Next Article in Journal
Detection of Fundamental Quality Traits of Winter Jujube Based on Computer Vision and Deep Learning
Next Article in Special Issue
The Physiological Mechanism of Melatonin Enhancing the Tolerance of Oat Seedlings under Saline–Alkali Stress
Previous Article in Journal
Effects of Combined Application of Biological Agent and Fertilizer on Fungal Community Structure in Rhizosphere Soil of Panax notoginseng
Previous Article in Special Issue
Physiological and Biochemical Mechanisms of Exogenous Melatonin Regulation of Saline–Alkali Tolerance in Oats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Seed Treatment Modulates Phenolic and Fatty Acid Metabolism of Sunflower under Water Scarcity

1
Department of Botany, Government College University, Faisalabad 38000, Pakistan
2
Institute of Molecular Biology and Biotechnology, The University of Lahore, Lahore 54000, Pakistan
3
Department of Botany, Bahrain University, Sakhir 32038, Bahrain
4
Department of Botany, Division of Science and Technology, University of Education Lahore, Lahore 54770, Pakistan
5
Department of Botany, University of Agriculture, Faisalabad 38000, Pakistan
6
Botany and Microbiology Department, Faculty of Science, King Saud University, Riyadh 11451, Saudi Arabia
7
Instituto de Conservación y Mejora de la Agrodiversidad Valenciana, Universitat Politècnica de València, 46022 Valencia, Spain
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Agronomy 2023, 13(8), 2094; https://doi.org/10.3390/agronomy13082094
Submission received: 17 June 2023 / Revised: 8 July 2023 / Accepted: 10 July 2023 / Published: 10 August 2023
(This article belongs to the Special Issue Strategies for Enhancing Abiotic Stress Tolerance in Crops)

Abstract

:
Environmental and anthropogenic activities are pushing the earth towards warmer years, which is reducing agricultural land and causing water scarcity. It is well documented that sunflower (being drought tolerant) crops can be grown under water deficit conditions with some additional supportive priming applications to compensate for drought-induced challenges. However, finding the most efficient and eco-friendly priming tools is always a top priority among researchers to improve plant growth, adaptive traits, and productivity. In this study, an experiment was performed on oil-producing crops (sunflower) using seed magnetic treatment. The seeds were subjected to 0.1, 0.2, and 0.3 Tesla for 20 min, respectively. Plants were exposed to water limitation (100 and 50%) after 30 days of germination. Sunflower showed its tolerance to water limitation by maintaining the majority of growth parameters, nutritive value, metabolizable energy, and higher proline content. Nevertheless, a reduction in the achene number per capitulum, capitulum weight, chlorophyll, catalase activity, unsaturated to saturated fatty acids, anthocyanin, and hydroxyl derivatives of cinnamic acid supported our hypothesis about the need for some supportive techniques. Better metabolic adjustment and percentage of oil yield were manifested by 0.3 T magnetic seed treatment, which was used for phenolic and fatty acid profiling. To conclude, magnetic treatment of seeds may improve their primary metabolic capacity and antioxidation potential, which in turn may activate their secondary metabolism as evidenced by an excess of gallic acid, quercetin, benzoic acid, curcumin acid, sinapic acid, and chlorogenic acid.

1. Introduction

Sunflower is one of the important oil seed crops with high nutraceutical importance. Its leaves are traditionally used as herbal remedies due to the presence of nutraceutical secondary metabolites, including alkaloid, tannin, steroid, and sesquiterpenoid compounds [1]. Sunflower oil used in the food industry can also be reused as CI engine fuel [2]. The husks of sunflower seeds after oil consumption are utilized for electricity production [1].
Sunflower has an additional benefit due to its moderate drought tolerance. This has made it possible to grow in areas with water deficit conditions. Water scarcity is a universal problem for the flora and fauna of the ecosystem. It is reducing the agricultural area day by day. In addition to environmental calamities, anthropogenic contributions are making the situation worse, which further poses a serious threat to sustainable agriculture [3,4,5,6]. The most practiced and simplest strategy to resolve this issue at the agronomic level is to grow drought-tolerant crops like sunflowers. Despite being tolerant, plants experience disturbed metabolism under stressful conditions. Photosynthesis, the efficiency of electron transport [7], the genomes of the hormonal system, the redox system, secondary metabolism, and hampered signaling are some of these phenomena [8].
A search for strategic applications is needed to alleviate drought-induced metabolic disturbances in sensitive and tolerant plants. In agriculture, the static magnetic field (SMF) is an environmentally friendly physical method of seed priming. It supports growth, metabolism, and development in normal [9], biotic, and abiotic [10] stressed environments. However, the efficacy of magnetic seed priming varies from species to species/variety to variety and magneto-reception (µT to 200 SMF) [9]. SMF activates cell magneto-receptor cryptochromes. It involves differences in electron spinning that vary with cell type, energy, and the dose of SMF [11]. Cryptochrome-1 can mediate stress tolerance [11,12].
Sunflowers with improved growth proved their potential to respond to applied SMF [13]. It is well studied at the agronomic and physiological levels. The proteomic, transcriptomic, and metabolomic levels of studies of SMF-treated sunflowers are scarce in the literature. A series of experiments using these approaches for SMF-treated sunflowers in normal and stressed environments are needed. As SMF is dose- and species-specific, we aimed to optimize the best dose for sunflower seeds, followed by an assessment of its effect on metabolic manipulation in sunflowers under limited water conditions using a metabolomics approach.

2. Materials and Methods

2.1. Seed Source and Selection of Levels

The present work was performed in an experimental field at Govt College University, Faisalabad. Sunflower seeds (FH-129) were collected from Ayub Agricultural Research Institute. Seeds were subjected to magnetic treatment following Mahajan and Pandey’s [14] method, with minor changes. With the help of magnetic field generators, a variable horizontal magnetic field was applied at 0.0, 0.1, 0.2, and 0.3 T for 20 min. A steady DC supply source (0–45 v/0–7.5 A) was supplied to the electromagnet of the magnetic field. The distance between magnetic poles was maintained at 5 mm. Magnetic strength was checked thoroughly using a gauss meter. Three sets of samples (each comprising 25 seeds) were organized in plastic bags to be exposed to the magnetic field between the pole pieces of the electromagnet.

2.2. Experimental Plan

Seeds were planted in two equal-sized plots: one plot was irrigated with regular irrigation, and the other was irrigated with 50% irrigation. Suggested doses (150-60-60 NPK kg ha−1) of fertilizer were applied. At sowing time, 1/3 of nitrogen was added to the soil as urea. The leftover dose was divided into two doses: one was applied at the vegetative stage and one was applied later at the flowering stage. Phosphorus (P) and potassium (K) were applied as triple super phosphate and sulphate of potash (K2SO4) during seedbed preparation. All regular practices such as irrigation, weed management, hoeing, plant protection, etc., were kept normal for the crop during the experiment. The growth and biochemical aspects were evaluated during the flowering stages. The final harvest was taken when the backs of the heads turned yellow and the bracts were brown and dried for 4 to 5 days. Five capitula were selected for the determination of different yield components. After measuring the capitulum weight, achenes were separated. The number of achenes per capitulum was counted and the weight of 100 achenes was measured. The individual parameters were quantified by averaging five readings.

2.3. Estimation of Photosynthetic Pigments

Leaf total chlorophyll and carotenoid contents were estimated following the protocol of Davies and Goodwin [15]. The absorbance of the leaves extracted with 80% acetone was recorded at 645, 663, and 480 nm.

2.4. Determination of Total Phenolics, Flavonoids, and Total Protein Contents

Leaf total phenolics were determined following Julkunen-Tiitto [16] using FolinCiocalteu’s reagent. Sample extraction was prepared as described in the procedure. After 90 min of incubation, the absorbance was measured at 765 nm by spectrophotometer. Total flavonoid contents were determined following the detailed method of Marinova et al. [17], whereas total protein contents were estimated using the method explained by Bradford [18]. The reaction mixture was prepared and the absorbance of the samples was determined using a spectrophotometer at 595 nm.

2.5. Estimation of Reducing Sugar Contents and Total Soluble Sugars

The conten of reducing sugars was determined as described by Miller [19]. The reaction mixture was vortexed and absorbance was measured at 540 nm. The total soluble sugars were estimated using the method of Yoshida et al. [20]. After the vortex, the reaction mixture was heated at 95 °C for 15 min. The absorbance of the samples was measured at 625 nm.

2.6. Determination of Total Anthocyanin Content

The quantitative analysis of anthocyanin was performed spectrophotometrically using the method of Nozzolillo [21]. Acidified methanolic leaf extract of 3 mL was taken, and the absorbance was recorded by using a spectrophotometer at 536 and 600 nm.

2.7. Determination of % Oil Content

The percent oil content was determined by following Horwitz [22]. In a 50 mL capacity plastic test tube, three grams of the plant sample was used and 30 mL n-hexane was added to it. At 100 rpm, the shaker was adjusted and the tubes were left for over 24 h after the samples were centrifuged and the supernatant was collected. Two successive repetitions of extractions were performed on the residue via vortexing and centrifugation. Clear extracts from all three collections were preserved for oil estimation. The residues were dried as fat-free samples with a difference in weight after extraction was recorded as the fat content.

2.8. Estimation of Nutritional Value and Metabolizable Energy

This value was determined using the method of Indrayan et al. [23] as
N u t r i t i v e v a l u e = ( 4 × % a g e o f p r o t e i n ) + ( 9 × % a g e o f f a t s ) + ( 4 × % a g e o f c a r b o h y d r a t e s )
The metabolizable energy of samples was determined using the following expressions by Macrae et al. [24]
Energy (kJ) = (4 kcal/g × gprotein × 4.186) + (4 kcal/g × gfat × 4.186) + (9 kcal/g × gcarbohydrate × 4.186)
E n e r g y k J = 4   k c a l g × g p r o t e i n × 4.186 + ( 4   k c a l / g × g f a t × 4.186 ) + ( 9   k c a l / g × g c a r b o h y d r a t e × 4.186 )

2.9. Determination of Growth and Yield Attributes

Plant growth in terms of leaf fresh weight, dry weight, and yield attributes like cake weight and achene weight were determined in grams using an electronic balance. Achenes per cake were counted and noted. For growth and yield, five plants were used as a single replicate.

2.10. Quantification of Free Proline

Proline content was measured following the method of Bates et al. [25]. Frozen leaves (0.2 g) were homogenized in 5 mL 3% sulpho-salicylic acid using a pestle and mortar. After centrifugation, the supernatant was collected. One mL of the samples ninhydrin and glacial acetic acid was taken, respectively. The solution was kept in a water bath for one hour at 100 °C and then shifted to an ice bath. Two mL of toluene was mixed and kept at room temperature for 30 min until two layers were formed. The absorbance was measured at 520 nm wavelength. The same sequence was run with a blank (2 mL of toluene).

2.11. Determination of Catalase (CAT) Activity

Fresh frozen leaves were taken (0.2 g) and 50 mM 5 mL of cooled phosphate buffer (pH 7.8) was ground using a tissue grinder mortar and pestle, which was placed in an ice bath for the antioxidant enzyme extraction. The mixture solution was centrifuged at 4 °C at 14,000× g for 10 min. The supernatants from each sample were stored in Eppendorf tubes at −20 °C and utilized for antioxidant activities on CAT. Catalase activity was determined by calculating the conversion rate of H2O2 to H2O and oxygen molecules using the protocol explained by Chance and Maehly [26]. The reaction mixture was prepared by using t 300 µL of 30% H2O2 mixed in 200 mL of phosphate buffer (pH 7.0). The reaction solution (3 mL) and 0.1 mL of supernatant were added to it later. Then, the reaction was started by the addition of enzyme extract. A change in absorbance of the reaction solution was interpreted every 60 s at 240 nm. CAT activity was described as an absorbance change of 0.01 units per min.

2.12. HPLC Based Phenolic Profiling

The internal standards used were as follows: Quercetin, gallic acid, caffeic acid, benzoic acid, vanillic acid, cinnamic acid, syringic acid, p-coumaric acid, m-coumaric acid, ferulic acid, sinapic acid, and chlorogenic acid. Phenolics were extracted with minor modifications using the method of Stalikas [27]. Powdered leaves were extracted in 100% methanol (1:10) and filtered through a 0.45 µm cellulose acetate filter (EMD Millipore, Billerica, MA, USA). An aqueous suspension of the extract was then prepared using double distilled water, the pH was adjusted to 2 with 6 M HCl, and the mixture was kept at 100 °C for 3 h. All of the standards and extracts were filtered through a 0.451 m syringe membrane filter (Type Millipore) and sonicated for 15 min in a micro clean 109 bath prior to HPLC analysis. Phenolic compounds were analyzed using Gradient HPLC (LC-10AT, SCTL, Shimadzu, Kyoto, Japan), and elution was performed for 60 min at a flow rate of 1 mL/min in a gradient system of two mobile phases A (H2O2:AA-94:6, pH 2.27) and B (ACN100%).

2.13. GC-Based Fatty Acid Profiling

Fatty acid methyl esters (FAMEs) were organized by IUPAC, 1987 standard method 2.301 and determined using a Perkin Elmer gas chromatograph model Clarus 500 fitted with an Rt-2340 NB (RESTEK, Corp., 800-356-1638, Centre County, PA, USA), methyl-lignocerate-coated (film thickness 0.20 μM), polar capillary column (60 m × 0.25 mm), and an FID detector. At a flow rate of 67.4 Psi, oxygen-free nitrogen was used as the carrier gas. Other conditions were as follows: initial oven temperature, 80 °C; final temperature, 210 °C; ramp rate, 3 °C/min; injector temperature, 210 °C; detector temperature, 220 °C; and FAMEs were determined by analyzing their relative and absolute retention times with authentic standards bought from Sigma-Aldrich (Buchs, Switzerland). Fatty acid methyl esters (FAMEs) were prepared following the IUPAC official method. Paquot’s [28] standard protocol was followed for the preparation of FAMEs, which was carried out by deriving samples from fatty acid methyl esters of triglycerides via saponification of the glycosides liberation, and esterification using methanol from the fatty acids. The oil sample (0.2 g) was weighed into a 100 mL round neck round bottom flask. One pellet of KOH and 30 mL of methanol were then mixed in the flask, and the flask was then refluxed for 25 min until the droplets of fats disappeared. The reaction mixture was cooled and then gently shifted to a separating funnel, and a small amount of n-hexane was mixed. The separating funnel was stirred gently by rotating it many times, and the upper layer of hexane was separated and washed several times with 10 mL of deionized water. The hexane solution was dried over anhydrous sodium sulfate, filtered, and used for GC analysis. The dry and solvent-free methyl esters were preserved in a sealed sample tube in a deep freezer and used for further analysis.

2.14. Statistical Analysis

All collected data were subjected to statistical analysis using the software CoStat CoHort 6.4. Analysis of variance and LSD were used to evaluate the effects of factors and comparison of means.

3. Results

Statistical analysis (analysis of variance) of the data showed a non-significant effect of water reduction and SMF on leaf area, leaf dry weight, and plant height (Table 1). With water deficit, no reduction was observed in leaf number. The SMF increased this attribute under limited water conditions (Table 1). There was negligible effect of water deficit on leaf fresh weight. The comparison of means (LSD @ 0.1 level) showed that SMF of 0.2 T enhanced leaf fresh weight under stressed and non-stressed conditions.
The yield experienced a marked reduction at 50% irrigation in terms of achenes per capitulum (p ≤ 0.001) and capitulum weight (p ≤ 0.001), whereas the % of oil yield, 50 achenes weight, and achene weight/plant did not show significant differences with variation in irrigation level. The yield experienced a marked reduction (p ≤ 0.001) with reduced irrigation (50%) in terms of achenes per capitulum. All doses of SMF increased achenes number per capitulum in stressed plants. The capitulum weight notably reduced (p ≤ 0.001) under water deficit (50% irrigation) conditions. Seed treatment with 0.1 T improved this attribute in stress (Table 1).
The percentage of oil yield remained invariable in both irrigation regimes (100 and 50%). SMF seed treatment at the highest dose (0.3 T) markedly improved the percent of oil yield under limited water conditions. Fifty achene weight and achene weight per plant remained similar for both factors (irrigation and SMF) (Table 1)
On the other hand, chlorophyll a levels in untreated plants were reduced by 50% via irrigation reduction (Figure 1A). For plants grown under regular irrigation, none of the treatments had a discernible impact on this trait. However, magnetic treatment (0.1, 0.2, and 0.3 T) significantly improved this attribute in plants grown under 50% irrigation. For chlorophyll b, neither factor (irrigation and priming) was significant (Figure 1A). With 50% less irrigation, there was a noticeable reduction in the overall chlorophyll content. The total chlorophyll content in sunflower leaf samples varied with the treatment doses. In 100% irrigation, plants grown after 0.1 T magneto-priming showed a marked decrease in the chlorophyll level. However, other levels (0.2 and 0.3 T) were not significantly different from the control (unprimed plants). Fifty % irrigation in plants raised from SMF (0.1, 0.2, and 0.3 T) equally improved this attribute in sunflower leaves (Figure 1C).
Water deficiency in the growth medium had a non-significant effect on the total leaf-soluble proteins. However, seed treatment (0.1 T and 0.3 T) had a positive effect (p < 0.001) in improving total leaf soluble proteins in both water regimes (Figure 1D). Total soluble sugars experienced a decline (p < 0.01) with water deficit (50% irrigation). Higher doses (0.2 and 0.3 T) of magnetic application were successful in improving (p < 0.001) this attribute in normal and stressed plants (100 and 50% irrigations) (Figure 1E). The leaf-reducing sugars were ineffective to stress. In the case of untreated plants, the results were similar under both irrigation conditions. Plants grown after magnetic treatment showed significantly enhanced (Figure 1F) accumulation in plants grown under 100% irrigation.
All plants grown under 50% irrigation had notably (p < 0.001) higher levels of proline in comparison to 100% irrigation. None of the magnetic applications was statistically effective for this attribute (Figure 1G). The statistical analysis of the data recorded for leaf percent oil content indicated a significant difference between treatments (p ≤ 0.001). All untreated and treated plants remained indifferent under both water conditions, except for the plants grown under normal irrigation after the 0.1 T pre-sowing treatment (Figure 1H). The activity of catalase experienced a marked decline with a reduction in water in the growth medium. SMF enhanced its activity with an increase in treatment intensity (0.1 to 0.3 T), and there was a linear increase in activity (Figure 1I).
Total leaf phenolics did not experience any deterioration with water reduction (50% irrigation). Magneto-priming of 0.1 and 0.3 T further enhanced (p < 0.001) its accumulation in both water environments (100 and 50%). Magnetic application of 0.1 and 0.3 T enhanced its accumulation under 100% irrigation, whereas under 50% condition, this enhancement was observed in plants grown after 0.2 and 0.3 T magnetic treatment only (Figure 2A). Statistical analysis of leaf flavonoid data was not significant for stress and priming (Figure 2B). Leaf anthocyanin content experienced a marked (Figure 2C) decline with reduced irrigation (50% irrigation) in untreated plants. Under 100% irrigation, 0.1 T decreased its level but other higher levels remained non-significant to unprimed plants. After performing 50% irrigation, 0.1, 0.2, and 0.3 T equally enhanced their level in sunflower leaves. Data observed for leaf carotenoids varied with treatment levels and growing conditions in terms of irrigation percentage. Under normal irrigation, 0.1 treated (magnetic application) seeds had reduced content, whereas the other two levels were insignificant. Under 50% irrigation, lower levels of improved leaf carotenoids were observed (Figure 2D).
The metabolizable energy of the leaf samples remained stable under water scarcity conditions. SMF improved (p < 0.001) its level under both growing conditions at 0.1 T and 0.3 T radiation doses (Figure 2E). Water limits (under 50% irrigation) failed to affect leaf nutritional values (Figure 2F). The priming of 0.1 T and 0.3 T treatment levels significantly increased, whereas under 50% irrigation, the highest level (0.3 T magnetic seed treatment) significantly enhanced this characteristic (Figure 2A). The metabolomics profiling of selected plants grown from 0.3 T treated seeds showed better oil and achene production.
HPLC chromatograms for phenolics showed variable patterns for irrigation levels and treatment levels. In untreated plants, 50% irrigation reduced the hydroxy derivative of cinnamic acid (26.6%) compared to untreated plants. With water reduction (50% irrigation), there was a 26.6% reduction in leaf hydroxy derivatives, as described by the HPLC chromatogram (Figure 3A,B). Fifty percent irrigation in comparison to 100% irrigation showed a percent increase in quercitin (34.80), gallic acid (100), vanillic acid (62.2577), chlorogenic acid (100), syringic acid (100), p-coumaric acid (5.59), m-coumaric acid (6.29), cinnamic acid (81.15), and sinapic acid (100). Caffiec acid (121.08), benzoic acid (146.51), and ferulic acid (0.04) underwent a percent decrease in untreated plants (Figure 3). Magneto-primed plants grown under 100% irrigation had greater accumulation of gallic acid (8.2 ppm), caffiec acid (5.5 ppm), and cinnamic acid (9.5 ppm), with a reduction in caffiec acid, chlorogenic acid, and syringic acid (7.53, 291.25, 229.53, and 55.43%, respectively). Plants grown under 50% irrigation responded positively towards seed treatment in terms of enhanced accumulation of benzoic acid, p-coumaric acid, chlorogenic acid, and ferulic acid (16.87, 32.3, 10.7, and 9.72, respectively) (Figure 3C,D).
The majority of fatty acids that were determined in sunflower achenes included palmitic acid (8.01%), stearic acid (4.11%), oleic acid (36.5%), and linoleic acid (50.76%). The unsaturation-to-saturation ratio was 46.96. Water deficit reduced the unsaturation-to-saturation ratio by up to 43.63. Magnetic treatment of 0.3 T under 50% irrigation enhanced unsaturated fatty acids and decreased the level of saturated fatty acids. Ultimately, there was a marked improvement in the unsaturation-to-saturation ratio (50.81) (Table 2).

4. Discussion

As a biotechnological tool, the static magnetic field (SMF) has gained an eminent position in the literature in recent years. This is due to the successful outcomes in the number of species, including wheat [29], alafa [30], almond seeds [31], and soybean [10]. The dose and timing of positive results varies from specie to specie. In the present study, sunflower plants were treated with SMF (0.1, 0.2, and 0.3 T) and varied in growth and metabolism regimes. SMF with activated meristematic cell division and auxin enhancement modulate metabolic activities and chloroplast growth, as per previous reports. Its genomic response is linked to cryptochrome and auxin signaling [32]. In previous findings, high doses have been reported to induce fragmentation, bond rupturing, and ROS production [33]. In the current study, the SMF range used (0.1–0.3 T) was found to be safe for sunflower seeds, even when 0.3 TSMF provided the highest oil yield. In agreement with the literature, the SMF-treated sunflower plants in the present study experienced better oil and capitulum yield under 50% irrigation.
In the present and previous investigations, it was shown that when there was a water shortage, sunflower plants had lower levels of photosynthetic pigments, chlorophyll a, and total chlorophyll. This was ascribed to stress-induced photoinhibition, with restricted light harvesting and electron transport mechanisms [7]. Magnetic field treatment of seeds was employed for different crops with positive outcomes [30,31]. Calcium-dependent strategic mechanisms behind plants’ stress tolerance are associated with SMF treatment, which enhances photosynthesis [34,35,36]. SMF has recently been known to activate key enzymes of carbon metabolism, thus facilitating CO2 provision to Rubisco for direct promotion of carbon fixation [10,37,38]. Our findings seem in line with the literature; the treated plants had better photosynthetic pigments and better accumulation of products of photosynthesis, which may be related to the possible activation of Ca2+ signaling pathways and carbonic anhydrase, although these parameters were not part of the determinations of this research.
Osmoregulation is a kinase cascade-mediated phenomenon accredited with a number of enzymatic and non-enzymatic proteins [39]. In addition to this, recent advancements in proteomic studies [40] have revealed a reconnoitered link between protein concentration in plants and their photosynthetic efficiency, defense system, folding, and degradation. Therefore, lowered amino acid metabolism and ROS scavenging proteins show the sensitivity of crops under different abiotic and biotic stresses. In the present study, no reduction in total soluble proteins was observed with water limits, indicating tolerance of the variety under 50% irrigation in terms of total soluble proteins. SMF further enhanced protein content under both irrigation conditions. Possible protein-mediated improvement in ROS scavenging under drought was observed under UV-B stress in soybean [10]. There is no direct evidence supporting SMF-induced molecular activation for protein overaccumulation. However, there is evidence of variation in protein expression in different cryptochrome-1 and 2 mutants. Therefore, external SMF with the potential to affect cryptochrome may be justified for protein enhancement [41].
Carbohydrate accumulation in plants plays a crucial role in osmoregulation under drought stress, directly or by producing amino acids and organic acid-like products to maintain osmotic balance [42]. In the present study, total soluble sugars were markedly reduced with stress. Recent molecular studies have revealed decreased carbohydrate levels under drought stress with the expression of autophagy-related genes. Stress enhances their expression. Leaf glucose and fructose experience a decline in drought, which resulted in a break in metabolic balance. Such a disorder of intracellular nutrients has a positive correlation with the expression of autophagy-related genes [43]. Uncontrolled autophagy in stressed cells may lead to the formation of bubble-like autophagic vesicles. Under extreme conditions, fragmentation of cytoplasmic components may occur. SMF seed pre-treatment enhanced carbohydrates in sunflowers, which is in agreement with the results of Zareei et al. [44]. Enhancement of the sugar content in plant leaves can support plants with active regulation of osmotic pressure. This plays a crucial role in drought stress tolerance [45,46]. There are possibilities of some suppression of autophagy-related genes too. Stressed plants of sunflowers with SMF showed better photosynthesis, soluble sugars, and yield, which is in line with the existing literature.
Proline, along with other osmolytes, can regulate osmotic pressure with consequential drought tolerance [45,46]. The higher proline content in stressed sunflower plants in the present study is in line with the literature [45]. Better accumulation of proline lessens osmotic potential and water evaporation results in the instability of the intracellular environment. In contrast to the literature regarding SMF-induced enhancement of proline, there was a negligible effect of treatment on sunflowers. In the present study, untreated plants experienced a marked decline in catalase, indicating a negative impact of drought. Sunflower leaves showed enhanced activity of anthocyanin in plants grown in limited water after treatment, in accordance with previous work [31]. SMF treatment is known to alter cryptochrome-mediated protein expression [41]. Catalase protein was enhanced with SMF in the present study linearly with the applied doses (0.1 < 0.2 < 0.3). Another justification from the literature may be a possible link between the magnetic field and Ca2+ ion pumps [35], which is in turn linked to enzyme production [47].
Phenolics and flavonoids showed resistance against stress [48] and nutraceutical value. SMF enhances the phenolics in the current and earlier studies [31]. When a magnetic solution is applied to plants, it activates key enzymes of phenolic metabolism. Currently, the positive effect of SMF on the quantitative and qualitative values of phenolics may be dependent upon the possible modulation of the PAL enzyme by SMF seed treatment. This was depicted by the variable phenolic profiles in SMF-treated and -untreated plants under both water regimes (Figure 3). Anthocyanin content was reduced with lowered irrigation. There was a positive correlation between plant anthocyanin content and drought tolerance [49,50]. The decline in antioxidant anthocyanin in drought was compensated by all doses of SMF in the present study. Anthocyanin biosynthesis in plant cells was regulated by cryptochromes [51]. In magnetically treated plants, the cryptochromes act as receptors [11] and regulate the expression of genes encoding for chalcone synthase, which is a key enzyme in anthocyanin metabolism [52,53]. Cryptochrome-1 regulation is also associated with drought [54] and UV-B [11] stress tolerance. Enhanced anthocyanin content in magnetically treated stressed sunflowers is in accordance with the literature [31,44].
The variable effect of applied SMF doses could also be justified by cryptochrome-mediated magneto-perception. Its modulation and activation generally rely on electron spinning of radical pairs, spin physics, and the chemistry behind this process. The external magnetic field splits the energy of the target atom (in terms of the Zeeman Effect). The coupling between the orbital momentum and spin angular momentum needs to be cancelled for a positive effect. The intensity of the magnetic field that successfully disrupts the coupling between the orbital momentum and spin angular momentum of cryptochrome could show positive results with a sufficient Paschen–Back effect. On the other hand, a magnetic field weaker than that level may be useless or less beneficial due to the dominant coupling of spin and orbit [11,55]. In our study, the strongest SMF (0.3 T) was successful. The present results of magnetically treated sunflower leaves with enhanced carotenoids under stress are in line with the findings of Hasan et al. [56] and de Souza-Torres et al. [57]. Carotenoid accumulation supports stable photosynthesis under drought [50] and acts as an antioxidant [45].
For the use of edible oil in the food industry, WHO, 1985, and the United Nations Food and Agriculture Organization (FAO) have recommended saturated fatty acid limits up to 10% of the total caloric intake. The expert opinions of WHO and FAO have been coined to attain better prevention of chronic diseases with a healthy use of nutrition [58]. Biodiesel production from sunflowers is also correlated with a better composition of unsaturated fatty acid in oil [59]. In the present study, sunflower seeds with 87.27% unsaturated fatty acids confirmed previous findings regarding the healthy use of sunflower as edible oil. Irrigation at 50% reduced its level up to 75.84% and enhanced the saturated fatty acid composition from 12.12 to 14.99%. This decline in nutritive value induced by water deficit supported our hypothesis that a moderately tolerant variety, if grown in a water-limited environment, could experience reduced nutritive value and may need an additional supportive strategy to cope with it.
Therefore, any method that increases the amount of unsaturated fatty acids, like oleic and linoleic acid, would be a potential strategy for future sunflower oil production. Water reduction lowered oleic acid and linoleic acid in the present study. Magnetic seed treatment (0.3 T) improved the oleic acid content and total unsaturated-to-saturated fatty acid ratio (87.85%) of sunflower achene oil. Previous reports indicated the potential of magnetically treated water to increase the essential fatty acids [60]. Oleic acid and linoleic acid increased in the current study, which has some association with plant stress tolerance as well [61]. Therefore, SMF with enhanced oleic acid improved the drought tolerance of sunflowers along with their nutritive value.

5. Conclusions

In conclusion, SMF at 0.3 T could manipulate sunflower metabolism to produce better nutraceutical metabolites with enhanced drought tolerance in sunflower.

Author Contributions

Conceptualization, B.A. and Z.u.N.; methodology, R.B.; validation, N.A., S.M. and M.A.; formal analysis, F.B.; investigation, A.W.; resources, M.I.; data curation, A.A.A.; writing—original draft preparation, P.K.; writing—review and editing, B.A. and S.M.; visualization, B.A.; supervision, S.M.; project administration, A.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data supporting the reported results can be found in the manuscript and can be requested from the corresponding authors upon reasonable request.

Acknowledgments

The authors would like to extend their sincere appreciation to the Researchers’ Support, Project Number (RSP2023R236), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Havrysh, V.; Kalinichenko, A.; Mentel, G.; Mentel, U.; Vasbieva, D.G. Husk energy supply systems for sunflower oil mills. Energies 2020, 13, 361. [Google Scholar] [CrossRef] [Green Version]
  2. Khan, M.Y.; Ahmad-II, N.; Khan, M.N.; Padhi, S.K. Characterization of Used Sunflower Oil (USO) for utilization as CI engine fuel. In Proceedings of the All India Seminar on Catalysing Vision, New Delhi, India, 12 February 2020; pp. 6–7. [Google Scholar]
  3. Stahle, D.W. Anthropogenic mega drought. Science 2020, 368, 238–239. [Google Scholar] [CrossRef]
  4. Kaya, C.; Şenbayram, M.; Akram, N.A.; Ashraf, M.; Alyemeni, M.N.; Ahmad, P. Sulfur-enriched leonardite and humic acid soil amendments enhance tolerance to drought and phosphorus deficiency stress in maize (Zea mays L.). Sci. Rep. 2020, 10, 6432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kosar, F.; Akram, N.A.; Ashraf, M.; Ahmad, A.; Alyemeni, M.N.; Ahmad, P. Impact of exogenously applied trehalose on leaf biochemistry, achene yield and oil composition of sunflower under drought stress. Physiol. Plant 2020, 172, 317–333. [Google Scholar] [CrossRef] [PubMed]
  6. Mansoor, S.; Kour, N.; Manhas, S.; Zahid, S.; Wani, O.A.; Sharma, V.; Wijaya, L.; Alyemeni, M.N.; Alsahli, A.A.; El-Serehy, H.A.; et al. Biochar as a tool for effective management of drought and heavy metal toxicity. Chemosphere 2021, 271, 129458. [Google Scholar] [CrossRef] [PubMed]
  7. Arslan, Ö.; Nalçaiyi, A.B.; Erdal, Ş.Ç.; Pekcan, V.; Kaya, Y.; Çiçek, N.U.N.; Ekmekçi, Y.İ.N. Analysis of drought response of sunflower inbred lines by chlorophyll a fluorescence induction kinetics. Photosynthetica 2020, 58, 163–172. [Google Scholar] [CrossRef] [Green Version]
  8. Escalante, M.; Vigliocco, A.; Moschen, S.; Fernández, P.; Heinz, R.; Garcia-Garcia, F.; Alemano, S. Transcriptomic analysis reveals a differential gene expression profile between two sunflower inbred lines with different ability to tolerate water stress. Plant Mol. Biol. Rep. 2020, 38, 222–237. [Google Scholar] [CrossRef]
  9. Sharma, R.; Pandey, S.T.; Verma, O.; Srivastava, R.C.; Guru, S.K. Physiological seedling vigour parameters of wheat as influenced by different seed invigoration techniques. Int. J. Chem. Stud. 2020, 8, 1549–1552. [Google Scholar] [CrossRef] [Green Version]
  10. Kataria, S.; Rastogi, A.; Bele, A.; Jain, M. Role of nitric oxide and reactive oxygen species in static magnetic field pre-treatment induced tolerance to ambient UV-B stress in soybean. Physiol. Mol. Biol. Plants 2020, 26, 931–945. [Google Scholar] [CrossRef]
  11. Hammad, M.; Albaqami, M.; Pooam, M.; Kernevez, E.; Witczak, J.; Ritz, T.; Ahmad, M. Cryptochrome mediated magnetic sensitivity in arabidopsis occurs independently of light-induced electron transfer to the flavin. Photochem. Photobiol. Sci. 2020, 19, 341–352. [Google Scholar] [CrossRef]
  12. Tissot, N.; Ulm, R. Cryptochrome-mediated blue-light signalling modulates UVR8 photoreceptor activity and contributes to UV-B tolerance in Arabidopsis. Nat. Commun. 2020, 11, 1323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Turker, M.; Temirci, C.; Battal, P.; Erez, M. The effects of an artificial and static magnetic field on plant growth, chlorophyll and phytohormone levels in maize and sunflower plants. Phyton Ann. Rei Bot. 2007, 46, 271–284. [Google Scholar]
  14. Mahajan, T.S.; Pandey, O.P. Effect of electric and magnetic treatments on germination of bitter gourd (Momordica charantia) seed. Int. J. Agric. Biol. 2015, 17, 351–356. [Google Scholar]
  15. Davies, B.H.; Goodwin, T.W. Chemistry and biochemistry of plant pigments. Carotenoids 1976, 2, 38–165. [Google Scholar]
  16. Julkunen-Tiitto, R. Phenolic constituents in the leaves of northern willows: Methods for the analysis of certain phenolics. J. Agric. Food Chem. 1985, 33, 213–217. [Google Scholar] [CrossRef]
  17. Marinova, D.; Atanassova, M.; Ribarova, F.; Atanassova, M. Total phenolics and flavonoids in Bulgarian fruits and vegetables. J. Univ. Chem. Technnol. Metall. 2005, 40, 255–260. [Google Scholar]
  18. Bradford, M.M. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 1976, 72, 248–254. [Google Scholar] [CrossRef] [PubMed]
  19. Miller, G.L. Estimation of reducing sugar by dinitrosalicylic acid method. Anal. Chem. 1972, 31, 426–428. [Google Scholar] [CrossRef]
  20. Yoshida, S.; Forno, D.A.; Cock, J.H. Laboratory manual for physiological studies of rice. Int. Rice Res. Philipp. 1976, 23, 61–66. [Google Scholar]
  21. Nozzolillo, C. Anthocyanin pigments in pea seedlings: Genetically controlled and environmentally influenced. Can. J. Bot. 1978, 56, 2890–2897. [Google Scholar] [CrossRef]
  22. Horwitz, W. Official Methods of Analysis of AOAC International. Volume I, Agricultural Chemicals, Contaminants, Drugs/Edited by William Horwitz; AOAC International: Gaithersburg, MD, USA, 1997. [Google Scholar]
  23. Indrayan, A.K.; Sharma, S.; Durgapal, D.; Kumar, N.; Kumar, M. Determination of nutritive value and analysis of mineral elements for some medicinally valued plants from Uttaranchal. Curr. Sci. 2005, 89, 1252–1255. [Google Scholar]
  24. Macrae, J.; Smith, J.; Dewey, P.; Brewer, A.; Brown, D.S.; Walker, A. The efficiency of utilization of metabolizable energy and apparent absorption of amino acids in sheep given spring-and autumn-harvested dried grass. Br. J. Nutr. 1985, 54, 197–209. [Google Scholar] [CrossRef] [PubMed]
  25. 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]
  26. Chance, B.; Maehly, A.C. Assay of catalases and peroxidases. Methods Enzymol. 1955, 2, 764–775. [Google Scholar]
  27. Stalikas, C.D. Extraction, separation, and detection methods for phenolic acids and flavonoids. J. Sep. Sci. 2007, 30, 3268–3295. [Google Scholar] [CrossRef]
  28. Paquot, C. Standard Methods for the Analysis of Oils, Fats and Derivatives; Elsevier: Amsterdam, The Netherlands, 2013. [Google Scholar]
  29. Hussain, M.S.; Dastgeer, G.; Afzal, A.M.; Hussain, S.; Kanwar, R.R. Eco-friendly magnetic field treatment to enhance wheat yield and seed germination growth. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100299. [Google Scholar] [CrossRef]
  30. Dziwulska-Hunek, A.; Kornarzyńska-Gregorowicz, A.; Niemczynowicz, A.; Matwijczuk, A. Influence of electromagnetic stimulation of seeds on the photosynthetic indicators in Medicago sativa L. leaves at various stages of development. Agronomy 2020, 10, 594. [Google Scholar] [CrossRef]
  31. Abdollahi, F.; Amiri, H.; Niknam, V.; Ghanati, F.; Mahdigholi, K. Effects of static magnetic fields on the antioxidant system of almond seeds. Russ. J. Plant Physiol. 2019, 66, 299–307. [Google Scholar] [CrossRef]
  32. Jin, Y.; Guo, W.; Hu, X.; Liu, M.; Xu, X.; Hu, F.; Huang, J. Static magnetic field regulates Arabidopsis root growth via auxin signaling. Sci. Rep. 2019, 9, 14384. [Google Scholar] [CrossRef] [Green Version]
  33. Krishnan, V.; Gothwal, S.; Dahuja, A.; Vinutha, T.; Singh, B.; Jolly, M.; Sachdev, A. Enhanced nutraceutical potential of gamma irradiated black soybean extracts. Food Chem. 2018, 245, 246–253. [Google Scholar] [CrossRef]
  34. Aliniaeifard, S.; Shomali, A.; Seifikalhor, M.; Lastochkina, O. Calcium Signaling in Plants under Drought. Salt and Drought Stress Tolerance in Plants: Signaling Networks and Adaptive Mechanisms; Springer International Publishing: Cham, Switzerland, 2020. [Google Scholar]
  35. Piacentini, M.P.; Fraternale, D.; Piatti, E.; Ricci, D.; Vetrano, F.; Dachà, M.; Accorsi, A. Senescence delay and change of antioxidant enzyme levels in Cucumis sativus L. etiolated seedlings by ELF magnetic fields. Plant Sci. 2001, 161, 45–53. [Google Scholar] [CrossRef]
  36. Navazio, L.; Formentin, E.; Cendron, L.; Szabò, I. Chloroplast calcium signaling in the spotlight. Front. Plant Sci. 2020, 11, 186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Momayyezi, M.; McKown, A.D.; Bell, S.C.; Guy, R.D. Emerging roles for carbonic anhydrase in mesophyll conductance and photosynthesis. Plant J. 2020, 101, 831–844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Ogée, J.; Wingate, L.; Genty, B. Estimating mesophyll conductance from measurements of C18OO photosynthetic discrimination and carbonic anhydrase activity. Plant Physiol. 2018, 178, 728–752. [Google Scholar] [CrossRef] [Green Version]
  39. Lin, Z.; Li, Y.; Zhang, Z.; Liu, X.; Hsu, C.C.; Du, Y.; Wang, P. A RAF-SnRK2 kinase cascade mediates early osmotic stress signaling in higher plants. Nat. Commun. 2020, 11, 613. [Google Scholar] [CrossRef] [Green Version]
  40. Azri, W.; Cosette, P.; Guillou, C.; Rabhi, M.; Nasr, Z.; Mliki, A. Physiological and proteomic responses to drought stress in leaves of two wild grapevines (Vitis sylvestris): A comparative study. Plant Growth Regul. 2020, 91, 37–52. [Google Scholar] [CrossRef]
  41. Yang, Y.J.; Zuo, Z.C.; Zhao, X.Y.; Li, X.; Klejnot, J.; Li, Y.; Lin, C.T. Blue-light-independent activity of Arabidopsis cryptochromes in the regulation of steady-state levels of protein and mRNA expression. Mol. Plant 2008, 1, 167–177. [Google Scholar] [CrossRef] [Green Version]
  42. Qu, X.; Wang, H.; Chen, M.; Liao, J.; Yuan, J.; Niu, G. Drought stress–induced physiological and metabolic changes in leaves of two oil tea cultivars. J. Am. Soc. Hortic. Sci. 2019, 144, 439–447. [Google Scholar] [CrossRef] [Green Version]
  43. Luo, X.; Zhang, Y.; Wu, H.; Bai, J. Drought stress-induced autophagy gene expression is correlated with carbohydrate concentrations in Caragana korshinskii. Protoplasma 2020, 257, 1211–1220. [Google Scholar] [CrossRef]
  44. Zareei, E.; Zaare-Nahandi, F.; Oustan, S.; Hajilou, J. Effects of magnetic solutions on some biochemical properties and production of some phenolic compounds in grapevine (Vitis vinifera L.). Sci. Hortic. 2019, 253, 217–226. [Google Scholar] [CrossRef]
  45. Babeanu, C.; Dodocioiu, A.M.; Buzatu, D.G. Variation of chlorophylls and osmolytes content in three cultivars of wheat under drought stress. Ann. Univ. Craiova-Agric. Mont. Cadastre Ser. 2019, 49, 293–298. [Google Scholar]
  46. Prazeres, C.S.; Coelho, C.M.M. Osmolyte accumulation and antioxidant metabolism during germination of vigorous maize seeds subjected to water deficit. Acta Sci. Agron. 2020, 42, e42476. [Google Scholar] [CrossRef] [Green Version]
  47. Rudenko, N.N.; Borisova-Mubarakshina, M.M.; Ignatova, L.K.; Fedorchuk, T.P.; Nadeeva-Zhurikova, E.M.; Ivanov, B.N. Role of plant carbonic anhydrases under stress conditions. Plant Stress Physiol. 2021, 4, 7–38. [Google Scholar]
  48. Ghonaim, M.M.; Mohamed, H.I.; Omran, A.A. Evaluation of wheat (Triticum aestivum L.) salt stress tolerance using physiological parameters and retrotransposon-based markers. Genet. Resour. Crop Evol. 2021, 68, 227–242. [Google Scholar] [CrossRef]
  49. He, Y.; Xie, Y.; Li, X.; Yang, J. Drought tolerance of transgenic rice overexpressing maize C4-PEPC gene related to increased anthocyanin synthesis regulated by sucrose and calcium. Biol. Plant 2020, 64, 136–149. [Google Scholar] [CrossRef] [Green Version]
  50. Naderi, S.; Fakheri, B.A.; Maali-Amiri, R.; Mahdinezhad, N. Tolerance responses in wheat landrace Bolani are related to enhanced metabolic adjustments under drought stress. Plant Physiol. Biochem. 2020, 150, 244–253. [Google Scholar] [CrossRef]
  51. Christie, J.M.; Briggs, W.R. Blue light sensing in higher plants. J. Biol. Chem. 2001, 276, 11457–11460. [Google Scholar] [CrossRef] [Green Version]
  52. Chattopadhyay, S.; Ang, L.H.; Puente, P.; Deng, X.W.; Wei, N. Arabidopsis bZIP protein HY5 directly interacts with light-responsive promoters in mediating light control of gene expression. Plant Cell 1998, 10, 673–683. [Google Scholar] [CrossRef] [Green Version]
  53. Harris, S.R.; Henbest, K.B.; Maeda, K.; Pannell, J.R.; Timmel, C.R.; Hore, P.J.; Okamoto, H. Effect of magnetic fields on cryptochrome-dependent responses in Arabidopsis thaliana. J. R. Soc. Interface 2009, 6, 1193–1205. [Google Scholar] [CrossRef] [Green Version]
  54. Hamidi, H.; Masoudian, N.; Ebadi, M.; Roudi, B.; Khajehzadeh, M.H. How nitric oxide down-regulates cryptochrome 1 from canola in blue but not red light, and reduces hypocotyl dwarfism in lead and drought stress. Polish J. Environ. Stud. 2020, 29, 1161–1167. [Google Scholar] [CrossRef]
  55. Abe, M.; Itoyama, R.; Komiyama, Y.; Ito, T.; Mashimo, T.; Tojo, S. Quantitative investigation of the Zeeman and Paschen-Back effects of the hyperfine structure during the rubidium 5 2 S 1/2→5 2 D 5/2 two-photon transition. Phys. Rev A 2019, 99, 053420. [Google Scholar] [CrossRef]
  56. Hasan, M.M.; Alharby, H.F.; Uddin, M.N.; Ali, M.A.; Anwar, Y.; Fang, X.W.; Hajar, A.S. Magnetized water confers drought stress tolerance in Moringa biotype via modulation of growth, gas exchange, lipid peroxidation and antioxidant activity. Pol. J. Environ. Stud. 2020, 29, 1625–1636. [Google Scholar] [CrossRef]
  57. de Souza-Torres, A.; Sueiro-Pelegrín, L.; Zambrano-Reyes, M.; Macías-Socarras, I.; González-Posada, M.; García-Fernández, D. Extremely low frequency non-uniform magnetic fields induce changes in water relations, photosynthesis and tomato plant growth. Int. J. Radiat. Biol. 2020, 96, 951–957. [Google Scholar] [CrossRef] [PubMed]
  58. Nishida, C.; Uauy, R.; Kumanyika, S.; Shetty, P. The joint WHO/FAO expert consultation on diet, nutrition and the prevention of chronic diseases: Process, product and policy implications. Public Health Nutr. 2004, 7, 245–250. [Google Scholar] [CrossRef] [Green Version]
  59. Vanhercke, T.; Petrie, J.R.; El Tahchy, A.; Singh, S.P.; Liu, Q. Industrial Research Organization CSIRO, 2014. Processes for Producing Hydrocarbon Products. U.S. Patent 8,735,111, 27 May 2014. [Google Scholar]
  60. Mazrou, R.M.; El-Shafie, S.E.; Afify, M. Integrated effect of magnetized water and biofertilizers for enhancing the productivity of Majorana hortensis L. MJPPF 2022, 7, 31–41. [Google Scholar] [CrossRef]
  61. Zamani, S.; Naderi, M.R.; Soleymani, A.; Nasiri, B.M. Sunflower (Helianthus annuus L.) biochemical properties and seed components affected by potassium fertilization under drought conditions. Ecotoxicol. Environ. 2020, 190, 110017. [Google Scholar] [CrossRef]
Figure 1. Biochemical attributes (primary metabolites) of sunflower plants grown under 100% versus 50% irrigation after magnetic seed treatment (0, 0.1, 0.2, and 0.3 T) at the flowering stage. Similar alphabets present non-significant differences in means at 0.01 level (LSD by CoStat CoHort 6.4).
Figure 1. Biochemical attributes (primary metabolites) of sunflower plants grown under 100% versus 50% irrigation after magnetic seed treatment (0, 0.1, 0.2, and 0.3 T) at the flowering stage. Similar alphabets present non-significant differences in means at 0.01 level (LSD by CoStat CoHort 6.4).
Agronomy 13 02094 g001
Figure 2. Secondary metabolites, metabolized energy, and nutritive values of sunflower plants grown under 100% versus 50% irrigation after magnetic seed treatment (0, 0.1, 0.2, and 0.3 T) at the flowering stage. Similar alphabets present non-significant differences in means at 0.01 level (LSD by CoStat CoHort 6.4).
Figure 2. Secondary metabolites, metabolized energy, and nutritive values of sunflower plants grown under 100% versus 50% irrigation after magnetic seed treatment (0, 0.1, 0.2, and 0.3 T) at the flowering stage. Similar alphabets present non-significant differences in means at 0.01 level (LSD by CoStat CoHort 6.4).
Agronomy 13 02094 g002
Figure 3. Phenolic profile of untreated ((A) 100%/(B) 50%) and magnetic (0.3 T) ((C) 100%/(D) 50%) sunflower leaves at the flowering stage under 100% irrigation.
Figure 3. Phenolic profile of untreated ((A) 100%/(B) 50%) and magnetic (0.3 T) ((C) 100%/(D) 50%) sunflower leaves at the flowering stage under 100% irrigation.
Agronomy 13 02094 g003
Table 1. Growth and yield of sunflower raised using magnetic seed treatment (0.1. 0.2 and 0.3 T), grown under two irrigation conditions (100 and 50%). Where similar alphabets present non-significant differences in means at 0.01 level (LSD by CoStat CoHort 6.4).
Table 1. Growth and yield of sunflower raised using magnetic seed treatment (0.1. 0.2 and 0.3 T), grown under two irrigation conditions (100 and 50%). Where similar alphabets present non-significant differences in means at 0.01 level (LSD by CoStat CoHort 6.4).
Stress LevelManetic TreatmentsLeaf AreaLeaf Number Leaf Fresh Weight Leaf Dry Weight Plant HeightCapitulum Weight% Oil YieldNo. of Achenes/Capitulum50 Achenes WeightAchene Weight/Plant
100% IrrigationControl242.3 ab19.0 ab2.00 c0.38 a99.66 ab129.3 a14.8 b533 b2.55 abc16.66 cd
0.1 T272.6 a24.3 a3.18 bc0.47 a118.6 a85.3 b15.6 b296 b2.78 a43 a
0.2 T216.5 ab20.3 ab4.11 ab0.54 a105.3 ab110.3 ab17.8 ab142 e2.91 a28.33 b
0.3 T125.3 b18.3 ab2.89 bc0.39 a81.66 b104.3 ab16.5.ab118 e1.97 c20 c
50% IrrigationControl139.7 b14.3 b3.32 bc0.54 a72.66 b036.0 c15.2 b194 de2.04 c11.00 de
0.1 T156.3 b26 a2.13 c0.32 a88.00 b103.6 ab5.7 a583 b2.44 abc12.00 de
0.2 T167.8 a25.5 a4.76 a0.61 a97.66 b38 c9.6 c733 a1.91 c6.33 f
0.3 T243.0 ab25.6 a3.13 bc0.43 a84.33 b41 c20.4 d350 c2.19 abc6.66 ef
Table 2. Fatty acid composition of oil yield obtained from sunflower grown under control (100%) and limited water (50% irrigation) conditions after presowing magnetic (0.3 T) irradiation. S indicates saturated fatty acid.
Table 2. Fatty acid composition of oil yield obtained from sunflower grown under control (100%) and limited water (50% irrigation) conditions after presowing magnetic (0.3 T) irradiation. S indicates saturated fatty acid.
IrrigationsSeed TreatmentPalmitic Acid (%)Stearic Acid (%)Oleic Acid (%)Linoleic Acid (%)SumU/SRatio
100%Control8.014.1136.5150.7699.3946.957
Magnetic8.94.1136.6950.86100.5646.51
50%Control8.416.5831.8244.0290.8343.634
Magnetic6.132.9875.2812.5796.9680.31
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Afzal, B.; Nisa, Z.u.; Bashir, R.; Ali, N.; Mahmood, S.; Azeem, M.; Batool, F.; Wahid, A.; Iqbal, M.; Alsahli, A.A.; et al. Magnetic Seed Treatment Modulates Phenolic and Fatty Acid Metabolism of Sunflower under Water Scarcity. Agronomy 2023, 13, 2094. https://doi.org/10.3390/agronomy13082094

AMA Style

Afzal B, Nisa Zu, Bashir R, Ali N, Mahmood S, Azeem M, Batool F, Wahid A, Iqbal M, Alsahli AA, et al. Magnetic Seed Treatment Modulates Phenolic and Fatty Acid Metabolism of Sunflower under Water Scarcity. Agronomy. 2023; 13(8):2094. https://doi.org/10.3390/agronomy13082094

Chicago/Turabian Style

Afzal, Beenish, Zaib un Nisa, Rohina Bashir, Naila Ali, Saqib Mahmood, Muhammad Azeem, Fatima Batool, Abdul Wahid, Muhammad Iqbal, Abdulaziz Abdullah Alsahli, and et al. 2023. "Magnetic Seed Treatment Modulates Phenolic and Fatty Acid Metabolism of Sunflower under Water Scarcity" Agronomy 13, no. 8: 2094. https://doi.org/10.3390/agronomy13082094

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