 and H2O2), leakage of electrolytes (EL), and peroxidation of membrane lipids (MDA) (Table 5 and Figure 1). However, leafy nourishment with Se-NPs positively modified these tested parameters of tissue cell integrity, as it resulted in significant promotion of RWC, MSI, free proline, soluble sugars, and Se contents due to minimized levels of oxidative stress markers, EL, and MDA comparing with the comparison treatment. The best-applied rate for Se-NPs was 1 mM, which preserved plant tissue cell integrity and increased endogenous Se to the desired optimal content. Both seasons (2019 and 2020) had the same outcome trends.

3.3. Enzyme Activity and Nutrient Content Responses of Salt-Stressed Common Bean to Foliar Nourishment with Se-NPs

Salt stress and spraying leaves with d.W (the comparison treatment) displayed the lowest activity of CAT, POX, APX, SOD, and GR; the lowest content of N, P, and K+; the lowest ratio of K+/Na+; and increased Na+ content (Figure 2 and Figure 3). However, foliar feeding with Se-NPs positively modified the tested enzyme activity and nutrient content, as this greatly improved all of the above parameters along with a considerable reduction in the Na+ content and a considerable rise in the ratio of K+/Na+ compared to the comparison treatment. The best-applied rate for Se-NPs was 1 mM, which sustained the plant antioxidative system and nutrient homeostasis. Both seasons (2019 and 2020) had the same outcome trends.

3.4. Anatomical Features Responses of Salt-Stressed Common Bean Leaf to Foliar Nourishment with Se-NPs

The data presented in Table 6 and Figure 4 indicate the lowest features of leaf anatomy in the common bean plant, as shown in the treatment of comparison (control; foliar spray with d.W), which reflected the negative effect of soil salinity. However, foliar nourishment with Se-NPs positively affected leaf anatomical features, as it led to significant elevations in all features of leaf anatomy compared to the comparison treatment. The best-applied rate for Se-NPs was 1 mM, which maximized the results of all features of leaf anatomy. It increased the thickness of each of blade, palisade tissue, spongy tissue, phloem, and xylem by 92.1%, 145.5%, 111.4%, 108.4%, and 100.3%, respectively, along with increasing midvein length and width by 85.2% and 107.3%, respectively. It also increased the diameter of the vessel by 118.7% compared to the comparison treatment.

4. Discussion

The agricultural sector in arid and semi-arid regions, including Egypt, faces many difficulties, including soil salinity. Salinity causes physiological drought even in the presence of water due to high osmotic pressure, which impedes plant growth and productivity due to disturbance of cell metabolism, division, elongation, meristematic activity, pigmentation, photosynthesis-related efficiency, stomatal closure, ionic imbalance, and gaseous exchange, along with elevated respiration rate and oxidative stress markers, including hydrogen peroxide (H2O2), superoxide (O2•−) radical, and hydroxyl anion (OH), due to the failure of antioxidant defenses to withstand salt stress [14,45,46,47,48,49,50].
Plants possess an antioxidative system to mitigate the damaging effects of ROS [51]. This system contains many active ingredients with low-molecular mass (non-enzymatic antioxidants) and high-molecular mass (antioxidant enzymes) that synergistically act with osmoprotectant compounds [52,53]. However, under severe stress, the natural tolerance of plants does not help with stress resistance since the active ingredients of the plant with low-molecular mass and high-molecular mass do not provide sufficient requirements for defense against stress. Therefore, exogenous adjuvants, including beneficial nutrients in nanoparticles and other auxiliary strategies, must be used to raise the efficiency of antioxidant defense in stressed plants to enable them to function effectively under different stresses [7,10,54,55,56,57,58,59], including salinity [6,60,61,62].
In the present study, the exogenous adjuvant was selenium nanoparticles (Se-NPs) applied as a foliar spray, which enabled common bean plants to perform effectively under soil salinity conditions (EC = 7.55–7.61 dS m−1; Table 1). Se-NPs enhanced the growth and production of salt-stressed common bean plants by promoting the aqueous state of plant cells, delaying plant aging [63,64], and improving the size and/or the number of plant cells. These distinct results may be attributed to the maintenance of cell turgidity, leading to cell elongation [65], increased rigidity of mature leaves [66,67], and improved all photosynthesis-related attributes [68], thus promoting the efficiency of photosynthesis [69]. Therefore, Se-NPs, applied as a foliar application, could be an effective biostimulator to promote tolerability of common bean plants under conditions of salinity stress [70]. Several works have used Se-NPs to improve plant growth under various stresses in different crops such as sorghum under heat stress [71], tomato under salinity stress [72], wheat under drought stress [73], wheat under heat stress [64], and maize under salinity stress [65]. Low concentrations of Se-NPs have been used to stimulate plant growth and yield; however, high doses cause toxic effects [65,67]. In this concern, Hawrylak-Nowak [74] found that application of low concentration of Se (5 μmol·dm−3) stimulates plant growth and root elongation, but high concentrations (50 and 100 μmol·dm−3) of Se leads to decreased growth, biomass, and root tolerance index resulting from high phosphorous (P) accumulation in the straw of plants. Our resulting data demonstrate that Se-NPs applied at 1.5 mM reduced growth, yield, and all biochemical indices compared to 1.0 mM, indicating that concentrations of Se-NPs greater than 1.0 mM caused damage to salt-stressed Phaseolus vulgaris plants.
Oxidative stress induced by salinity causes damage to plasma membranes and chlorophyll due to an excess of H2O2 and O2•− [15]. It is associated with raised uptake of Na+ and limited accumulation of Mg2+, which affects the synthesis of the chlorophylls [75]. Subjecting common bean plants to salinity-induced adverse conditions induces changes in the role of the complex of pigment protein [75]. Salt accumulation limits pigment synthesis by minimizing the activity of Mg-chelatase, 5-aminolevulinic acid dehydratase, porphobilinogen deaminase, protochlorophyllide oxidoreductase, and porphyrinogen IX oxidase [76]. These undistinguished findings are associated with increased activity of chlorophyllase [77], diminished leaf water potential, reduced uptake of N, and thus minimized efficiency of photosynthesis [78,79]. However, Alyemeni et al. [80] reported that Se application helps eliminate reactive oxygen species (ROS) and nutrient build-up due to improved photosynthesis. Additionally, Elkelish et al. [77] documented that exogenous Se application enhanced chlorophyll and carotenoid pigments and conferred a positive effect on stomatal conductance and photosynthetic efficiency. Besides, Iqbal et al. [64] showed that applying the optimal amount of Se leads to an increase in chlorophyll and protects the chloroplast structure from oxidative damage. The decrease in oxidative damage occurs due to the positive effect of Se on the chlorophyll content and its role in the structure of chloroplasts [65,81]. The Se-induced improvement in photosynthesis has also been reported in tomato plants that interfered with salinity [72]. Se had a considerable potency for maintaining the ultrastructure of mitochondria and chloroplasts to promote photosynthesis capacity and acclimatize to salt stress [81]. The Se-NPs exhibit a profitable influence on the performance of the photosystem in stressed plants [21,82,83].
Principally, Rios et al. [84] documented that the decrease of photosynthesis under salinity stress occurs due to oxidative and osmotic stress and nutritional imbalance. Our resulting findings reveal that foliar nourishment with 1 mM Se-NPs as an optimal level increased photosynthetic effectiveness, transpiration rate (Tr), net photosynthesis rate (Pn), and leaf gas exchange in salinity-stressed Phaseolus vulgaris plants (Table 4). Optimum supplementation of Se boosted the performance of photosynthesis by promoting indices of chlorophyll fluorescence, assimilation of CO2, and photosynthesis rate under the stress conditions of salinity [80,85]. Our data show that Se-NPs applied at 1 mM significantly enhanced cell integrity (i.e., RWC and MSI) under salt stress (Table 5) by increasing hydraulic conductivity of roots which caused an increase in the flow of water from roots to shoots of plants [86,87,88,89]. Therefore, Se-NPs application eliminates the toxic effect of salinity and increases RWC by increasing water flow from roots to shoots. The increased RWC by Se-NPs is associated with an accumulation of sugar content that reflects positively on plant productivity under stress conditions [90]. Our results also show that Se-NPs increased proline accumulation that improved the effectiveness of photosynthesis, production of ATP, and efficiency of water use by plants, as formerly reported [91]. Proline is a low-molecular-mass antioxidant that has an important role in the osmotic modification of salinity-stressed plant cells [92]. It reduces ROS damage and promotes the tolerability of plants by minimizing ROS detoxification under salt stress [93]. Besides, proline improves plant antioxidant system efficiency [94]. Likewise, Chandrasekhar and Sandhyarani [95] stated that the accumulation of proline assists plants replenish energy under salinity conditions and increases their survival under stress. Se-NPs application regulates the accumulation of proline and increases enzymatic activity and protects photosynthesis by improving RWC and Rubisco protection along with preserving proline and soluble sugar contents to help clean up ROS [78].
Similar to proline, soluble sugar accumulation keeps the regularity between the osmotic quality of the cellular cytosol and the vacuole [96]. Se appears to be able to positively modulate accumulations of soluble sugars and proline in plant leaves under salinity stress due to their significant roles in maintaining osmotic and ionic balance in plant cells, thus leading to stress tolerance [77,78]. There are additional mechanisms to minimize ROS damage and promote plant tolerance by proline. Among these mechanisms, it reduces the influences of salinity stress by cleaning up ROS generated by salinity poisoning and may react with hydroxyl radicals, directly, or quench singlet oxygen in a physical manner [93].
Our resulting data indicate the determination of the peroxidation level of membrane lipids as the content of MDA, and this is rated as a stress biochemical indicator, as it minimizes biomass production and decreases the plant’s stress adaptation [97]. Our data reveal that the control treatment (soil salinity with foliar spray with distilled water) exceeded all Se-NPs at O2•− and H2O2 levels as markers of oxidative stress and their unwanted consequences such as increased electrolyte leakage (EL) and MDA levels (Figure 1). This realistic result was demonstrated by the high ROS concentrations generated in the control treatment plants. Commonly, the levels of peroxidation of membrane lipids and EL in higher plants through free radicals encourage damage to cell membranes or retrogradation when plants undergo ecological stresses [98]. Under these conditions, acceptance of electrons by NADP will be restricted, thus O2 can act as an electron acceptor, resulting in the production of more ROS, which cause peroxidation of cell membranes and increased EL [99,100]. However, the application of Se-NPs significantly controlled O2•− and H2O2 (Figure 1) and reduced MDA and EL levels (Figure 1). Salinity stress causes oxidative cell damage through an imbalance in ROS generation and alters antioxidant activity [101]. To avert the oxidative damage caused by salt stress, plants possess a well-developed antioxidative system, with low- (non-enzymatic) and high-molecular-mass (enzymatic) active ingredients. Among the enzymes, SOD (first line of defense) defends against O2•− [102] by converting it to H2O2, which in turn is converted to O2 and H2O by peroxidases in the presence of APX, which is the most important enzyme in reducing ROS [103,104]. GR and DHAR as regenerating enzymes have been described as a base portion of the Halliwell–Asada cycle. They form a portion of AsA regeneration from DHA using GSH as a reducing power [105]. The present study indicated that Se-NPs treatments, especially that applied at 1 mM, significantly increased antioxidant enzyme activities (e.g., CAT, APX, GR, SOD, and POX) compared to the control treatment (Figure 2). These increased enzyme activities upregulate the AsA-GSH pathway, protecting the photoelectron transport chain by respecting toxic radical formation and NADP level maintenance. Besides, applying Se reduces ROS and promotes plant growth by stimulating various antioxidants and modified osmoprotectant levels. Likewise, former studies have documented that applying Se increases the activity of antioxidant enzymes in different crops [80,81,82].
The data from our study show minimal accumulation of Na+ in common bean plants due to the application of Se-NPs, thus reducing levels of oxidative stress markers induced by salinity (Figure 3). Zhang et al. [106] hypothesized that the application of Se-NPs elevates the expression of Na+/H+ antiport and tonoplast H+-ATPase at the root membranes, limiting the translocation of Na+ ions to the upper plant tissues, thus limiting Na+ toxic effects. Our data also show that the application of Se-NPs significantly increased the uptake and thus the content of N, P, and K+ to promote the regulation of plant growth by inference to antioxidant metabolism, cellular stress signaling, and nitrogen assimilation [78,107,108]. Besides, high N, P, and K+ uptake and minimization of Na+ level in common bean plants lead to improved stress signaling and amino acid and metabolite production, thus boosting plant tolerance to salinity [77]. As another important mechanism, Astaneh et al. [109] stated that spraying leaves with Se leads to Na+ and K+ homeostasis under salt stress due to that Se plays a positive role at the level of membrane transport and improves expression of the ZmNHX1 gene, thus improving K+ uptake and reducing Na+ accumulation under salt stress [17,110]. This improved ratio of K+/Na+ has been shown to be beneficial for protecting vital processes and maintaining osmotic balance [17,111]. In plant roots, the ZmMPK5, ZmMPK7, and ZmCPK11 genes are upregulated by Se under osmotic stress. Besides, overexpression and upregulation of ZmNHX1 and NHX genes in some transgenic plant species are responsible for Na+ compartmentalization and enhanced salt resistance [17,65,112]. In addition, OsNHX1 (vacuolar Na+/H+ antiporter gene) is responsible for maintaining plant osmotic balance by reducing Na+ ions obstruction during water movement towards plant shoots [17,112], which may be attributed to the sequestration of Na+ ions in vacuoles of roots and/or shoots [17,113]. It could be visualized that the higher OsNHX1 transcriptional level promoted Na+ sequestration within the root vacuoles, thus reducing Na+ ion accumulation in the plant shoot, which ultimately promoted plant growth and antioxidant defense mechanisms [17].
Our resulting data show an increase in endogenous Se as a result of exogenous feeding with Se-NPs, and the optimum endogenous content of Se (28.2–29.4 mg kg−1 DW) was obtained by exogenous application of 1 mM Se-NPs (Table 5). The promoted growth of Phaseolus vulgaris plants by externally used Se-NPs could be a positive measure of salt tolerance associated with increased plant productivity, and thus the plants produced more metabolites, which were essential for plant growth. Foliar nourishment with Se-NPs improved salinity stress tolerance in Phaseolus vulgaris plants through optimal growth and productivity due to the optimal content of endogenous Se. These desirable findings are attributed to the restoration of the damaged chloroplast structure along with stimulation of chlorophyll biosynthesis and increased photosynthetic capacity and antioxidant activities due to the osmoprotection and osmoregulation role of Se [114,115]. The osmoregulation role of Se is effectively increased through its upregulation of choline/choline monooxygenase biosynthesis, which catalyzes glycine betaine biosynthesis [116]. Thus, Se plays some crucial roles, as it helps in stabilizing and maintaining membrane integrity and keeps the turgor of plant cells under stress conditions of salinity.
Our findings also show that the application of Se-NPs improved the leaf anatomical features in salinity-stressed Phaseolus vulgaris plants, indicating that exogenous Se-NPs recovered the deleterious effects of salt stress on the features related to leaf anatomy. This improvement in features related to leaf anatomy awarded an opportunity for a good translocation of assimilates along with absorbed nutrients into cells for use in various metabolic processes that are reflected in a positive manner in robust growth and satisfying productions under salt stress conditions. Leafy nourishment with Se-NPs could improve the protective tissue development in Phaseolus vulgaris leaves, thus strengthening their anti-dehydration capacity. Similar anatomical findings have been obtained previously [3,55,117].
Under salinity stress, our resulting data display that low- and high-molecular-mass antioxidant activity was sufficiently increased by application of Se-NPs (at a concentration of 1.0 mM), in conjunction with higher RWC, MSI, stomatal conductance, and photosynthesis efficiency and decreased EL and MDA, thus reducing cell membrane injury. Finally, our data suggest that Se helps minimize salt stress influences in Phaseolus vulgaris plants by promoting antioxidative activity. In addition, there is evidence that the resulting yields from plants treated with Se are found to be safe for human health [118].

5. Conclusions

The results obtained indicate that the decreased parameters (pigments and other indices related to photosynthetic efficiency, osmoprotectants, various antioxidants, and nutrients) related to Phaseolus vulgaris plant growth and yield obtained under salty field conditions (EC = 7.55–7.61 dS m−1) were restored with nourishment with Se-NPs applied as a foliar spray for plants, especially at 1 mM. Se-NPs application ameliorated photosynthetic efficiency, antioxidant defense system, and osmoregulation and promoted plant growth and productivity. Thus, foliar nourishment with Se-NPs provided a noticeable role in alleviating the adverse effects of salt stress on all indices related to Phaseolus vulgaris plant growth, physiology, biochemistry, and green yield. Therefore, leafy nourishment with Se-NPs can be recommended as a noteworthy strategic approach to promote the growth and productivity of the Phaseolus vulgaris plant under soil salinity stress at the open field level.

Author Contributions

Conceptualization, M.M.R. and E.-S.M.D.; data curation, M.M.R., E.-S.M.D., S.M.A., and E.S.; formal analysis, M.M.R., E.-S.M.D., S.M.A., A.M., E.F.A., and E.S.; investigation, M.M.R., E.-S.M.D., S.M.A., A.M., E.F.A., S.M.A.I.A., and E.S.; methodology, M.M.R., E.-S.M.D., S.M.A., A.M., E.F.A., S.M.A.I.A., and E.S.; resources, M.M.R., E.-S.M.D., S.M.A., A.M., E.F.A., S.M.A.I.A., and E.S.; software, M.M.R., E.-S.M.D., S.M.A., A.M., E.F.A., S.M.A.I.A., and E.S.; writing—original draft, M.M.R., E.-S.M.D., S.M.A., A.M., S.M.A.I.A., and E.S.; and writing—review and editing, M.M.R., E.-S.M.D., and E.F.A. All authors have read and agreed to the published version of the manuscript.

Funding

The Deanship of Scientific Research at Taif University through the research number TURSP-2020/110 is acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

The authors are thankful to the Taif University Researchers Supporting Project number (TURSP-2020/110), Taif University, Taif, Saudi Arabia for providing the financial support and research facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Broughton, W.J.; Hernandez, G.; Blair, M.; Beebe, S.; Gepts, P.; Vanderleyden, J. Beans (Phaseolus spp.)—Model food legumes. Plant Soil 2003, 252, 55–128. [Google Scholar] [CrossRef] [Green Version]
  2. Rady, M.M.; Bhavya Varma, C.; Howladar, S.M. Common bean (Phaseolus vulgaris L.) seedlings overcome NaCl stress as a result of presoaking in Moringa oleifera leaf extract. Sci. Hortic. 2013, 162, 63–70. [Google Scholar] [CrossRef]
  3. Sitohy, M.Z.; Desoky, E.M.; Osman, A.; Rady, M.M. Pumpkin seed protein hydrolysate treatment alleviates salt stress effects on Phaseolus vulgaris by elevating antioxidant capacity and recovering ion homeostasis. Sci. Hortic. 2020, 271, 109495. [Google Scholar] [CrossRef]
  4. Mass, E.V.; Hoffman, G.J. Crop salt tolerance. Current assessment. J. Irrig. Drain. 1977, 103, 115–134. [Google Scholar] [CrossRef]
  5. Centro Internacional de Agricultura Tropical (CIAT). Contraints to and opportunities for improving bean production. In A Planting Document 1993–98 and Achieving Document 1987–92; CIAT: Cali, Colombia, 1992. [Google Scholar]
  6. Seleiman, M.F.; Semida, W.M.; Rady, M.M.; Mohamed, G.F.; Hemida, K.A.; Alhammad, B.A.; Hassan, M.M.; Shami, A. Sequential Application of Antioxidants Rectifies Ion Imbalance and Strengthens Antioxidant Systems in Salt-Stressed Cucumber. Plants 2020, 9, 1783. [Google Scholar] [CrossRef]
  7. ElSayed, A.I.; Boulila, M.; Rafudeen, M.S.; Mohamed, A.H.; Sengupta, S.; Rady, M.; Omar, A.A. Melatonin Regulatory Mechanisms and Phylogenetic Analyses of Melatonin Biosynthesis Related Genes Extracted from Peanut under Salinity Stress. Plants 2020, 9, 854. [Google Scholar] [CrossRef]
  8. Desoky, E.M.; Mansour, E.-S.; Yasin, M.A.T.; El-Sobky, E.E.A.; Rady, M.M. Improvement of drought tolerance in five different cultivars of Vicia faba with foliar application of ascorbic acid or silicon. Span. J. Agric. Res. 2020, 18, e0802. [Google Scholar] [CrossRef]
  9. Rady, M.M. Effect of 24-epibrassinolide on growth, yield, antioxidant system and cadmium content of bean (Phaseolus vulgaris L.) plants under salinity and cadmium stress. Sci. Hortic. 2011, 129, 232–237. [Google Scholar] [CrossRef]
  10. Taha, R.S.; Seleiman, M.F.; Alotaibi, M.; Alhammad, B.A.; Rady, M.M.; Mahdi, A.H.A. Exogenous Potassium Treatments Elevate Salt Tolerance and Performances of Glycine max L. by Boosting Antioxidant Defense System under Actual Saline Field Conditions. Agronomy 2020, 10, 1741. [Google Scholar] [CrossRef]
  11. Marschner, H. Mineral Nutrition of Higher Plants, 2nd ed.; Academic Press: London, UK, 1995; ISBN 012-473-5-436. [Google Scholar]
  12. Sudhir, P.; Murthy, S.D.S. Effect of salt stress on basic process of photosynthesis. Photosynthetica 2004, 42, 481–486. [Google Scholar] [CrossRef]
  13. Desoky, E.M.; EL-Maghraby Lamiaa, M.M.; Awad, A.E.; Abdo, A.I.; Rady, M.M.; Semida, W.M. Fennel and ammi seed extracts modulate antioxidant defence system and alleviate salinity stress in cowpea (Vigna unguiculata). Sci. Hortic. 2020, 272, 109576. [Google Scholar] [CrossRef]
  14. Rady, M.M.; Elrys, A.S.; Abo El-Maati, M.F.; Desoky, E.M. Interplaying roles of silicon and proline effectively improve salt and cadmium stress tolerance in Phaseolus vulgaris plant. Plant Physiol. Biochem. 2019, 139, 558–568. [Google Scholar] [CrossRef]
  15. Rady, M.M.; Desoky, E.M.; Elrys, A.S.; Boghdady, M.S. Can licorice root extract be used as an effective natural biostimulant for salt-stressed common bean plants? S. Afr. J. Bot. 2019, 121, 294–305. [Google Scholar] [CrossRef]
  16. Shardendu, S.N.; Boulyga, S.F.; Stengel, E. Phytoremediation of selenium by two helophyte species in subsurface flow constructed wetland. Chemosphere 2003, 50, 967–973. [Google Scholar] [CrossRef]
  17. Kamran, M.; Parveen, A.; Ahmar, S.; Hussain, S.; Chattha, M.S.; Saleem, M.H.; Adil, M.; Heidari, P.; Chen, J. An Overview of Hazardous Impacts of Soil Salinity in Crops, Tolerance Mechanisms, and Amelioration through Selenium Supplementation. Int. J. Mol. Sci. 2020, 21, 148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Pennanen, A.; Xue, T.; Hartikainen, H. Protective role of selenium in plant subjected to severe UV irradiation stress. J. Appl. Bot. 2002, 76, 66–76. [Google Scholar]
  19. Sharma, C.S.; Nema, R.K.; Sharma, V.K. Synthesis, anticonvulsant activity and in silico study of some novel amino acids incorporated bicyclo compounds. Stamford J. Pharm. Sci. 2009, 2, 42–47. [Google Scholar] [CrossRef]
  20. Babajani, A.; Iranbakhsh, A.; Ardebili, Z.O.; Eslami, B. Differential growth, nutrition, physiology, and gene expression in Melissa officinalis mediated by zinc oxide and elemental selenium nanoparticles. Environ. Sci. Pollut. Res. 2019, 26, 24430–24444. [Google Scholar] [CrossRef]
  21. Djanaguiraman, M.; Belliraj, N.; Bossmann, S.H.; Prasad, P.V. High-temperature stress alleviation by selenium nanoparticle treatment in grain sorghum. ACS Omega 2018, 3, 2479–2491. [Google Scholar] [CrossRef] [Green Version]
  22. Zsiros, O.; Nagy, V.; Párducz, Á.; Nagy, G.; Ünnep, R.; El-Ramady, H.; Prokisch, J.; Lisztes-Szabó, Z.; Fári, M.; Csajbók, J.; et al. Effects of selenate and red Se-nanoparticles on the photosynthetic apparatus of Nicotiana tabacum. Photosynth. Res. 2019, 139, 449–460. [Google Scholar] [CrossRef] [PubMed]
  23. Black, C.A. Soil Plant Relationships, 2nd ed.; John Wiley and Sons: Hoboken, NJ, USA, 1968. [Google Scholar]
  24. Jackson, M.L. Soil Chemical Analysis; Prentice Hall, Inc.: Englewood Califfs, NJ, USA, 1973. [Google Scholar]
  25. Dahnke, W.C.; Whitney, D.A. Measurement of soil salinity. In Recommended Chemical Soil Test Procedures for the North Central Region, 499; North Central Regional Publication 221; Dahnke, W.C., Ed.; North Dakota Agricultural Experiment Station: Fargo, ND, USA, 1988; pp. 32–34. [Google Scholar]
  26. Arnon, D.I. Copper enzymes in isolated chloroplasts, polyphenoxidase in Beta vulgaris. Plant Physiol. 1949, 24, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Bates, L.S.; Waldren, R.P.; Teare, I.D. Rapid determination of free proline for water stress studies. Plant Soil 1973, 39, 205–207. [Google Scholar] [CrossRef]
  28. Irigoyen, J.J.; Emerich, D.W.; Sanchez-Diaz, M. Water stress induced changes in the concentrations of proline and total soluble sugars in nodulated alfalfa (Medicago sativa) plants. Plant Physiol. 1992, 8, 455–460. [Google Scholar] [CrossRef]
  29. 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]
  30. Kubis, J. Exogenous spermidine differentially alters activities of some scavenging system enzymes, H2O2 and superoxide radical levels in water-stressed cucumber leaves. J. Plant Physiol. 2008, 165, 397–406. [Google Scholar] [CrossRef] [PubMed]
  31. Madhava Rao, K.V.; Sresty, T.V.S. Antioxidative parameters in the seedlings of pigeonpea (Cajanus cajan (L.) Millspaugh) in response to Zn and Ni stresses. Plant Sci. 2000, 157, 113–128. [Google Scholar] [CrossRef]
  32. Dionisio-Sese, M.L.; Tobita, S. Antioxidant responses of rice seedlings to salinity stress. Plant Sci. 1998, 135, 1–9. [Google Scholar] [CrossRef]
  33. Osman, A.S.; Rady, M.M. Effect of humic acid as an additive to growing media to enhance the production of eggplant and tomato transplants. J. Hortic. Sci. Biotechnol. 2014, 89, 237–244. [Google Scholar]
  34. Chapman, H.D.; Pratt, F.P. Determination of Minerals by Titration Method: Methods of Analysis for Soils, Plants and Water, 2nd ed.; Agriculture Division, California University: San Diego, CA, USA, 1982; pp. 169–170. [Google Scholar]
  35. Watanabe, F.S.; Olsen, S.R. Test of ascorbic acid method for determine phosphorus in water and NaHCO3 extracts from soil. Soil Sci. Soc. Am. Proc. 1965, 29, 677–678. [Google Scholar] [CrossRef]
  36. Emilio, A.C.; Francisco, P.A.; Vicente, M.; Manuel, C.; Maria, C.B. Evaluation of salt tolerance in cultivated and wild tomato species through in vitro shoot apex culture. Plant Cell Tissue Organ Cult. 1998, 53, 19–26. [Google Scholar]
  37. Aebi, H. Catalase in vitro. Methods Enzymol. 1984, 105, 121–126. [Google Scholar]
  38. Thomas, R.L.; Jen, J.J.; Morr, C.V. Changes in soluble and bound peroxidase-IAA oxidase during tomato fruit development. J. Food Sci. 1982, 47, 158–161. [Google Scholar] [CrossRef]
  39. Nakano, Y.; Asada, K. Hydrogen peroxide is scavenged by ascorbate-specific peroxidase in spinach chloroplasts. Plant Cell Physiol. 1981, 22, 867–880. [Google Scholar]
  40. Foster, J.G.; Hess, J.L. Responses of Superoxide Dismutase and Glutathione Reductase Activities in Cotton Leaf Tissue Exposed to an Atmosphere Enriched in Oxygen. Plant Physiol. 1980, 66, 482–487. [Google Scholar] [CrossRef]
  41. Bradford, M.N. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 1976, 12, 248–254. [Google Scholar] [CrossRef]
  42. Yu, Q.; Rengel, Z. Waterlogging influences plant growth and activities of superoxide dismutases in narrow-leafed lupin and transgenic tobacco plants. J. Plant Physiol. 1999, 155, 431–438. [Google Scholar] [CrossRef]
  43. Sass, J.A. Botanical Microtechnique, 3rd ed.; The Iowa State University Press: Ames, IA, USA, 1961. [Google Scholar]
  44. Nassar, M.A.; El-Sahhar, K.F. Botanical Preparations and Microscopy (Microtechnique); Academic Bookshop: Giza, Egypt, 1998; p. 219. (In Arabic) [Google Scholar]
  45. Kaydan, D.; Okut, M.Y. Effects of salicylic acid on the growth and some physiological characters in salt stressed wheat (Triticum aestivum L.). Tarim Bİlimleri Dergisi. 2007, 13, 114–119. [Google Scholar]
  46. Abdul Qados, A.M.S. Effects of salicylic acid on growth, yield and chemical contents of pepper (Capsicum annuum L.) plants grown under salt stress conditions. Int. J. Agric. Crop Sci. 2015, 8, 107–113. [Google Scholar]
  47. Tsai, Y.C.; Hong, C.Y.; Liu, L.F.; Kao, C.H. Relative importance of Na+ and Cl in NaCl-induced antioxidant systems in roots of rice seedlings. Physiol. Plant. 2004, 122, 86–94. [Google Scholar] [CrossRef]
  48. Cheeseman, J. Hydrogen peroxide and plant stress: A challenging relationship. Plant Stress 2007, 1, 4–15. [Google Scholar]
  49. Zaki, S.S.; Rady, M.M. Moringa oleifera leaf extract improves growth, physio-chemical attributes, antioxidant defence system and yields of salt-stressed Phaseolus vulgaris L. plants. Int. J. ChemTech Res. 2015, 8, 120–134. [Google Scholar]
  50. Desoky, E.M.; Elrys, A.S.; Rady, M.M. Integrative moringa and licorice extracts application improves performance and reduces fruit contamination content of pepper plants grown on heavy metals-contaminated saline soil. Ecotoxicol. Environ. Saf. 2019, 169, 50–60. [Google Scholar] [CrossRef] [PubMed]
  51. Halliwell, B.; Gutteridge, M.J.C. Free Radicals in Biology and Medicine; Oxford University Press: Oxford, UK, 2015. [Google Scholar]
  52. Schutzendubel, A.; Polle, A. Plant responses to abiotic stresses: Heavy metal induced oxidative stress and protection by mycorrhization. J. Exp. Bot. 2002, 53, 1351–1365. [Google Scholar] [CrossRef] [PubMed]
  53. Semida, W.M.; Rady, M.M. Presoaking application of propolis and maize grain extracts alleviates salinity stress in common bean (Phaseolus vulgaris L.). Sci. Hortic. 2014, 68, 210–217. [Google Scholar] [CrossRef]
  54. Desoky, E.-S.M.; Mansour, E.; Ali, M.M.A.; Yasin, M.A.T.; Abdul-Hamid, M.I.E.; Rady, M.M.; Ali, E.F. Exogenously Used 24-Epibrassinolide Promotes Drought Tolerance in Maize Hybrids by Improving Plant and Water Productivity in an Arid Environment. Plants 2021, 10, 354. [Google Scholar] [CrossRef]
  55. Semida, W.M.; Abdelkhalik, A.; Mohamed, G.F.; Abd El-Mageed, T.A.; Abd El-Mageed, S.A.; Rady, M.M.; Ali, E.F. Foliar Application of Zinc Oxide Nanoparticles Promotes Drought Stress Tolerance in Eggplant (Solanum melongena L.). Plants 2021, 10, 421. [Google Scholar] [CrossRef]
  56. Alharby, H.F.; Al-Zahrani, H.S.; Hakeem, K.R.; Alsamadany, H.; Desoky, E.-S.M.; Rady, M.M. Silymarin-Enriched Biostimulant Foliar Application Minimizes the Toxicity of Cadmium in Maize by Suppressing Oxidative Stress and Elevating Antioxidant Gene Expression. Biomolecules 2021, 11, 465. [Google Scholar] [CrossRef]
  57. Alharby, H.F.; Al-Zahrani, H.S.; Alzahrani, Y.M.; Alsamadany, H.; Hakeem, K.R.; Rady, M.M. Maize Grain Extract Enriched with Polyamines Alleviates Drought Stress in Triticum aestivum through Up-Regulation of the Ascorbate-Glutathione Cycle, Glyoxalase System, and Polyamine Gene Expression. Agronomy 2021, 11, 949. [Google Scholar] [CrossRef]
  58. Rady, M.M.; Boriek, S.H.K.; Abd El-Mageed, T.A.; Seif El-Yazal, M.A.; Ali, E.F.; Hassan, F.A.S.; Abdelkhalik, A. Exogenous Gibberellic Acid or Dilute Bee Honey Boosts Drought Stress Tolerance in Vicia faba by Rebalancing Osmoprotectants, Antioxidants, Nutrients, and Phytohormones. Plants 2021, 10, 748. [Google Scholar] [CrossRef]
  59. Mekdad, A.A.A.; El-Enin, M.M.A.; Rady, M.M.; Hassan, F.A.S.; Ali, E.F.; Shaaban, A. Impact of Level of Nitrogen Fertilization and Critical Period for Weed Control in Peanut (Arachis hypogaea L.). Agronomy 2021, 11, 909. [Google Scholar] [CrossRef]
  60. Azzam, C.R.; Al-Taweel, S.K.; Abdel-Aziz, R.M.; Rabea, K.M.; Abou-Sreea, A.I.B.; Rady, M.M.; Ali, E.F. Salinity Effects on Gene Expression, Morphological, and Physio-Biochemical Responses of Stevia rebaudiana Bertoni In Vitro. Plants 2021, 10, 820. [Google Scholar] [CrossRef]
  61. Mekdad, A.A.A.; Rady, M.M.; Ali, E.F.; Hassan, F.A.S. Early Sowing Combined with Adequate Potassium and Sulfur Fertilization: Promoting Beta vulgaris (L.) Yield, Yield Quality, and K- and S-Use Efficiency in a Dry Saline Environment. Agronomy 2021, 11, 806. [Google Scholar] [CrossRef]
  62. Mekdad, A.A.A.; Shaaban, A.; Rady, M.M.; Ali, E.F.; Hassan, F.A.S. Integrated Application of K and Zn as an Avenue to Promote Sugar Beet Yield, Industrial Sugar Quality, and K-Use Efficiency in a Salty Semi-Arid Agro-Ecosystem. Agronomy 2021, 11, 780. [Google Scholar] [CrossRef]
  63. Djanaguiraman, M.; Devi, D.D.; Shanker, A.K.; Sheeba, J.A.; Bangarusamy, U. Selenium—An antioxidative protectant in soybean during senescence. Plant Soil 2005, 272, 77–86. [Google Scholar] [CrossRef]
  64. Iqbal, M.; Hussain, I.; Liaqat, H.; Ashraf, M.A.; Rasheed, R.; Rehman, A.U. Exogenously applied selenium reduces oxidative stress and induces heat tolerance in spring wheat. Plant Physiol. Biochem. 2015, 94, 95–103. [Google Scholar] [CrossRef]
  65. Jiang, C.; Zu, C.; Lu, D.; Zheng, Q.; Shen, J.; Wang, H.; Li, D. Effect of exogenous selenium supply on photosynthesis, Na+ accumulation and antioxidative capacity of maize (Zea mays L.) under salinity stress. Sci. Rep. 2017, 7, 42039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Kahakachchi, C.; Boakye, H.T.; Uden, P.C.; Tyson, J.F. Chromatographic speciation of anionic and neutral selenium compounds in Se-accumulating Brassica juncea (Indian mustard) and in selenized yeast. J. Chromatogr. A. 2004, 1054, 303–312. [Google Scholar] [CrossRef]
  67. Jia, H.; Song, Z.; Wu, F.; Ma, M.; Li, Y.; Han, D.; Yang, Y.; Zhang, S.; Cui, H. Low selenium increases the auxin concentration and enhances tolerance to low phosphorous stress in tobacco. Environ. Exp. Bot. 2018, 153, 127–134. [Google Scholar] [CrossRef]
  68. Haghighi, M.; Sheibanirad, A.; Pessarakli, M. Effects of selenium as a beneficial element on growth and photosynthetic attributes of greenhouse cucumber. J. Plant Nutr. 2016, 39, 1493–1498. [Google Scholar] [CrossRef] [Green Version]
  69. Rezende, R.A.L.S.; Rodrigues, F.A.; Soares, J.D.R.; Silveira, H.R.D.O.; Pasqual, M.; Dias, G.D.M.G. Salt stress and exogenous silicon influence physiological and anatomical features of in vitro-grown cape gooseberry. Cienc. Rural. 2018, 48, e20170176. [Google Scholar] [CrossRef] [Green Version]
  70. Zahedi, S.M.; Abdelrahman, M.; Hosseini, M.S.; Hoveizeh, N.F.; Tran, L.P. Alleviation of the effect of salinity on growth and yield of strawberry by foliar spray of selenium-nanoparticles. Environ. Pollut. 2019, 253, 246–258. [Google Scholar] [CrossRef]
  71. Djanaguiraman, M.; Prasad, P.V.V.; Seppanen, M. Selenium protects sorghum leaves from oxidative damage under high temperature stress by enhancing antioxidant defense system. Plant Physiol. Biochem. 2010, 48, 999–1007. [Google Scholar] [CrossRef] [PubMed]
  72. Diao, M.; Ma, L.; Wang, J.; Cui, J.; Fu, A.; Liu, H. Selenium promotes the growth and photosynthesis of tomato seedlings under salt stress by enhancing chloroplast antioxidant defense system. J. Plant Growth Regul. 2014, 33, 671–682. [Google Scholar] [CrossRef]
  73. Nawaz, F.; Ashraf, M.Y.; Ahmad, R.; Waraich, E.A.; Shabbir, R.N. Selenium (Se) regulates seedling growth in wheat under drought stress. Adv. Chem. 2014, 2014, 143567. [Google Scholar] [CrossRef] [Green Version]
  74. Hawrylak-Nowak, B. Effect of selenium on selected macronutrients in maize plants. J. Elem. 2008, 13, 513–519. [Google Scholar]
  75. Levitt, J. Responses of Plants to Environmental Stresses, Volume 2. Water, Radiation, Salt, and Other Stresses; Academic Press: New York, NY, USA, 1980; pp. 365–380. [Google Scholar]
  76. Pattanayak, G.K.; Tripathy, B.C. Overexpression of protochlorophyllide oxidoreductase C regulates oxidative stress in Arabidopsis. PLoS ONE 2011, 6, e26532. [Google Scholar] [CrossRef]
  77. Elkelish, A.A.; Soliman, M.H.; Alhaithloul, H.A.; El-Esawi, M.A. Selenium protects wheat seedlings against salt stress-mediated oxidative damage by up-regulating antioxidants and osmolytes metabolism. Plant Physiol. Biochem. 2019, 137, 144–153. [Google Scholar] [CrossRef]
  78. Ahanger, M.A.; Tomar, N.S.; Tittal, M.; Argal, S.; Agarwal, R.M. Plant growth under water/salt stress: ROS production; antioxidants and significance of added potassium under such conditions. Physiol. Mol. Biol. Plants 2017, 23, 731–744. [Google Scholar] [CrossRef]
  79. Khan, M.I.R.; Asgher, M.; Khan, N.A. Alleviation of salt-induced photosynthesis and growth inhibition by salicylic acid involves glycine betaine and ethylene in mung bean (Vigna radiata L.). Plant Physiol. Biochem. 2014, 80, 67–74. [Google Scholar] [CrossRef]
  80. Alyemeni, M.N.; Ahanger, M.A.; Wijaya, L.; Alam, P.; Bhardwaj, R.; Ahmad, P. Selenium mitigates cadmium-induced oxidative stress in tomato (Solanum lycopersicum L.) plants by modulating chlorophyll fluorescence, osmolyte accumulation, and antioxidant system. Protoplasma 2018, 255, 459–469. [Google Scholar] [CrossRef]
  81. Kong, L.; Wang, M.; Bi, D. Selenium modulates the activities of antioxidant enzymes, osmotic homeostasis and promotes the growth of sorrel seedlings under salt stress. Plant Growth Regul. 2005, 45, 155–163. [Google Scholar] [CrossRef]
  82. Nawaz, F.; Naeem, M.; Ashraf, M.Y.; Tahir, M.N.; Zulfiqar, B.; Salahuddin, M.; Shabbir, R.N.; Aslam, M. Selenium supplementation affects physiological and biochemical processes to improve fodder yield and quality of maize (Zea mays L.) under water deficit conditions. Front. Plant Sci. 2016, 7, 1438. [Google Scholar] [CrossRef] [Green Version]
  83. Saffaryazdi, A.; Lahouti, M.; Ganjeali, A.; Bayat, H. Impact of selenium supplementation on growth and selenium accumulation on spinach (Spinacia oleracea L.) plants. Not. Sci. Biol. 2012, 4, 95–102. [Google Scholar] [CrossRef] [Green Version]
  84. Rios, J.J.; Martínez-Ballesta, M.C.; Ruiz, J.M.; Blasco, B.; Carvajal, M. Silicon mediated improvement in plant salinity tolerance: The role of aquaporins. Front. Plant Sci. 2017, 8, 948. [Google Scholar] [CrossRef] [Green Version]
  85. Desoky, E.M.; ElSayed, A.I.; Merwad, A.M.A.; Rady, M.M. Stimulating antioxidant defenses, antioxidant gene expression, and salt tolerance in Pisum sativum seedling by pretreatment using licorice root extract (LRE) as an organic biostimulant. Plant Physiol. Biochem. 2019, 142, 292–302. [Google Scholar] [CrossRef]
  86. Stępień, P.; Kłobus, G. Water relations and photosynthesis in Cucumis sativus L. leaves under salt stress. Biol. Plant. 2006, 50, 610–616. [Google Scholar] [CrossRef]
  87. Trapp, S.; Feificova, D.; Rasmussen, N.F.; Bauer-Gottwein, P. Plant uptake of NaCl in relation to enzyme kinetics and toxic effects. Environ. Exp. Bot. 2008, 64, 1–7. [Google Scholar] [CrossRef]
  88. Ma, X.; Zhang, J.; Huang, B. Cytokinin-mitigation of salt-induced leaf senescence in perennial ryegrass involving the activation of antioxidant systems and ionic balance. Environ. Exp. Bot. 2016, 125, 1–11. [Google Scholar] [CrossRef]
  89. Ahanger, M.A.; Agarwal, R.M. Salinity stress induced alterations in antioxidant metabolism and nitrogen assimilation in wheat (Triticum aestivum L) as influenced by potassium supplementation. Plant Physiol. Biochem. 2017, 115, 449–460. [Google Scholar] [CrossRef]
  90. Munns, R.; Tester, M. Mechanisms of salinity tolerance. Annu. Rev. Plant Biol. 2008, 59, 651–681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Guo, C.Y.; Wang, X.Z.; Chen, L.; Ma, L.N.; Wang, R.Z. Physiological and biochemical responses to saline-alkaline stress in two halophytic grass species with different photosynthetic pathways. Photosynthetica 2015, 53, 128–135. [Google Scholar] [CrossRef]
  92. Zhu, J.K. Plant salt tolerance. Trends Plant Sci. 2001, 6, 66–71. [Google Scholar] [CrossRef]
  93. Howladar, S.M. A novel Moringa oleifera leaf extract can mitigate the stress effect of salinity and cadmium in bean (Phaseolus vulgaris L.) plants. Ecotoxicol. Environ. Saf. 2014, 100, 69–75. [Google Scholar] [CrossRef]
  94. Farhangi-Abriz, S.; Torabian, S. Antioxidant enzyme and osmotic adjustment changes in bean seedlings as affected by biochar under salt stress. Ecotoxicol. Environ. Saf. 2017, 137, 64–70. [Google Scholar] [CrossRef] [PubMed]
  95. Chandrasekhar, K.R.; Sandhyarani, S. Salinity induced chemical changes in Crotalaria striata DC plants. Indian J. Plant Physiol. 1996, 1, 44–48. [Google Scholar]
  96. Sairam, R.K.; Rao, K.V.; Srivastava, G.C. Differential response of wheat genotypes to long term salinity stress in relation to oxidative stress, antioxidant activity and osmolyte concentration. Plant Sci. 2002, 163, 1037–1046. [Google Scholar] [CrossRef]
  97. Hernández, J.A.; Almansa, M.S. Short-term effects of salt stress on antioxidant systems and leaf water relations of pea leaves. Physiol. Plant. 2002, 115, 251–257. [Google Scholar] [CrossRef]
  98. Allen, D.J.; Kee, I.F.M.; Farage, P.K.; Baker, N.R. Analysis of the limitation to CO2 assimilation to exposure of leaves of two Brassica napus cultivars to UV-B. Plant Cell Environ. 1997, 20, 633–640. [Google Scholar] [CrossRef]
  99. Jiang, Y.; Huang, N. Drought and Heat stress injury to two cool season turf grasses in relation to antioxidant metabolism and lipid peroxidation. Crop Sci. 2001, 41, 436–442. [Google Scholar] [CrossRef]
  100. Alireza, Y.; Aboueshaghi, R.S.; Dehnavi, M.M.; Balouchi, H. Effect of micronutrients foliar application on grain qualitative characteristics and some physiological traits of bean (Phaseolus vulgaris L.) under drought stress. Ind. J. Fund. Appl. Life Sci. 2014, 4, 124–131. [Google Scholar]
  101. Hernández, J.A.; Francisco, F.J.; Corpas, G.M.; Gómez, L.A.; Del Río, F.S. Salt induced oxidative stresses mediated by activated oxygen species in pea leaves mitochondria. Plant Physiol. 1993, 89, 103–110. [Google Scholar] [CrossRef]
  102. Alscher, R.G.; Erturk, N.; Heath, L.S. Role of superoxide dismutases (SODs) in controlling oxidative stress in plants. J. Exp. Biol. 2002, 53, 1331–1341. [Google Scholar]
  103. Feierabend, J. Catalases in plants: Molecular and functional properties and role in stress defense. In Antioxidants and Reactives Oxygen Species in Plants; Smirnoff, N., Ed.; Blackwell Publishing: London, UK, 2005; pp. 101–140. [Google Scholar]
  104. Foyer, C.H. Free radical processes in plants. Biochem. Soc. Transact. 1996, 24, 427–434. [Google Scholar] [CrossRef] [Green Version]
  105. Foyer, C.H.; Noctor, G. Redox Regulation in Photosynthetic Organisms: Signaling: Acclimation and Practical Implications. Trends Plant Sci. 2009, 6, 486–492. [Google Scholar] [CrossRef]
  106. Zhang, Y.; Wang, L.; Liu, Y.; Zhang, Q.; Wei, Q.; Zhang, W. Nitric oxide enhances salt tolerance in maize seedlings through increasing activities of proton-pump and Na+/H+ antiport in the tonoplast. Planta 2006, 224, 545–555. [Google Scholar] [CrossRef] [PubMed]
  107. Ahmad, P.; Sarwat, M.; Bhat, N.A.; Wani, M.R.; Kazi, A.G.; Tran, L.S. Alleviation of cadmium toxicity in Brassica juncea L. (Czern. & Coss.) by calcium application involves various physiological and biochemical strategies. PLoS ONE 2015, 10, e0114571. [Google Scholar]
  108. Drahonovsky, J.; Szákova, J.; Mestek, O.; Tremlova, J.; Kanab, A.; Najmanova, J.; Tlustoa, P. Selenium uptake, transformation and inter-element interactions by selected wildlife plant species after foliar selenate application. Environ. Exp. Bot. 2016, 125, 12–19. [Google Scholar] [CrossRef]
  109. Astaneh, R.K.; Bolandnazar, S.; Nahandi, F.Z.; Oustan, F. The effects of selenium on some physiological traits and K, Na concentration of garlic (Allium sativum L.) under NaCl stress. Inf. Process Agric. 2018, 5, 156–161. [Google Scholar] [CrossRef]
  110. Shabala, S.; Pottosin, I. Regulation of potassium transport in plants under hostile conditions: Implications for abiotic and biotic stress tolerance. Physiol. Plant. 2014, 151, 257–279. [Google Scholar] [CrossRef]
  111. Gupta, M.; Gupta, S. An overview of selenium uptake, metabolism, and toxicity in plants. Front. Plant Sci. 2016, 7, 2074. [Google Scholar]
  112. Jiang, C.; Zheng, Q.; Liu, Z.; Xu, W.; Liu, L.; Zhao, G.; Long, X. Overexpression of Arabidopsis thaliana Na+/H+ antiporter gene enhanced salt resistance in transgenic poplar (Populus × euramericana “Neva”). Trees 2012, 26, 685–694. [Google Scholar] [CrossRef]
  113. Almeida, D.M.; Gregorio, G.B.; Oliveira, M.M.; Saibo, N.J.M. Five novel transcription factors as potential regulators of OsNHX1 gene expression in a salt tolerant rice genotype. Plant Mol. Biol. 2017, 93, 61–77. [Google Scholar] [CrossRef] [PubMed]
  114. Feng, R.; Wei, C.; Tu, S. The roles of selenium in protecting plants against abiotic stresses. Environ. Exp. Bot. 2013, 87, 58–68. [Google Scholar] [CrossRef]
  115. Rady, M.O.A.; Semida, W.M.; Howladar, S.M.; Abd El-Mageed, T.A. Raised beds modulate physiological responses, yield and water use efficiency of wheat (Triticum aestivum L) under deficit irrigation. Agric. Water Manag. 2021, 245, 106629. [Google Scholar] [CrossRef]
  116. Khattab, H.I.; Emam, M.A.; Emam, M.M.; Helal, N.M.; Mohamed, M.R. Effect of selenium and silicon on transcription factors NAC5 and DREB2A involved in drought-responsive gene expression in rice. Biol. Plant. 2014, 58, 265–273. [Google Scholar] [CrossRef]
  117. Al-Elwany, O.A.; Mohamed, G.F.; Abdehrahman, H.A.; Rady, M.M.; Abdel Latef, A.H. Exogenous glutathi-one-mediated tolerance to deficit irrigation stress in salt-affected Capsicum frutescence (L.) plants is connected with higher antioxidant content and proper ion homeostasis. Not. Bot. Horti Agrobot. Cluj Napoca 2020, 48, 1957–1979. [Google Scholar] [CrossRef]
  118. Hussein, H.A.A.; Darwesh, O.M.; Mekki, B.B.; El-Hallouty, S.M. Evaluation of cytotoxicity, biochemical profile and yield components of groundnut plants treated with nano-selenium. Biotechnol. Rep. 2019, 24, e00377. [Google Scholar] [CrossRef]
Figure 1. Oxidative stress markers and their outcomes responses of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons. Above bars, different letters are considered significantly different at p ≤ 0.05. Con, control (spraying leaves with distilled water); F1, spraying leaves with 0.5 mM Se-NPs; F2, spraying leaves with 1 mM Se-NPs; F3, spraying leaves with 1.5 mM Se-NPs; MDA, lipid peroxidation measured as malondialdehyde; H2O2, hydrogen peroxide; O2•−, superoxide radical.
Figure 1. Oxidative stress markers and their outcomes responses of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons. Above bars, different letters are considered significantly different at p ≤ 0.05. Con, control (spraying leaves with distilled water); F1, spraying leaves with 0.5 mM Se-NPs; F2, spraying leaves with 1 mM Se-NPs; F3, spraying leaves with 1.5 mM Se-NPs; MDA, lipid peroxidation measured as malondialdehyde; H2O2, hydrogen peroxide; O2•−, superoxide radical.
Plants 10 01189 g001
Figure 2. Enzyme activity response of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons. Above bars, different letters are considered significantly different at p ≤ 0.05. Con, control (spraying leaves with distilled water); F1, spraying leaves with 0.5 mM Se-NPs; F2, spraying leaves with 1 mM Se-NPs; F3, spraying leaves with 1.5 mM Se-NPs; CAT, catalase; APX, ascorbate peroxidase; GR, glutathione reductase; SOD, superoxide dismutase; POX, peroxidase.
Figure 2. Enzyme activity response of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons. Above bars, different letters are considered significantly different at p ≤ 0.05. Con, control (spraying leaves with distilled water); F1, spraying leaves with 0.5 mM Se-NPs; F2, spraying leaves with 1 mM Se-NPs; F3, spraying leaves with 1.5 mM Se-NPs; CAT, catalase; APX, ascorbate peroxidase; GR, glutathione reductase; SOD, superoxide dismutase; POX, peroxidase.
Plants 10 01189 g002
Figure 3. Nutrient content response of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons. Above bars, different letters are considered significantly different at p ≤ 0.05. Con, control (spraying leaves with distilled water); F1, spraying leaves with 0.5 mM Se-NPs; F2, spraying leaves with 1 mM Se-NPs; F3, spraying leaves with 1.5 mM Se-NPs; N, nitrogen; P, phosphorus; K+, potassium ion; Na+, sodium ion.
Figure 3. Nutrient content response of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons. Above bars, different letters are considered significantly different at p ≤ 0.05. Con, control (spraying leaves with distilled water); F1, spraying leaves with 0.5 mM Se-NPs; F2, spraying leaves with 1 mM Se-NPs; F3, spraying leaves with 1.5 mM Se-NPs; N, nitrogen; P, phosphorus; K+, potassium ion; Na+, sodium ion.
Plants 10 01189 g003aPlants 10 01189 g003b
Figure 4. Leaf anatomy (transverse section through the leaflet blade) responses of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in 2020 (second) season: (A) control (spraying leaves with distilled water); (B) spraying leaves with 0.5 mM Se-NPs; (C) spraying leaves with 1 mM Se-NPs; and (D) spraying leaves with 1.5 mM Se-NPs. UE, upper epidermis; LE, lower epidermis; PT, palisade tissue; ST, spongy tissue; XT, xylem tissue; PhT, phloem tissue. X = 200 µm.
Figure 4. Leaf anatomy (transverse section through the leaflet blade) responses of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in 2020 (second) season: (A) control (spraying leaves with distilled water); (B) spraying leaves with 0.5 mM Se-NPs; (C) spraying leaves with 1 mM Se-NPs; and (D) spraying leaves with 1.5 mM Se-NPs. UE, upper epidermis; LE, lower epidermis; PT, palisade tissue; ST, spongy tissue; XT, xylem tissue; PhT, phloem tissue. X = 200 µm.
Plants 10 01189 g004
Table 1. Some major properties of the investigated soil.
Table 1. Some major properties of the investigated soil.
Soil CharacteristicsUnitValues
First SeasonSecond Season
Sand%44.244.7
Silt31.730.8
Clay24.124.5
Texture classLoam
Field capacity%16.216.3
CaCO3g kg−160.162.3
Organic matter7.558.25
pH (in soil paste) 7.527.61
EC (in soil paste extract)dS m−17.557.61
Soluble ions (anions and cations) **
Mg2+mmolc L−119.920.1
Ca2+15.615.7
SO42–8.198.4
K+6.726.53
HCO320.920.7
CO32–--
Na+19.119.5
Cl31.732.7
Available nutrient
Nmg kg−1 soil59.059.4
P9.809.90
K98.799.5
Se0.060.08
** Mg2+ = magnesium cation, Ca2+ = calcium cation, SO42 = sulfate anion, K+ = potassium cation, HCO3 = bicarbonate anion, CO32 = carbonate anion, Na+= sodium cation, and Cl = chloride anion.
Table 2. Some major properties of the Se-NPs used in this study.
Table 2. Some major properties of the Se-NPs used in this study.
The PropertyThe UnitThe Value
Diameternm10–45
Surface aream2 g−130–50
Densityg cm−33.89
Purity%99.95
MorphologySpherical
Table 3. Growth and productivity responses of salt-stressed common bean plants (soil EC = 7.55−7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons.
Table 3. Growth and productivity responses of salt-stressed common bean plants (soil EC = 7.55−7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons.
Foliar SprayShL (cm)NoL-PLeA-P (cm2)DW-Sh (g)NoP-PGPY-H (ton ha−1)
The first season (2019)
Distilled water16.1 ± 1.1 d6.87 ± 0.4 b7.26 ± 0.3 c3.61 ± 0.1 d10.6 ± 0.9 d2.17 ± 0.19 c
Se-NPs (0.5 mM)21.5 ± 1.4 c8.24 ± 0.6 b7.55 ± 0.4 c6.17 ± 0.3 c12.6 ± 1.1 c4.33 ± 0.21 bc
Se-NPs (1 mM)32.3 ± 2.1 a10.7 ± 0.8 a9.25 ± 0.6 a9.31 ± 0.6 a18.8 ± 1.5 a7.88 ± 0.26 a
Se-NPs (1.5 mM)26.1 ± 1.6 b9.82 ± 0.4 a8.65 ± 0.5 b7.64 ± 0.4 b14.9 ± 1.3 b6.74 ± 0.19 ab
The second season (2020)
Distilled water16.9 ± 1.3 d7.77 ± 0.3 c7.76 ± 0.2 d4.11 ± 0.2 c11.1 ± 0.8 d2.21 ± 0.12 c
Se-NPs (0.5 mM)22.4 ± 1.5 c9.21 ± 0.5 b8.29 ± 0.4 c6.39 ± 0.3 b13.7 ± 1.2 c4.40 ± 0.14 bc
Se-NPs (1 mM)33.6 ± 2.5 a12.0 ± 0.7 a10.0 ± 0.8 a9.39 ± 0.6 a19.9 ± 1.2 a7.95 ± 0.31 a
Se-NPs (1.5 mM)27.2 ± 2.3 b10.9 ± 0.6 a9.35 ± 0.6 b7.71 ± 0.5 b15.8 ± 1.3 b6.81 ± 0.21 ab
Data are means (n = 9) ± SE. In each column, means with different letters are considered significantly different at p ≤ 0.05. ShL, shoot length; NoL-P, plant leaf number; LeA-P, plant leaf area; DW-Sh, shoot dry weight; NoP-P, plant pods number; GPY-H, hectare green pods yield.
Table 4. Photosynthetic efficiency responses of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons.
Table 4. Photosynthetic efficiency responses of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons.
Foliar SprayChlorophyll “a”Chlorophyll “b”CarotenoidsPn (µmol CO2 m−2 s−1)Tr (mmol H2 O m−2 s−1)
(mg g−1 FW)
The first season (2019)
Distilled water0.94 ± 0.07 d0.53 ± 0.02 c0.85 ± 0.03 c5.56 ± 0.2 d3.18 ± 0.1 d
Se-NPs (0.5 mM)1.11 ± 0.08 c0.62 ± 0.03 b0.98 ± 0.04 bc9.77 ± 0.4 c4.76 ± 0.2 c
Se-NPs (1 mM)1.33 ± .0.06 a0.71 ± 0.04 a1.20 ± 0.06 a12.9 ± 0.5 a7.01 ± 0.5 a
Se-NPs (1.5 mM)1.20 ± 0.08 b0.65 ± 0.02 ab1.07 ± 0.07 ab11.2 ± 0.6 b5.97 ± 0.3 b
The second season (2020)
Distilled water0.97 ± 0.04 c0.55 ± 0.01 b0.90 ± 0.02 b5.86 ± 0.3 d3.44 ± 0.1 d
Se-NPs (0.5 mM)1.14 ± 0.06 b0.65 ± 0.02 a1.01 ± 0.04 b9.88 ± 0.6 c4.94 ± 0.1 c
Se-NPs (1 mM)1.36 ± 0.07 a0.74 ± 0.02 a1.24 ± 0.04 a13.0 ± 0.8 a7.26 ± 0.4 a
Se-NPs (1.5 mM)1.23 ± 0.05 b0.68 ± 0.03 a1.10 ± 0.03 ab11.5 ± 0.7 b6.18 ± 0.3 b
Data are means (n = 9) ± SE. In each column, means with different letters are considered significantly different at p ≤ 0.05. Pn, rate of net photosynthetic; Tr, rate of transpiration.
Table 5. Tissue cell integrity response of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons.
Table 5. Tissue cell integrity response of salt-stressed common bean plants (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in two consecutive seasons.
Foliar SprayRWC (%)MSI (%)Free proline (µmol g1 DW)Soluble sugars (mg g1 DW)Se content (mg kg−1 DW)
The first season (2019)
Distilled water77.1 ± 2.2 c36.9 ± 1.2 d27.5 ± 1.3 d18.6 ± 1.1 d10.4 ± 0.2 d
Se-NPs (0.5 mM)79.6 ± 2.6 c48.9 ± 1.5 c29.5 ± 1.2 c21.3 ± 1.3 c21.6 ± 0.4 c
Se-NPs (1 mM)90.6 ± 3.5 a81.4 ± 2.5 a34.1 ± 1.3 a26.2 ± 1.4 a28.2 ± 0.5 b
Se-NPs (1.5 mM)86.3 ± 3.3 b61.0 ± 1.8 b32.4 ± 1.5 b24.1 ± 1.5 b36.4 ± 0.7 a
The second season (2020)
Distilled water79.1 ± 2.5 b38.2 ± 1.1 d28.1 ± 1.4 c19.1 ± 1.2 d12.0 ± 0.2 d
Se-NPs (0.5 mM)80.9 ± 3.2 b50.6 ± 1.9 c30.1 ± 1.5 b21.8 ± 1.3 c24.2 ± 0.3 c
Se-NPs (1 mM)92.5 ± 3.4 a83.5 ± 2.9 a34.5 ± 1.4 a26.6 ± 1.4 a29.4 ± 0.6 b
Se-NPs (1.5 mM)88.1 ± 3.5 a63.1 ± 2.3 b33.1 ± 1.6 a24.5 ± 1.5 b37.0 ± 0.8 a
Data are means (n = 9) ± SE. In each column, means with different letters are considered significantly different at p ≤ 0.05. RWC, relative water content; MSI, membrane stability index; Se, selenium.
Table 6. Anatomical features responses of salt-stressed common bean leaf (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in 2020 (second) season.
Table 6. Anatomical features responses of salt-stressed common bean leaf (soil EC = 7.55–7.61 dS m−1) to foliar nourishment with selenium dioxide nanoparticles (Se-NPs) in 2020 (second) season.
Foliar SprayBlade Thic. (μm)Palisade Thick. (μm)Spongy Thick. (μm)Length of Midvein (μm)Width of Midvein (μm)Phloem Thick. (μm)Xylem Thick. (μm)Diameter of Vessel (μm)
Distilled water130.6 ± 1.1d40.5 ± 0.9d54.3 ± 1.3d750.4 ± 2.6d619.1 ± 4.1d101.6 ± 1.1d105.4 ± 1.2d15.5 ± 0.4d
Se-NPs (0.5 mM)165.5 ± 1.5c49.2 ± 1.2c60.6 ± 1.2c904.5 ± 3.5c722.2 ± 2.5c117.3 ± 1.4c113.32 ± 1.4c20.3 ± 0.5c
Se-NPs (1 mM)250.8 ± 2.2a99.5 ± 1.3a114.8 ± 2.2a1389.9 ± 2.9a1283.6 ± 4.3a211.8 ± 1.5a211.2 ± 1.9a33.9 ± 0.8a
Se-NPs (1.5 mM)230.3 ± 1.9b85.4 ± 1.1b100.6 ± 2.5b1270.6 ± 3.5b1172.7 ± 3.6b181.2 ± 1.6b145.6 ± 1.7b29.2 ± 0.7b
Data are means (n = 9) ± SE. In each column, means with different letters are considered significantly different at p ≤ 0.05. Thick., thickness.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rady, M.M.; Desoky, E.-S.M.; Ahmed, S.M.; Majrashi, A.; Ali, E.F.; Arnaout, S.M.A.I.; Selem, E. Foliar Nourishment with Nano-Selenium Dioxide Promotes Physiology, Biochemistry, Antioxidant Defenses, and Salt Tolerance in Phaseolus vulgaris. Plants 2021, 10, 1189. https://doi.org/10.3390/plants10061189

AMA Style

Rady MM, Desoky E-SM, Ahmed SM, Majrashi A, Ali EF, Arnaout SMAI, Selem E. Foliar Nourishment with Nano-Selenium Dioxide Promotes Physiology, Biochemistry, Antioxidant Defenses, and Salt Tolerance in Phaseolus vulgaris. Plants. 2021; 10(6):1189. https://doi.org/10.3390/plants10061189

Chicago/Turabian Style

Rady, Mostafa M., El-Sayed M. Desoky, Safia M. Ahmed, Ali Majrashi, Esmat F. Ali, Safaa M. A. I. Arnaout, and Eman Selem. 2021. "Foliar Nourishment with Nano-Selenium Dioxide Promotes Physiology, Biochemistry, Antioxidant Defenses, and Salt Tolerance in Phaseolus vulgaris" Plants 10, no. 6: 1189. https://doi.org/10.3390/plants10061189

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