Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Bulk and nanoparticles of zinc oxide exerted their beneficial effects by conferring modifications in transcription factors, histone deacetylase, carbon and nitrogen assimilation, antioxidant biomarkers, and secondary metabolism in soybean

  • Tahereh Mirakhorli,

    Roles Conceptualization, Formal analysis, Investigation, Methodology, Resources, Writing – review & editing

    Affiliation Department of Biology, Garmsar Branch, Islamic Azad University, Garmsar, Iran

  • Zahra Oraghi Ardebili,

    Roles Conceptualization, Formal analysis, Investigation, Methodology, Software, Supervision, Writing – original draft

    Affiliation Department of Biology, Garmsar Branch, Islamic Azad University, Garmsar, Iran

  • Alireza Ladan-Moghadam ,

    Roles Conceptualization, Investigation, Resources, Supervision, Validation, Writing – review & editing

    ladanmoghadam.alireza@gmail.com

    Affiliation Department of Horticulture, Garmsar Branch, Islamic Azad University, Garmsar, Iran

  • Elham Danaee

    Roles Conceptualization, Investigation, Supervision, Validation, Writing – review & editing

    Affiliation Department of Horticulture, Garmsar Branch, Islamic Azad University, Garmsar, Iran

Abstract

Nanoscience paves the way for producing highly potent fertilizers and pesticides to meet farmer’s expectations. This study investigated the physiological and molecular responses of soybean seedlings to the long-time application of zinc oxide nanoparticles (ZnO NPs) and their bulk type (BZnO) at 5 mg L-1 under the two application methods (I- foliar application; II- soil method). The ZnO NPs/BZnO treatments in a substance type- and method-dependent manner improved plant growth performance and yield. ZnO NPs transactionally upregulated the EREB gene. However, the expression of the bHLH gene displayed a contrary downward trend in response to the supplements. ZnO NPs moderately stimulated the transcription of R2R3MYB. The HSF-34 gene was also exhibited a similar upward trend in response to the nano-supplements. Moreover, the ZnONP treatments mediated significant upregulation in the WRKY1 transcription factor. Furthermore, the MAPK1 gene displayed a similar upregulation trend in response to the supplements. The foliar application of ZnONP slightly upregulated transcription of the HDA3 gene, while this gene showed a contrary slight downregulation trend in response to the supplementation of nutrient solution. The upregulation in the CAT gene also resulted from the nano-supplements. The concentrations of photosynthetic pigments exhibited an increasing trend in the ZnONP-treated seedlings. The applied treatments contributed to the upregulation in the activity of nitrate reductase and the increase in the proline concentrations. ZnO NPs induced the activity of antioxidant enzymes, including peroxidase and catalase by averages of 48.3% and 41%, respectively. The utilization of ZnO NPs mediated stimulation in the activity of phenylalanine ammonia-lyase and increase in soluble phenols. The findings further underline this view that the long-time application of ZnO NPs at low concentrations is a safe low-risk approach to meet agricultural requirements.

Introduction

Zinc (Zn) is an essential micronutrient for different kinds of living organisms, including plants. Zn performs significant fundamental roles in many biological processes such as protein synthesis, gene transcription, gene regulation, and metabolism of phytohormones [13]. In alkaline soils, plants face reduced access to Zn, a limiting factor that restricts crop yields and reduces the quality of crop-derived foods. Zn deficiency in crops not only restricts their productivity but also influences human health [4]. Hence, attempts have been made to introduce the novel, safe, and high-potent Zn fertilizers to counteract with Zn crisis and fulfill the agricultural requirements [3,4]. Taking sustainable agriculture into account, different modern technologies, in particular nanotechnology, offer novel strategies to meet the farmer’s expectations and the food demand of humans [3,4]. Nanotechnology paves the way for improving the efficiency of a vast array of commercial products such as fertilizers/pesticides in agriculture by exploitations of engineered nanoparticles (NPs), especially metal-based NPs. Nowadays, nano fertilizers can significantly contribute to fulfilling the target challenges in the sustainable agriculture and food industry. Numerous studies support this hypothesis that nanotechnology confers a considerable opportunity to meet agricultural requirements by providing novel highly-potent fertilizers/pesticides, reducing the excessive usages of conventional chemicals, and improving stress tolerance [27]. Contrary to this view, several reports underline the necessity of paying special attention to the cytotoxicity and ecotoxicity of nanoproducts [2,3,811]. Considering inadequate knowledge, the biological risk assessment of nano-based products should be, however, explored.

Taking zinc oxide nanoparticles (ZnO NPs; one of the most commonly used NPs) into account, the application of these nanomaterials improved growth and productivity in plant species, like Datura stramonium [2], wheat [12], and soybean [13]. Moreover, its application enhanced the production of secondary metabolites, like essential oils in Feverfew [14], phenolics in Melissa officinalis [15], alkaloids in Datura stramonium [2], and phenylpropanoids in tomato [4]. Exposure to ZnO NPs also affected phytohormones [3,16,17]. Besides, ZnO NPs mitigated the risk associated with diverse abiotic [7,18,19] and biotic [20] stress conditions. For instance, the foliar application of ZnO NPs mitigated the hazardous impact of chilling stress via affecting several stress-responsive genes in rice [7]. Therefore, the molecular assessment may fill the knowledge gap on mechanisms by which ZnO NPs may potentially confer stress tolerance in plants. On the other hand, ZnO NPs were associated with cytotoxicity and genotoxicity risk in Barley [11], Arabidopsis [9], and tomato [10,11].

Although a plethora of studies displayed the effects of ZnO NPs on vegetative growth at the early developmental phase, few researchers reported that these nanoparticles may positively or negatively affect plant productivity and yield at the reproductive stage, as well [4,13,21]. There, however, remains a scientific knowledge gap on how plant cells interact with ZnONP, thereby triggering these physiological responses. Several recent studies support this hypothesis that exposure of nanoparticles to plant cells may influence the transcriptome [25,7,11]. In this concern, investigation on plant transcriptional responses to ZnO NPs is necessary to elucidate the molecular mechanisms involved.

Transcription factors, mitogen-activated protein kinases (MAPKs), and epigenetic modification are critical mechanisms by which plant responses to both internal and external cues can be modulated [7,2227]. Transcription factors are proteins that contribute to the regulation of a multitude of downstream genes by recognizing specific regulatory DNA sequences of domains during signal transduction. Among large families of transcription factors, some members, such as WRKY1 [7,26], R2R3MYB [4,7], bZIP [25], bHLH [4,22], and EREB [23] play fundamental roles in regulating growth, metabolism, and stress responses. In eukaryotic living organisms, including plant cells, MAPKs act as vital mediators contributed to the gene regulation events by controlling the phosphorylation of chromatin-related proteins, transcription factors, and co-regulatory components [3,27]. In addition to the transcription factors and MAPKs, the chromatin structure is another main factor in determining the gene accessibility to the transcriptional machinery. In this regard, histone deacetylases (HDA) are key enzymes that remove the acetyl group from histone proteins on DNA, thereby epigenetically determining gene accessibility and regulating gene expression [2]. However, there is inadequate information on whether nanoparticles can change the expression of transcription factors and histone deacetylases (HDA) genes [2,22,27]. Herein, the potential involvements of these highlighted genes in controlling plant responses to ZnO NPs and the bulk counterpart were explored.

Nearly a majority of existing scientific reports on ZnO NPs-mediated responses have monitored short-time responses of plants at the early developmental stage. However, little information is gained considering the effect of multiple applications of ZnO NPs at low concentrations, mainly to avoid the potential risk associated with the application of these nanomaterials. Taking plant responses to ZnO NPs into account, there is inadequate scientific knowledge in terms of [i] long-time application at low doses, [ii] molecular mechanism, especially at a signal transduction level, and [iii] comparative evidence on foliar and soil application methods. This study was carried out to respond following main research questions; i- does ZnONP trigger epigenetic responses? ii- Which of the target transcription factors are ZnONP responsive? iii- does the supplementation of nutrient solution with ZnONP induce a long-distance response in leaves? iv- Which application method is more efficient to improve plant growth, productivity, and immunity? v- is the plant’s response to nanoparticles different from the bulk counterpart?.

Material and methods

Treatments and experimental conditions

ZnO NPs (CAS Number: 1314-13-2; Stock number: US3590) were supplied by US research nanomaterials, Inc; 3302 Twig Leaf Lane Houston, TX 77084, USA). The nanoproduct had sizes ranging from 10 to 30 nm and displayed the acceptable zeta potential amount, (- 20 mV). Field emission scanning electron microscopy (FESEM) image, UV-Vis spectrum, and Zeta potential distribution graph are depicted in S1 Fig. Zinc oxide of Sigma company was utilized as a corresponding bulk control to provide comparative data.

Soybean (Glycine max L.) seed was provided by Seed and Plant Improvement Institute, Karaj, Iran. In this experiment, plants were cultured in a soilless medium with high leaching capacity to keep the concentration applied in the foliar method equal to the amount used in the soil method. Moreover, this experiment was performed in pots under a soilless medium consisting of Cocopeat: Perlite at a ratio of 60: 40 to avoid potential interference of soil factors. In this regard, the field capacity of the soilless medium was determined, which was 500 mL per pot. To keep the concentrations of the two methods equal, each pot was irrigated with a 300 mL solution containing BZnO/ZnO NPs per pot that was equal to the volume used for the spray method as well. To keep the concentrations almost constant and prevent over-accumulation of materials in the soil medium, the pots were irrigated with the nutrient solution above the field capacity (800 mL per pot) at intervals of the bulk or nano treatments for leaching the previous materials. In soil, especially heavy clay soil, the concentrations would exponentially increase and the concentration would not remain constant. Plants were cultivated under the same natural conditions in Garmsar (southeast of Tehran, Iran; relative humidity of 60%; day/night; light intensity: 90 μmol m−2 s−1; mean temperature: 27/16°C). Thirty-day-old plants were grouped in 5 treatment groups and treated with two concentrations of ZnONP or BZnO, including 0 and 5 mg L-1, 10 times with 72 h intervals under two different application methods; I- foliar application; II- soil application. it should be noted that the main experiment was designed according to the findings of the preliminary experiment. To fully disperse the nanoparticles, ZnO NPs dissolved in deionized water were ultra-sonicated using an ultrasonicator (40 kHz) for 30 min before the treatments. The treatment groups were called as follows; C: Control; BZnO-F: Foliar application of BZnO; ZnO NPs-F: Foliar application of ZnO NPs; BZnO-S: supplementation of nutrient solution with BZnO; ZnO NPs-S: supplementation of nutrient solution with ZnO NPs; The treated plants were harvested at two developmental stages; I- the first harvest was performed 48 h after the last treatments for assessment of biomass, molecular, and physiological characteristics; II- second harvest was carried out at the end of the plant life cycle to evaluate the crop yield.

Real-time quantitative PCR (qRT-PCR)

Total RNA was purified from leaves by utilizing an RNA isolation kit (Denazist, Iran), Trizol reagent (GeneAll Biotechnology Co, South Korea), DEPC Water (BioBasic, Canada), and DNase I (Thermo Fisher, USA) according to the manufacturer’s protocol. Next, the RNA purity was confirmed based on the absorbance ratio at 260/280 nm (Nanodrop instrument; Thermo Scientific™NanoDrop Model 2000c). At the next step, the complementary DNA (cDNA) was prepared using a thermocycler instrument (PEQLAB, 96Grad) was utilized to perform. Moreover, Oligo7 and AllelID software were utilized to design the forward and reverse primer sequences for each target gene investigated. The forward and reverse sequences of primers for the monitored genes, including histone deacetylase (HDA, XM_006592247.2), heat stress transcription factor (Hsf-34; XM_003550218.2), ethylene-responsive element-binding protein (EREB; NM_001349033.1), MYB transcription factors (R2R3MYB; NM_001370267.1), WRKY1 (EU019552.1), basic/helix-loop-helix (bHLH; KT031116.1), Catalase (CAT; NM_001247898), mitogen-activated protein kinase 1 (MAPK1; AF104247), and Elongation factor (a housekeeping gene) are presented in Table 1. To investigate transcription variation among the treatment groups, the real-time quantitative PCR (Applied Biosystems StepOne™ Real-Time PCR) was conducted under a cycling program of 94°C: 120 s, 94°C: 15s, 57°C: 25 s, and 72°C: 20 s. The relative transcription of the target genes was estimated using the equation of 2-ΔΔCT in which ΔCT refers to subtracting the internal control CT amount from the CT value of each gene investigated [4,22,28].

thumbnail
Table 1. The forward and reverse sequences of primers for target genes and Elongation factor (a housekeeping gene).

https://doi.org/10.1371/journal.pone.0256905.t001

Enzyme extraction and assessments of activities of enzymes

The possible differences in some important enzymes involved in the antioxidant system, nitrogen assimilation, and phenylpropanoid metabolism were investigated. These enzymes were catalase and peroxidase, nitrate reductase, and phenylalanine ammonia-lyase (PAL). First, the leaves were homogenized in the phosphate buffer (0.1 M; pH 7.5) supplemented with ascorbate and Na2EDTA. The next step was the centrifugation of the resulting homogenates at 4°C. After that, the resulting supernatants were kept at − 80°C for the enzymatic assessments.

The catalase activity in leaves was spectrophotometrically investigated by recording the decrease in absorbance at 240 nm per min (ΔA 240 nm) in an enzyme reaction medium containing phosphate buffer and H2O2. Then, the micromole of the degraded substrate (H2O2) was calculated using extinction coefficient (ε = 39.4 mM-1 cm-1). Definition of a unit of catalase activity is the amount of enzyme that is required to decompose 1.0 μmole of the substrate (H2O2) per min. To estimate peroxidase activity, the amount of increase in the absorbance at 470 nm (ΔA 470 nm) per min following the addition of enzyme extract into the reaction medium containing guaiacol, phosphate buffer, and H2O2. Enzyme activity was calculated using the extinction coefficient (ε = 26.6 mM-1 cm-1). Unit enzyme activity is expressed as the amount of enzyme that is required for the oxidation of guaiacol into the tetra-guaiacol at 470 nm. To evaluate the nitrate reductase activity, the common protocol provided by Sym [29] was conducted. After keeping the reaction mixture (potassium nitrate, enzyme extract, and phosphate buffer) at dark conditions for 1 h, Griess reagent I and Griess reagent II were added. The procedure was continued by recording the absorbance at 540 nm. The nitrite concentration was quantified based on the standard curve of sodium nitrite and the enzyme activity was expressed in terms of μmol NO2 h-1 g−1 fw. PAL activity was estimated based on the conversion rate of phenylalanine to cinnamate during 1 h after the addition of enzyme extract to the reaction medium [30]. The reaction mixture containing phenylalanine and Tris-HCl buffer was kept at 37 ˚C for 1 hour. After that, the conversion reaction of phenylalanine to cinnamate was stopped by adding HCl. Finally, the cinnamate concentration PAL activity was quantified based on the standard curve equation of cinnamate. The PAL activity was finally expressed in terms of microgram cinnamate per hour per gram fresh weight (μg Cin. h−1 g−1 fw).

Quantification of photosynthetic pigments, proline, and soluble phenols

The ZnONP/BZnO-associated variations in the concentrations of photosynthetic pigments, including chlorophyll a (Chl a), chlorophyll b (Chl b), and carotenoids were investigated. The photosynthetic pigments were purified by homogenizing the leaves in the acetone solvent inside the mortar. The absorptions of filtrated acetonic extracts were spectrophotometrically recorded at wavelengths, including 470, 663, and 645 nm. The concentrations of Chl a, Chl b, and carotenoids were calculated using the equations presented by Lichtenthaler and Welburn [31].

To determine the concentration of proline, the sulfa salicylic acid of 3% (w/v) was utilized to extract proline. The reaction was initiated by adding 2 mL leaf extract to the reaction mixture containing 2 mL ninhydrin reagent and an equal volume of glacial acetic acid. Then, the reaction continued by heating under the water bath for 1 h and followed by immediately cooling down at an ice bath. The next step was the addition of toluene solvent and vigorously shaking by a vortex instrument. Finally, the optical density of the resulting toluene phase was spectrophotometrically monitored at 520 nm and the concentration was quantified based on the standard curve of proline [32].

To measure total soluble phenols, leaf extract was prepared with homogenizing leaves in ethanol solvent (80%) and incubating in a boiling water bath. Then, leaf extract was added to a mixture containing the Folin-Ciocalteu reagent and saturated Sodium carbonate (21%). The reaction was associated with developing a blue color. After centrifugation, the optical density of the supernatant was recorded at the wavelength of 760 nm [32]. The standard curve of tannic acid was served to determine the concentration of soluble phenols.

Statistical analysis

The experimental design was completely randomized. All data were subjected to analysis of variance (ANOVA) using GraphPad software. The differences among the mean values of treatment groups were statistically compared according to the Tukey test analysis at a level of 5% of probability. The heatmap correlation matrix was also prepared using GraphPad software to express the potential correlation between the traits monitored.

Results

The BZnO-F, ZnO NPs-F, BZnO-S, and ZnO NPs-S treatments significantly increased biomass accumulation by 24.8%, 52.7%, 14%, and 37.2%, respectively, in the shoot (Table 2). The root biomass displayed a similar upward trend in response to the BZnO-F (16.4%), ZnO NPs-F (34.9%), BZnO-S (31.7%), and ZnO NPs-S (48%) treatments (Table 2). The applied treatments significantly enhanced leaf fresh mass by an average of 37.7% over the control (Table 2). With a similar trend, both supplementation of nutrient solution and foliar application, especially the latter method, were slightly increased the number of leaves (Fig 1D). The BZnO-F (27%), ZnO NPs-F (60%), BZnO-S (15%), and ZnO NPs-S (43%) treatments contributed to improvement in crop yield, indicated by the significantly higher numbers of pods compared to the control (Table 2). The foliar application of the supplements slightly decreased the pod biomass which was the non-significant difference (Table 2). The BZnO-F and ZnO NPs-F treatments significantly increased Zn concentrations in leaves by 27% and 34%, respectively over the control group (Table 2). While the BZnO-S and ZnO NPs-S treatments led to a significant increase in leaf Zn contents by 11.5% and 19.3% compared with the control (Table 2). Likewise, the BZnO-F (23%), ZnO NPs-F (30%), BZnO-S (9.7%), and ZnO NPs-S (15%) treatment groups had higher concentrations of Zn in seeds in comparison with the control (Table 2). The foliar application method was a more efficient way to improve Zn bioaccumulation than the soil method (Table 2).

thumbnail
Fig 1.

The effects of BZnO or ZnONP in two application methods (foliar and supplementation of nutrient solution) on the transcription of several genes, including EREB [a], bHLH [b], R2R3MYB [c], HSF-34 [d], WRKY1 [e], and MAPK1 [f]; The treatment groups: C: Control; BZnO-F: Foliar application of BZnO; ZnO NPs-F: Foliar application of ZnO NPs; BZnO-S: Supplementation of nutrient solution with BZnO; ZnO NPs-S: Supplementation of nutrient solution with ZnO NPs; Different letters on the columns refer to statistically significant differences according to Tukey’s multiple comparisons test. A comparison between the two BZnO and ZnO NPs groups at the same method is presented by placing an asterisk [*] on the line, while asterisks [*] on the dashed lines define the comparison between two ZnO NPs-F and ZnO NPs-S groups. ns: Non-significant; *: 0.01 < p ≤ 0.05; **: 0.001 < p ≤ 0.01; ***: 0.0001 < p ≤ 0.001; ****: p ≤ 0.0001.

https://doi.org/10.1371/journal.pone.0256905.g001

thumbnail
Table 2. The effect of BZnO or ZnO NPs and application methods on various characteristics related to growth, yield, and Zn concentration.

https://doi.org/10.1371/journal.pone.0256905.t002

The BZnO-F, ZnO NPs-F, and ZnO NPs-S treatments significantly upregulated EREB by 3.8, 7.6, and 5.14-fold, respectively. While the difference between the BZnO-S and control groups was not statistically significant (Fig 1A). The expression of bHLH was slightly downregulated in response to the applied supplements (Fig 1B). The BZnO-F and BZnO-S treatments led to a slight significant up-regulation in the expression of R2R3MYB by an average of 2.9-fold (Fig 1C). However, the ZnO NPs-F and ZnO NPs-S treatments moderately stimulated the transcription of R2R3MYB by an average of 7.6-fold relative to the control (Fig 1C). The HSF-34 gene was significantly upregulated in response to the BZnO-F, ZnO NPs-F, and ZnO NPs-S treatments by an average of 8.3-fold, while the BZnO-S treatment made no significant change in comparison to the control (Fig 1D). The BZnO-F, ZnO NPs-F, BZnO-S, and ZnO NPs-S treatments mediated significant upregulation in WRKY1 by 6.3, 10.1, 2.3, and 7.1-fold, respectively compared to the control (Fig 1E). MAPK1 also displayed a similar upregulation trend in response to the supplements (Fig 1F).

The BZnO-F and ZnO NPs-F slightly upregulated HDA3 by an average of 1.9-fold, while this gene showed a contrary slight downregulation trend in response to the ZnO NPs-S treatment in comparison to the control (Fig 2A). The BZnO-F, ZnO NPs-F, BZnO-S, and ZnO NPs-S treatments significantly upregulated the CAT gene by 5.9, 8.4, 3.4, and 8.9-fold, respectively relative to the control (Fig 2B).

thumbnail
Fig 2.

The transcriptional responses of HDA3 [a] and CAT [b] genes following the BZnO or ZnONP treatments. The treatment groups: C: Control; BZnO-F: Foliar application of BZnO; ZnO NPs-F: Foliar application of ZnO NPs; BZnO-S: Supplementation of nutrient solution with BZnO; ZnO NPs-S: Supplementation of nutrient solution with ZnO NPs; Different letters on the columns refer to statistically significant differences according to Tukey’s multiple comparisons test. A comparison between the two BZnO and ZnO NPs groups at the same method is presented by placing an asterisk [*] on the line, while asterisks [*] on the dashed lines define the comparison between two ZnO NPs-F and ZnO NPs-S groups. ns: Non-significant; *: 0.01 < p ≤ 0.05; **: 0.001 < p ≤ 0.01; ***: 0.0001 < p ≤ 0.001; ****: p ≤ 0.0001.

https://doi.org/10.1371/journal.pone.0256905.g002

The concentrations of photosynthetic pigments, including Chl a, Chl b, and carotenoids exhibited an upward trend in response to the supplements (Fig 3A–3C). Among the treatment groups, the ZnO NPs-F group contained the highest concentration of photosynthetic pigments. The BZnO-F, ZnO NPs-F, BZnO-S, and ZnO NPs-S treatments contributed to the significant upregulation in the activity of nitrate reductase by an average of 36.8% compared to the control (Fig 3D). The BZnO-F, ZnO NPs-F, and ZnO NPs-S treatments contributed to the significant increase in the proline concentrations by 13.4%, 45.1%, and 23.6%, respectively, compared with the control (Fig 3E). The BZnO-F, ZnO NPs-F, BZnO-S, and ZnO NPs-S treatments induced the activity of the peroxidase enzyme by 22.7%, 61%, 16.8%, 35.6%, respectively, over the control (Fig 3F). Likewise, the activity of the catalase enzyme was upregulated by an average of 41% in comparison with the control (Fig 3G). The BZnO-treated plants showed higher activities of PAL enzyme by an average of 35.8%, while the ZnONP-supplemented seedlings led to up-regulation in the activity of this enzyme by 74% over the control (Fig 3H). The soluble phenols displayed a similar upward trend by an average of 41.7% (Fig 3I). The heatmap correlation matrix presented in Fig 4, clearly exhibits the strong significant correlation among the majority of molecular and physiological characteristics investigated in this study.

thumbnail
Fig 3.

The BZnO or ZnONP-mediated changes in several physiological traits, including Chl a [a], Chl b [b], carotenoids [c], nitrate reductase [d], proline [e], peroxidase activity [f], catalase activity [g], PAL activity [h], and soluble phenols [i]; The treatment groups: C: Control; BZnO-F: Foliar application of BZnO; ZnO NPs-F: Foliar application of ZnO NPs; BZnO-S: Supplementation of nutrient solution with BZnO; ZnO NPs-S: Supplementation of nutrient solution with ZnO NPs; Different letters on the columns refer to statistically significant differences according to Tukey’s multiple comparisons test. A comparison between the two BZnO and ZnO NPs groups at the same method is presented by placing an asterisk [*] on the line, while asterisks [*] on the dashed lines define the comparison between two ZnO NPs-F and ZnO NPs-S groups. ns: Non-significant; *: 0.01 < p ≤ 0.05; **: 0.001 < p ≤ 0.01; ***: 0.0001 < p ≤ 0.001; ****: p ≤ 0.0001.

https://doi.org/10.1371/journal.pone.0256905.g003

thumbnail
Fig 4. Heatmap correlation matrix among molecular and physiological traits investigated.

https://doi.org/10.1371/journal.pone.0256905.g004

Discussion

The findings confirmed that the supplements in substance type- and application method-dependent manners influenced soybean seedlings at different growth, physiological, and molecular aspects. Scholars firmly believe that Zn plays fundamental regulatory roles in a multitude of biological processes such as chromatin architecture, meristem performance, cell division, cell cycle, and metabolism of phytohormones [3]. In both application methods, the ZnO NPs or their bulk form not only did not show phytotoxic impacts but also improved plant growth performance and metabolism. The results confirmed that both material type and application method are two vital factors affecting plant responses and the efficacy of nanomaterials. The foliar application method was more efficient than the soil method, indicated by the observed significant differences in various physiological and molecular traits investigated. These findings further underline the opinion that the multiple foliar applications of ZnO NPs at low concentrations are safe low-risk approaches to improve growth, physiology, immunity, and productivity in crops. Moreover, it can be concluded that the potential risk of the foliar method is lower than the soil method as nanoparticles may adversely affect the soil microbiome [3]. The recorded physiological responses are consistent with the findings of Vafaee et al., [2] in Datura stramonium and Pejam et al., [4] in tomato. Pejam et al., [4] recently reported that the foliar utilization of ZnO NPs enhanced biomass, yield, nutritional status, and metabolism in the tomato plant.

Taking carbon metabolism into account, the supplements, especially the nano form, enhanced photosynthetic pigments, among which the increase in carotenoids is a vital protective mechanism, thereby improving plant tolerance to photoinhibition phenomenon during the stress conditions. Considering the observed variation in nitrate reductase activity and proline content, it was concluded that the efficacy of the nanoform in both application methods was more than the bulk to enhance nitrogen metabolism. The statistical analysis also confirmed the strong correlation among these markers in carbon and nitrogen primary metabolism. ZnO NPs at non-toxic doses were associated with improvement in nutritional status [4,15], carbon assimilation through photosynthesis [4,33], and nitrogen metabolism [2,4,33]. Contrary to these reports, the phytotoxicity of ZnO NPs contributed to the down-regulation in the transcription of several genes involved in the chlorophyll synthesis and photosystem structure, whereas ZnO NPs affected the production of carotenoids by upregulating genes such as PSY in Arabidopsis [9].

One of the proposed functions for ZnO NPs is to improve the resistance of plants to stress as one of the main needs of agriculture. Several studies exhibited the effectiveness of using ZnO NPs in activating defense system, stimulating secondary metabolism, and mitigating the risk of various stress conditions, such as cadmium in wheat [34], salinity in tomato [18], drought in sorghum [35], and chilling stress in rice [7]. Moreover, ZnO NPs treatment conferred systemic resistance against Tobacco Mosaic Virus through upregulation in the transcription of PR-1 (salicylic acid marker gene), CHS, POD, and PAL genes, implying activation of defense machinery [20]. However, the underlying molecular mechanisms by which ZnO NPs may confer these responses remain unknown. For this reason, this study intended to address the ZnO NPs-mediated changes in antioxidant biomarkers, secondary metabolism (PAL and soluble phenols), and molecular indices involved in regulating plant stress responses and developmental programs.

The nano form, especially in the foliar method, was more capable of remodeling the transcription of the target genes than the bulk. However, the soil application method also displayed high efficiency to trigger long-distance (root to shoot) signaling, thereby modifying the expression of genes investigated. Although the Zn concentration in the leaves of the BZnO-F group was higher than the ZnO NPs-S group, the effectiveness of the ZnO NPs-S treatment in affecting the expression of the target genes was higher than the BZnO-F group. This finding indirectly suggests the opinion that the plant response to ZnO NPs may, at least in part, resulted from the exclusive signaling of the nanomaterials rather than the effects of Zn ions. The foliar application of ZnO NPs upregulated the HDA gene, while the soil application method reversed the expression of this gene, implying the critical roles of the application methods. Little is known about epigenetic responses to nanoparticles. Several epigenetic responses such as DNA methylation [25,26,27] and histone deacetylase [2,4,27] have been reported following the application of nanomaterials. It is worth mentioning that this is the first investigation highlighting how the different application methods may be associated with differential epigenetic responses.

This study, therefore, provides a piece of molecular evidence on how exposure to ZnO NPs may contribute to activation of the defense system and stress tolerance via mediating changes in the expression of transcription factors, MAPKs, and epigenetic modification. It appears that transient alteration in cellular redox status, changes in phytohormones, and redox-based regulation are potentially responsible for the observed alterations in the expression of the investigated genes and activation of the plant immunity system. There is no denying the fact that MAPKs along with transcription factors are major components of regulatory networks by which a wide spectrum of downstream genes are regulated. It has been well documented that HSFs [4,22,36], EREB [23], WRKY1 [7,26,37], and R2R23MYB [2,4,7,22] are largely involved in the regulation of genes of the defense system and metabolism. The ZnO NPs-mediated transcriptional responses have been supported by several molecular evidence in different plant species such as Datura stramonium [2], maize [38], brassica [5], tomato [4,33], and rice [7]. The molecular assessment revealed that several transcription factors, including OsbZIP52, OsMYB4, OsMYB30, OsNAC5, OsWRKY76, and OsWRKY94, were upregulated in response to the foliar application of ZnO NPs, thereby improving plant resistance against chilling stress in rice [7]. ZnO NPs also induced genes that contributed to the antioxidant system, such as OsCATA, OsCATB, OsPRX65, OsPRX89, OsPRX11, OsCu/ZnSOD1, OsCu/ZnSOD2, and OsCu/ZnSOD3 [7]. The involvement of microRNAs in plant responses to nanoparticles has been supported in barley [11] and tomato [24]. It has been well confirmed that microRNAs have close crosslinks with transcription factors and chromatin remodeling systems [27].

In brassica, ZnO NPs in a dose-dependent manner influenced the transcriptions of the cellular expression of cation efflux transporter gene (BjCET2) and metal tolerant protein (BjMTP) and [5], implying the transcriptional involvement of ZnO NPs in plant nutrition. The application of ZnO NPs was associated with slight changes in expression of the transcription factors (by approximately 5-fold) and activation of plant immunity. These responses might result from a transient change in redox status and subsequent redox-based control of genes. The transcriptions of stress-responsive genes are modulated by redox-based management, in particular at transcriptional and post-translational levels [27]. Besides, these molecular responses explain how the application of ZnO NPs at optimum doses potentially mitigates the risk of stress conditions. The transactional pattern of bHLH displayed a contrary trend compared to the other genes investigated. This response can be a significant feedback regulatory marker, thereby controlling the expression of downstream genes. Current evidence supports the involvement of bHLHs in a multitude of biological events, including phytohormone signaling, stomatal conductance, light signaling, shoot organogenesis and morphogenesis, root development, and stress/stimulus responses [34,39]. Nakata et al., [40] provided molecular evidence on how bHLH acts as a transcriptional repressor of jasmonate and adversely influenced signaling of jasmonate, a phytohormone involved in the regulation of the tradeoff between plant growth and stress responses. In tobacco, ZnO NPs contributed to the morphological, physiological, and anatomical responses by adjusting auxin levels, tissue differentiation, metabolism, and immune system [41]. As is well known, phytohormones have close dual crosslinks with MAPKs and transcription factors. It may be, therefore, expected that the ZnO NPs-mediated changes in expression of the target genes correlate with alteration in phytohormones and signaling molecules such as nitric oxide, H2O2, and H2S; needs to be explored in the future.

Overall, the physiological and molecular assessments provide convincing evidence comparing the efficacy of ZnO NPs and bulk ZnO to modify growth performance, yield, photosynthesis, nitrogen assimilation, secondary metabolism markers, antioxidant biomarkers, and transcriptional responses. This study also provided a piece of molecular evidence indicating the fundamental role of the type of substance and the application method in influencing epigenetic modification.

Conclusion

Nanoscience paves the way for producing highly potent fertilizers and pesticides to meet farmer’s expectations. This study was conducted to fill the knowledge gap through monitoring the physiological and molecular responses of soybean seedlings to the long-time application of ZnO NPs and their bulk type at low doses under the two-application method. In both application methods, the ZnO NPs or their bulk form (especially the nano type) not only did not show phytotoxic impacts but also improved plant growth performance and metabolism. The findings further underline the opinion that the multiple application of ZnO NPs at low concentrations is a safe low-risk approach to meet agricultural requirements. The molecular findings manifest the hypothesis that exposure to ZnO NPs may widely affect the expression of transcription factors and subsequently their downstream genes, thereby influencing wide aspects of plant growth, metabolism, productivity, and immunity. The nano-form, especially in the foliar method, was more capable of remodeling the transcription of the target genes than the bulk. However, the supplementation of nutrient solution also displayed high efficiency to trigger long-distance signaling, thereby modifying the expression of genes investigated. This study is the first investigation highlighting how the different application methods may be associated with differential epigenetic responses. Taken collectively, ZnO NPs at low doses can be considered as a highly potent substance for potential utilization in plant-related sciences and industries.

Supporting information

S1 Fig.

The physiocochemical traits of ZnO NPs, including FESEM image (a), UV-Vis spectrum (b), and Zeta potential distribution graph (c).

https://doi.org/10.1371/journal.pone.0256905.s001

(TIF)

Acknowledgments

The authors would like to thank Dr. K. Khosraviani for his benevolent and professional collaborations in the research procedure.

References

  1. 1. Sheteiwy MS, Dong Q, An J, Song W, Guan Y, He F, et al. Regulation of ZnO nanoparticles-induced physiological and molecular changes by seed priming with humic acid in Oryza sativa seedlings. Plant Growth Regul. 2017; 83: 27–41.
  2. 2. Vafaee Moghadam A, Iranbakhsh A, Saadatmand S, Ebadi M, Ardebili ZO. New Insights into the Transcriptional, Epigenetic, and Physiological Responses to Zinc Oxide Nanoparticles in Datura stramonium; Potential Species for Phytoremediation. J Plant Growth Regul. 2021; pmid:33649694
  3. 3. Iranbakhsh A, Ardebili ZO, Ardebili NO. Synthesis and characterization of zinc oxide nanoparticles and their impact on plants. In: Singh V.P., Singh S., Tripathi D.K., Prasad S.M., Chauhan D.K. (eds) Plant Responses to Nanomaterials. Nanotechnology in the Life Sciences. Springer, Cham. 2021; 33–93. https://doi.org/10.1007/978-3-030-36740-4_3.
  4. 4. Pejam F, Ardebili ZO, Ladan-Moghadam A, Danaee E. Zinc oxide nanoparticles mediated substantial physiological and molecular changes in tomato. Plos one. 2021; 16(3): e0248778. pmid:33750969
  5. 5. Mazumder JA, Khan E, Perwez M, Gupta M, Kumar S, Raza K, et al. Exposure of biosynthesized nanoscale ZnO to Brassica juncea crop plant: Morphological, biochemical and molecular aspects. Sci Rep. 2020; 10(1): 1–13. pmid:31913322
  6. 6. Shang Y, Hasan M, Ahammed GJ, Li M, Yin H, Zhou J. Applications of nanotechnology in plant growth and crop protection: a review. Molecules 2019; 24(14):2558. pmid:31337070
  7. 7. Song Y, Jiang M, Zhang H, Li R. Zinc Oxide Nanoparticles Alleviate Chilling Stress in Rice (Oryza Sativa L.) by Regulating Antioxidative System and Chilling Response Transcription Factors. Molecules 2021; 26(8):2196. pmid:33920363
  8. 8. Li M, Ahammed GJ, Li C, Bao X, Yu J, Huang C, et al. Brassinosteroid ameliorates zinc oxide nanoparticles-induced oxidative stress by improving antioxidant potential and redox homeostasis in tomato seedling. Front Plant Sci. 2016; 7:615. pmid:27242821
  9. 9. Wang X, Yang X, Chen S, Li Q, Wang W, Hou C, et al. Zinc oxide nanoparticles affect biomass accumulation and photosynthesis in Arabidopsis. Front Plant Sci. 2016; 6:1243. pmid:26793220
  10. 10. Amooaghaie R, Norouzi M, Saeri M. Impact of zinc and zinc oxide nanoparticles on the physiological and biochemical processes in tomato and wheat. Botany 2017; 95(5):441–455.
  11. 11. Plaksenkova I, Kokina I, Petrova A, Jermaļonoka M, Gerbreders V, Krasovska M. The Impact of Zinc Oxide Nanoparticles on Cytotoxicity, Genotoxicity, and miRNA Expression in Barley (Hordeum vulgare L.) Seedlings. The Scientific World Journal. 2020; https://doi.org/10.1155/2020/6649746.
  12. 12. Sheoran P, Grewal S, Kumari S, Goel S. Enhancement of growth and yield, leaching reduction in Triticum aestivum using biogenic synthesized zinc oxide nanofertilizer. Biocatal Agric Biotechnol. 2021; 32: 101938.
  13. 13. Yusefi-Tanha E., Fallah S., Rostamnejadi A. and Pokhrel L.R. Zinc oxide nanoparticles (ZnO NPs) as a novel nanofertilizer: Influence on seed yield and antioxidant defense system in soil grown soybean (Glycine max cv. Kowsar). Sci Total Environ. 2020; 738: 140240. pmid:32570083
  14. 14. Shahhoseini R, Azizi M, Asili J, Moshtaghi N., Samiei L. Effects of zinc oxide nanoelicitors on yield, secondary metabolites, zinc and iron absorption of Feverfew (Tanacetum parthenium L. Schultz Bip.). Acta Physiol Plant. 2020; 42(4): 1–18.
  15. 15. Babajani A, Iranbakhsh A, Ardebili ZO. 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(24): 24430–44. https://doi.org/10.1007/s11356-019-05676-z.
  16. 16. Jiang M, Wang J, Rui M, Yang L, Shen J, Chu H, et al. OsFTIP7 determines metallic oxide nanoparticles response and tolerance by regulating auxin biosynthesis in rice. J Hazard Mater. 2021; 403: 123946. pmid:33264991
  17. 17. Vankova R., Landa P, Podlipna R, Dobrev PI, Prerostova S, Langhansova L. et al. ZnO nanoparticle effects on hormonal pools in Arabidopsis thaliana. Sci Total Environ. 2017; 593: 535–542. pmid:28360003
  18. 18. Faizan M, Bhat J, Chen C, Alyemeni MN, Wijaya L, Ahmad P, et al. Zinc oxide nanoparticles (ZnO-NPs) induce salt tolerance by improving the antioxidant system and photosynthetic machinery in tomato. Plant Physiol Biochem. 2021; 161: 122–130. pmid:33581620
  19. 19. Khan AR, Wakeel A, Muhammad N, Liu B, Wu M, Liu Y, et al. Involvement of ethylene signaling in zinc oxide nanoparticle-mediated biochemical changes in Arabidopsis thaliana leaves. Environ Sci Nano. 2019; 6(1):341–55.
  20. 20. Abdelkhalek A, Al-Askar AA. Green synthesized ZnO nanoparticles mediated by Mentha Spicata extract induce plant systemic resistance against Tobacco mosaic virus. Appl Sci. 2020; 10(15): p.5054.
  21. 21. Medina-Perez G, Fernandez-Luqueno F, Trejo-Tellez LI, Lopez-Valdez F, Pampillon-Gonzalez L. Growth and development of common bean (Phaseolus vulgaris l.) var. pinto saltillo exposed to iron, titanium, and zinc oxide nanoparticles in an agricultural soil. Appl Ecol Environ Res. 2018; 16:1883–1897.
  22. 22. Mirakhorli T, Ardebili ZO, Ladan-Moghadam A, Danaee E. Nitric Oxide Improved Growth and Yield in Soybean (Glycine max) by Mediating Physiological, Anatomical, and Transcriptional Modifications. J Plant Growth Regul. 2021; pmid:33649694
  23. 23. Abedi S, Iranbakhsh A, Oraghi Ardebili Z. Nitric oxide and selenium nanoparticles confer changes in growth, metabolism, antioxidant machinery, gene expression and flowering in chicory (Cichorium intybus L.), potential benefits and risk assessment. Environ Sci Pollut Res. 2021; 28: 3136–3148. pmid:32902749
  24. 24. Neysanian M, Iranbakhsh A, Ahmadvand R, Oraghi Ardebili Z, Ebadi M. Comparative efficacy of selenate and selenium nanoparticles for improving growth, productivity, fruit quality, and postharvest longevity through modifying nutrition, metabolism, and gene expression in tomato; potential benefits and risk assessment. PloS one. 2020; 15(12): e0244207. pmid:33338077
  25. 25. Sotoodehnia-Korani S, Iranbakhsh A, Ebadi M, Majd A, Oraghi Ardebili Z. Selenium nanoparticles induced variations in growth, morphology, anatomy, biochemistry, gene expression, and epigenetic DNA methylation in Capsicum annuum; an in vitro study. Environ Pollut. 2020; 265: 114727. pmid:32806441
  26. 26. Rajaee Behbahani S, Iranbakhsh A, Ebadi M, Majd A, Ardebili ZO. Red elemental selenium nanoparticles mediated substantial variations in growth, tissue differentiation, metabolism, gene transcription, epigenetic cytosine DNA methylation, and callogenesis in bitter melon (Momordica charantia); an invitro experiment. Plos One. 2020; 15(7): e0235556. pmid:32614916
  27. 27. Iranbakhsh A, Aredebili ZO, Ardebili NO. Gene regulation by H2S in plants. Chapter 9 in Hydrogen Sulfide in Plant Biology, Academic press, Elsevier; https://doi.org/10.1016/B978-0-323-85862-5.00014-2.
  28. 28. Ghasempour M, Iranbakhsh A, Ebadi M and Ardebil ZO. Multi-walled carbon nanotubes improved growth, anatomy, physiology, secondary metabolism, and callus performance in Catharanthus roseus, an in vitro study. 3 Biotech. 2019; 9(11): 404. pmid:31681525
  29. 29. Sym GJ. Optimisation of the in-vivo assay conditions for nitrate reductase in barley (Hordeum vulgare L cv Igri). J. Sci. Food Agric.1984; 35: 725–30.
  30. 30. Beaudoin-Eagan LD, Thorpe TA. Tyrosine and phenylalanine ammonia lyase activities during shoot initiation in tobacco callus cultures. Plant Physiol.1985; 78: 438–441. pmid:16664262
  31. 31. Lichtenthaler H, Wellburn A. Determination of total carotenoids and chlorophylls a and b of leaf extracts in different solvents. Biochem Soc Trans.1983; 603: 591–592.
  32. 32. Tabatabaee S, Iranbakhsh A, Shamili M, Oraghi Ardebili Z. Copper nanoparticles mediated physiological changes and transcriptional variations in microRNA159 (miR159) and mevalonate kinase (MVK) in pepper; potential benefits and phytotoxicity assessment. J. Environ. Chem. Eng. 2021; https://doi.org/10.1016/j.jece.2021.106151.
  33. 33. Sun L, Wang Y, Wang R, Wang R, Zhang P, Ju Q, Xu J. Physiological, transcriptomic, and metabolomic analyses reveal zinc oxide nanoparticles modulate plant growth in tomato. Environ Sci Nano. 2020; 7(11): 3587–3604.
  34. 34. Hussain A, Ali S, Rizwan M, Rehman MZ, Javed MR, Imran M et al. Zinc oxide nanoparticles alter the wheat physiological response and reduce the cadmium uptake by plants. Environ Pollut. 2018; 242: 1518–1526. pmid:30144725
  35. 35. Dimkpa CO, Singh U, Bindraban PS, Elmer WH, Gardea-Torresdey JL, White J. Zinc oxide nanoparticles alleviate drought-induced alterations in sorghum performance, nutrient acquisition, and grain fortification. Sci Total Environ. 2019; 688: 926–934. pmid:31726574
  36. 36. Safari M, Ardebili ZO, Iranbakhsh A. Selenium nano-particle induced alterations in expression patterns of heat shock factor A4A (HSFA4A), and high molecular weight glutenin subunit 1Bx [Glu-1Bx] and enhanced nitrate reductase activity in wheat (Triticum aestivum L.). Acta Physiol Plant. 2018; 40: 117.
  37. 37. Iranbakhsh A, Ardebili ZO, Molaei H, Ardebili NO, Amini M. Cold plasma up-regulated expressions of WRKY1 transcription factor and genes involved in biosynthesis of cannabinoids in Hemp (Cannabis sativa L.). Plasma Chem Plasma Process. 2020; 40: 527–537.
  38. 38. Xun H, Ma X, Chen J, Yang Z, Liu B, Gao X, et al. Zinc oxide nanoparticle exposure triggers different gene expression patterns in maize shoots and roots. Environmental Pollution. 2017; 229: 479–88. pmid:28624629
  39. 39. Mao K., Dong Q., Li C., Liu C. and Ma F. Genome wide identification and characterization of apple bHLH transcription factors and expression analysis in response to drought and salt stress. Frontiers in plant science. 2017; 8: p.480. pmid:28443104
  40. 40. Nakata M, Mitsuda N, Herde M, Koo AJ, Moreno JE, Suzuki K. et al. Howe, A bHLH-type transcription factor, ABA-INDUCIBLE BHLH-TYPE TRANSCRIPTION FACTOR/JA-ASSOCIATED MYC2-LIKE1, acts as a repressor to negatively regulate jasmonate signaling in Arabidopsis. Plant Cell. 2013; 25(5): 1641–1656. pmid:23673982
  41. 41. Tirani MM, Haghjou MM, Ismaili A. Hydroponic grown tobacco plants respond to zinc oxide nanoparticles and bulk exposures by morphological, physiological and anatomical adjustments. Functional Plant Biology. 2019; 46(4):360–75. pmid:32172745