Next Article in Journal
Chemical and Biological Control of Wet Bubble Disease (Hypomyces perniciosus) in Mushroom Crops
Previous Article in Journal
A Bibliometric Analysis of Research on Apple Snails (Ampullariidae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of 24-Epibrassinolide on the Energetic Parameters and Early Stages of Growth and Development in Seedlings of Two Maize (Zea mays L.) Genotypes

1
Faculty of Ecology and Environmental Protection, University Union-Nikola Tesla, Cara Dušana 62-64, 11000 Belgrade, Serbia
2
Institute for Plant Protection and Environment, University of Belgrade, Teodora Drajzera 9, 11000 Belgrade, Serbia
3
Institute of General and Physical Chemistry, Studentski trg 12/V, 11000 Belgrade, Serbia
4
Faculty of Chemistry, University of Belgrade, Studentski trg 12-16, 11158 Belgrade, Serbia
5
“Vinča” Institute of Nuclear Sciences—National Institute of the Republic of Serbia, University of Belgrade, Mike Petrovića Alasa 12-14, 11000 Belgrade, Serbia
6
Faculty of Physical Chemistry, University of Belgrade, Studentski trg 12-16, 11158 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Agronomy 2023, 13(7), 1673; https://doi.org/10.3390/agronomy13071673
Submission received: 9 May 2023 / Revised: 12 June 2023 / Accepted: 18 June 2023 / Published: 21 June 2023
(This article belongs to the Section Crop Breeding and Genetics)

Abstract

:
Brassinosteroids (BRs) are a class of plant hormones that play important roles in regulating various physiological and developmental processes in plants. One of the most effective BRs involved in modulating crop growth is 24-epibrassinolide (24-EBL). The effects of different concentrations of 24-EBL on various biochemical and biophysical parameters critical to early growth stages and seedling development were investigated using two maize hybrids, ‘ZP 434’ (a new-generation hybrid) and ‘ZP 704’ (an older-generation hybrid). The evaluation of results is based on measurements of germination percentage, morphometric parameters, redox status, comparative analysis of thermodynamic parameters (such as Gibbs free energy, enthalpy, entropy), and the concentration of specific sugars in different parts of maize seedlings. The results indicate that the germination and initial growth of maize seedlings are influenced by the flow of crucial sugars from the remaining seed (as a source of nutrients) towards the plumule and radicle (as sink organs). Furthermore, alterations in Gibbs free energy play a significant role in these sugar transfers within the maize seedlings. The seed germination was most affected by the highest concentrations of 24-EBL, showing inhibitory effects, whereas lower and moderate concentrations of exogenously added 24-EBL exhibited a beneficial influence on the initial phases of seedling growth. The mentioned approach gives new insights into source–sink relationships and can be used as a quantitative measure of the germination energy, which until now has been a qualitative criterion in seed science.

Graphical Abstract

1. Introduction

Brassinosteroids (BRs) are one of the last identified classes of phytohormones [1], with a coordinating effect on the various developmental processes of plants, thereby decisively affecting the processes of growth and development in normal and stressful environmental conditions [2,3,4,5,6,7,8]. The coordination action of BRs is reflected in complex signaling pathways, triggered by the binding of BR molecules to a specific plasmalemma receptor [5,9]. The BR signaling pathways interfere with other signaling pathways in the plant cell: other phytohormones and light reception systems (e.g. phytochrome system, etc.), as well as other signaling pathways (Ca signals, signals from lipid molecules, etc.) [5,9,10,11,12].
Seed germination and seedling establishment are developmental stages of plants that are highly sensitive to variations in environmental conditions, where the impact of BRs is crucial (late frost, early drought, parasite attacks, etc.) [13]. 24-Epibrassinolide (24-EBL) is a phytohormone belonging to the BR group. Studies have shown that 24-EBL has a significant effect on seed germination, seedling growth, and plant development. It acts by interacting with specific receptors in plants, activating signaling pathways that regulate growth and development processes. In agricultural production, the application of 24-EBL can be useful for improving yield and improving plant quality and resistance to stressful conditions. However, it is important to note that the optimal concentrations and application time of 24-EBL may vary depending on the plant species, the stage of plant development, and growing conditions [14,15].
The energy required for developmental activities in young plants is derived from the breakdown of stored starch. The regulation of seed germination and growth is influenced by BRs affecting α-Amylase expression and activity and the degradation of starch in the endosperm in rice (Oryza sativa) [16]. This process involves the conversion of starch into intermediate saccharides and eventually into monosaccharides, primarily glucose. These monosaccharides then act as a substrate for respiration and other synthetic processes during the plant’s growth and development [17,18]. Glucose and its derivatives play a significant role in signaling processes as indicators of carbon status in plant cells via hexokinase enzymes as “sugar sensors” [19], thereby directing carbon metabolism toward catabolism or anabolism by modulating different sets of genes. Sucrose, as a disaccharide, is the major transport sugar in the phloem of the vast majority of higher plants [20] in addition to having a signaling function similar to glucose, reflecting the carbon status of plant cells [21]. Some other mono-, di- and tri-saccharides (e.g., fructose, arabinose, raffinose) have different functions in plant metabolism, such as acting as monomers or products of degradation of different more complex saccharides with different functions, e.g., in cell walls and in seeds [22,23]. BRs may influence the status of this saccharides, possibly to affect homeostasis of the cell wall of plants by acting on the ester linkages of pectin and polyphenolic compounds [24]. Observing the various effects of 24-EBL on different sugars, the influence of 24-EBL on disaccharide trehalose, which contributes to the protection of plant proteins under stress [25,26,27], common when forming or germinating seeds, is of great importance for germination and the early stages of seedling development.
During the initial stages of germination, seeds present a metastable and unstable system with a high risk of disruption of structures and functions, mainly due to the production of reactive oxygen and nitrogen species and oxidative stress [17,28,29], so it is important to determine what make it possible to overcome unstable stages (crisis stage) in the development of each plant, called seedling establishment. Recent results suggest that there is a certain tendency to maintain the stability of these metastable structures of living matter through controlled modulation of system entropy [30,31] and other thermodynamic parameters, such as enthalpy and Gibbs free energy, which may be of interest in considering seed germination and the first stages of seedling growth. The Gibbs free energy in seedlings refers to the amount of free energy available or required during growth and development [32]. During the growth of maize seedlings under the influence of 24-EBL, there are changes in the chemical composition of the plant, such as the breakdown of stored starch in the seed and conversion to saccharides, which provides energy for the initial growth of the seedling and which result in changes in the Gibbs free energy [33].
Among various thermodynamic approaches which found applications in plant systems [34,35], the cybernetic approach based on the Aristide Lindenmayer’s L-Systems [36,37] has proven fruitful in modeling the various situations that occur during plant growth and development [38,39] and even in the case of the very complex, so far mostly qualitatively described source–sink relationships, which are very important in determining the yield of cultivated plants [40]. The problem of quantitative modeling of plant systems, which would take into account the interconnections of biochemical, energetics, and information parameters, has not yet been satisfactorily resolved. This approach is very demanding, but it can be applied to plant systems such as germinating seeds, seedlings, or small model plants, such as Arabidopsis sp.
One of the areas where BRs have been shown to have an influence is in the regulation of source–sink relations in plants, where source tissues such as leaves are responsible for photosynthesis and the production of carbohydrates, while sink tissues such as roots, flowers, and developing seeds are responsible for using these carbohydrates for growth and development [41,42,43,44]. Studies have shown that BRs can influence source–sink relations by regulating the expression of genes involved in carbohydrate metabolism, transport, and partitioning [45,46]. A thorough explanation of the biochemical and biophysical stages that govern these processes still needs to be provided. It is crucial to comprehend the impact of BRs on the mechanisms underlying plant responses in order to anticipate an optimal source–sink relationship.
The intention of this work was to test the effect of BRs on processes in plants by testing physicochemical processes in maize seedlings, which are also a model system for both source–sink relationships and plant seed germination processes. The actual innovative approach provides a more comprehensive understanding of plant seedling growth and the effects of 24-EBL on maize seedlings from a thermodynamics standpoint. Instead of the existing qualitative indicators of plant seed germination (“germination energy”), the same process can be described quantitatively, in real thermodynamic terms, by using parameters, such as Gibbs free energy, enthalpy, entropy, so that the germination of seeds of various genotypes can be compared quantitatively. In addition, it can help determine the optimal dose of 24-EBL that can enhance the growth and development of the seedlings without compromising their energetics or thermodynamic stability.

2. Materials and Methods

2.1. Plant Material, Treatments, and Growth Conditions

Two maize hybrids, ZP434 (drought tolerant) and ZP 704 (older-generation hybrid, which is a standard hybrid, more susceptible to stressful conditions), were used in the experiments. The seeds were produced in the “Maize Research Institute ZEMUN POLJE”, Republic of Serbia. Concentrations of 24-EBL used for treatments and growth conditions were previously described and confirmed in several papers [2,47,48,49,50,51], and based on findings (inhibitory and promotion effect), a range of concentrations from 5.2 × 10−9 to 5.2 × 10−15 M were chosen [52].

2.2. Seed Germination and Determination of Morphometric Parameters of Seedlings

Seed germination and determination of morphometric parameters of seedlings was performed using International Seed Testing Association (ISTA) rules for seed testing [49,51]. The 800 previously measured seeds were divided into 4 equal portions, a control group and groups intended to be treated with different targeted concentrations of 24-EBL (5.2 × 10−9, 5.2 × 10−12, and 5.2 × 10−15 M). Seeds were surface sterilized with 0.5% (v/v) sodium hypochlorite solution and washed thoroughly with several changes of sterile distilled water. Each group of 200 seeds was germinated in 4 replicates (50 seeds per one box) in 2 L plastic boxes, on filter paper sheets. At the beginning of the experiment, each replicate was topped with 60 mL of different concentrations of 24-EBL solution, and the control was topped with distilled water [50], under the phytothrone (Loške tovarne hladilnikov Škofja Loka, d.d., Slovenia) conditions at 24 °C (over day) and 21 °C (overnight), with a 12 h of light (110–160 μmol photons m−2 s−1)/12 h of dark regime [51,52]. After 7 days, 25 uniformly grown seedlings from each box (in total, 100 seedlings per treatment) were divided with a scalpel into plumule, radicle, and RoS. Seedling parts chosen for the further experiments were measured using an analytical balance (Ohaus Pioneer, model PA413) and stored in a deep freezer at −70 °C.

2.3. Electron Paramagnetic Resonance Spectroscopy

Electron paramagnetic resonance spectroscopy (EPR) was used to assess the total redox status associated with the production of the reactive oxygen species (ROS) and reactive nitrogen species (RNS) in seedlings of the maize hybrids. All EPR recordings were performed on a Elexsys II E-540 X/L EPR spectrometer (Bruker, Rheinstetten, Germany) in the X-band. During the EPR measurements, experimental parameters were as follows: Center field 3506.3 G, field width 75 G, microwave frequency 9.85 GHz, microwave power 6.325 mW, modulation amplitude 2 G, and modulation frequency 100 kHz. Quantification of the EPR results consisted of measuring the remaining signal in the system after the reduction of the pyrrolidine membrane-permeable spin probe 3CP by maize seedlings. A 30 µL 0.075 mM 3CP solution dissolved in deionized water was used as the control (Sigma–Aldrich, Steinheim, Germany). Whole seedlings of measured mass were immersed in 10 mL of 0.075 µM 3CP solution. After 60 min, 30 µL of the 3CP solution was sampled and transferred into the EPR spectrometer for signal reading. Sampling was performed by drawing the solution into 5 cm long gas-permeable Teflon tubes (Zeus industries, Raritan, NJ, USA) with a wall thickness of 0.025 mm and an inside diameter of 0.6 mm. The obtained EPR signals were compared with the control by assessing the double integral values of the EPR spectra. All recordings were performed at room temperature and analyzed using Xepr-2.6b.170 software (Bruker GmbH, Billerica, MA, USA).

2.4. Thermodynamic Calculations

Isothermal measurements of thermodynamic parameters were performed at a high temperature in an oven (Carbolite Gero GmbH & Co. KG, Neuhausen, Germany) operating at a heating rate of 30 °C min−1. All measurements were carried out in a static oxygen atmosphere (without continuous gas flow). The previously separated portions of the 25 seedlings of both maize hybrids (ZP 434 and ZP 704) treated with appropriate concentrations of exogenously added 24-EBL (5.20 × 10−9 M, 5.20 × 10−12 M, and 5.20 × 10−15 M and control) were individually thermally treated at T = 60, 105, and 130 °C. Samples were placed in an oven and subjected to temperature treatment in the time range t = 8 min to t = 30 min. Subsequently, the samples were placed in a desiccator and then measured. The masses recorded represent the average mass losses of the sample. Each experiment at a given operating temperature was repeated three times. Water activity in the system can be described as the “effective” water content. The temperature dependence of the change in water activity (isothermal displacement) can be described by the Claussius–Clapeyron equation:
ln a w 2 a w 1 = q + λ w R = 1 T 2 1 T 1 ,
where q is the heat of sorption, λw is the latent heat of evaporation for water (44.0 kJ kg−1 at 25 °C), and R is the universal gas constant, while aw1 and aw2 are water activities at different temperatures (T1 and T2) at a given equilibrium water content. Changes in external (temperature and/or humidity) and internal factors (“glassy state” stability) affect the energy status of seedlings with respect to the respiratory processes of living cells, which can be estimated (calculated) using the sorption isotherm to which the following equations for different thermodynamic parameters apply [28]:
d H o = R T 1 T 2 T 2 T 1 ln a w 1 a w 2 ,
d G o = R T 2 T 1 ln a w 1 a w 2 ,
and
d S o = d H o d G o d T ,
where at a given water content, aw1 and aw2 represent relative humidity at lower and higher temperatures T1 and T2, respectively. dHo, dGo, and dSo represent differential enthalpy, differential Gibbs free energy, and differential entropy dehydration, respectively. The total (cumulative) change in Gibbs free energy can be obtained from equation [28]:
Δ G o = Δ H o T i Δ S o ,
where Ti represents the selected temperature. Other definitions of thermodynamic parameters are given in [49,53]. Obtained results are not shown and were used for statistical evaluation.

2.5. Statistical Evaluation

The results represent the mean values of the measurements on the three experimental samples. Multivariate linear regression analysis (Partial Least Square Regression—PLSR method) was used to evaluate the relationship between the content of selected sugars (trehalose, arabinose, glucose, fructose, sucrose, raffinose (results of sugars used for statistical evaluations can be founded in refs. [49,53])) and differential Gibbs free energy (ΔGo; kJ mol−1) as a measure of synthetic processes in the maize hybrids. PLSR can analyze strongly collinear data, reducing the high-dimensional data matrix to a much smaller and interpretable set of latent variables (LVs). LVs were calculated for both independent and dependent variable matrices, plus a relationship between them. The quality of the models was monitored with the following parameters: (i) R2cal, the cumulative sum of squares of the Ys explained by all extracted components, and R2CV, the cumulative fraction of the total variation of the Ys that can be predicted by all extracted components, and (ii) root mean square errors of calibration (RMSEC) and root mean square errors of cross-validation (RMSECV). Validation of the models was performed using venetian blinds cross-validation. The data were mean-centered and scaled to unit variance before statistical analyses. PLSR was carried out using PLS ToolBox, v. 6.2.1 (http://www.eigenvector.com/software/pls_toolbox.htm, accessed on 15 February 2023, Eigenvector Research, Inc., Wenatchee, WA 98801, USA) for MATLAB (7.12.0 (R2011a)).
The results of seed germination, morphometric parameters, and total redox status were obtained from the average of four biological replicates. Data significant difference was analyzed using one-way ANOVA with Fisher’s Least Significant Difference (LSD) post hoc test. Analyses were performed using Microsoft Excel 2016 (Data Analysis Package).

3. Results and Discussion

3.1. Influence of 24-EBL on Seed Germination

Based on the obtained results (Figure 1), it can be seen that different concentrations of 24-EBL have a different effect on the germination of hybrid seeds ZP 704 and ZP 434. The greatest effect is observed at the highest concentrations, where germination is 93% for ZP 434 and 92% for ZP 704, unlike the control samples, whose germination rate is almost 96%. Observing the differences in the influence of different concentrations on germination, small variations are observed at all tested concentrations except for the highest concentration. In addition, taking into consideration differences between the hybrids, it can be seen that the hybrid ZP 704 shows a higher sensitivity than the hybrid ZP 434. Taking into consideration the results shown in Figure 1 and Figure 2, as well as our previous study, there is an indication of a better germination and a higher Vigor index (as an important quality parameter combining seedling growth rate and dry weight) [53] of the maize hybrid ZP 434 in comparison to the hybrid ZP 704 at a higher gradient of concentration of the 24-EBL solution in which the seeds of these maize genotypes germinated. This finding is in accordance with the declared higher resistance of the plants of the ZP 434 hybrid to stressful environmental conditions confirmed by Waisi [49].

3.2. Determination of Morphometric Parameters of Treated Seedlings

The hybrid ZP 434 showed a positive impact on plumule elongation for all 24-EBL concentrations, with the exception of the highest concentration (5.2 × 10−9 M), as compared to the control. Previous results suggest that higher concentrations (5.2 × 10−8, 5.2 × 10−7 M) have an inhibitory effect, while concentrations of 5.2 × 10−6 M have a total inhibitory effect on the germination of seeds and growth [53]. The results from Figure 2A suggest that the hybrid ZP 704 exhibited plumule length values similar to control samples only at the lowest concentration of 24-EBL (5.2 × 10−15 M), while higher concentrations inhibited plumule length. It is important to note that the germination of seeds and the elongation of seedlings are highly influenced by the ABA and the combined effect of BRs and GAs [54,55,56]. In germinating cereal grains, GAs trigger the expression of a number of genes encoding enzymes such as α-Amylase, crucial for the breakdown of starch reserves [57,58].
The mechanism of action of BRs on root growth is complex and includes the interaction of hormones with receptors on the cell membrane. This interaction leads to the activation of signaling pathways that influence the growth and differentiation of root cells. One of the main effects of BRs on root growth is the stimulation of cell elongation, leading to root elongation [59]. A series of studies performed on different plant species showed that the application of BRs usually leads to an increase in root length. In a study conducted on barley (Hordeum vulgare L.), the application of 24-EBL led to a 30% increase in root length compared to the control group [60]. Similar effects of BRs have been reported in other crops, such as maize (Zea mays L.), tomato (Solanum lycopersicum L.), and soybean (Glycine max L.) [59]. However, the results of the present study show that there are slight differences, but they are not statistically significant (Figure 2B). Taking into account the differences in the literary data (some point to an increase and others to a decrease in root length), it can be assumed that the influence of 24-EBL is different for different species, even hybrids [59,61]. It is necessary to carry out analyses on a large number of species and varieties in order to be able to draw conclusions about the precise application of 24-EBL in agriculture.

3.3. Impact of 24-EBL on The total Redox Status

It can be observed from Figure 3 that the EPR signal magnitude of the 24-EBL-treated maize seedlings for the concentration range from 5.2 × 10−12 to 5.2 × 10−9 M is greater for the ZP 704 seedlings than for the ZP 434 seedlings, indicating a greater spin probe reduction potential of ZP 434 in comparison to ZP 704 hybrids and/or greater spin probe reoxidation potential of ZP 704 in comparison to ZP 434 hybrids. These results could derive from either lower antioxidant content in ZP 704 in comparison to in ZP 434 hybrids, higher production of ROS and RNS in ZP 704 in comparison to in ZP 434 hybrids, or both.
The increased presence of ROS and RNS in both hybrids could be attributed to the regular auxin-mediated relaxation of the cell wall, which is a necessary part of the cell elongation process for seedlings [62]. The level of response to stress is noticeably higher in hybrid ZP 434 compared to hybrid ZP 704, and this can be attributed to increased concentrations of polyphenolic compounds during the germination of hybrid ZP 434, which has more polar phenolics in both the root and shoot compared to ZP 704 [63]. Auxins and BRs are known to act independently but in synergy [2,3], especially since it was found that one of the first genes found to be induced by the action of BRs was xyloglucan endotransglucosylase (XET; [3]), one of the key enzymes in cell wall remodeling during plant cell elongation. The aforementioned enzyme is also known to be controlled by auxin-mediated acidification of the cell wall [23]. Furthermore, the exposure of treated ZP 704 seedlings to 24-EBL could lead to oxidative stress, which is a common occurrence during seed imbibition and rehydration in the first growth stages of maize seedling. Oxidative stress is a common event in the rehydration of dry seed tissue [64]. It is also known that the production of ROS occurs in the cell wall synthesis process [23], which is relatively common in the early stages of growth and development of maize seedlings, a plant of subtropical origin [29], especially in stressful conditions.

3.4. Influence of 24-EBL on Thermodynamic Changes in Maize Seedlings

The model parameters of the regression relationship between the content of selected sugars (trehalose, arabinose, glucose, fructose, sucrose, raffinose) and differential thermodynamic parameters (entropy, enthalpy, and Gibbs free energy) assessed in 7-day-old maize seedlings (hybrids ZP 434 and ZP 704) are presented in Table 1 and Table 2.
The Gibbs free energy can be used to estimate the thermodynamic feasibility of a particular reaction or process in maize. A negative Gibbs free energy value indicates that a reaction is spontaneous and thermodynamically favorable, while a positive Gibbs free energy value suggests that the reaction is non-spontaneous and requires energy input to occur. By measuring the Gibbs free energy changes of various biochemical reactions in maize, a better understanding of the metabolic pathways involved in processes such as photosynthesis, respiration, and carbohydrate metabolism could be achieved. As can be seen from Table 1 and particularly from Table 2, the regression between the content of sugars against differential entropy vs. enthalpy is inversely proportional, especially with respect to the monitored parameters in plumulas and radicles of the seedlings of both maize genotypes. This relationship is very reminiscent of the enthalpy–entropy compensatory dependence, already observed in related biological material in the works of [49,53], suggesting that the enthalpy–entropy relationship may influence the redistribution of assimilates in the maize seedlings examined and thus the changes in Gibbs free energy and the growth of seedlings.
The presented results are significant (Table 1 and Table 2), since there is a possibility of maintaining the stability of these metastable structures of living matter through the controlled modulation of system entropy (by the inverse relationship of the entropy and enthalpy changes) [30]. These results indicate that the changes in Gibbs free energy, in particular seedling organs (plumule (P), radicle (R), rest of seed (RoS/S)), in both corn hybrids and its linear dependence on monitored sugars are not statistically significant, contrary to the significant relationship obtained for whole young plants (Table 1 and Table 2). The observed correlations suggest that the regulation of these processes occurs at the plant seedling level. This indicates that different plant growth regulation processes (mainly regulated by different phytohormones) and/or source–sink relationships are important in the redistribution of assimilates and therefore in the regulation of plant growth and development in maize seedlings. This can be explained in more detail by examining the regression relationship between Gibbs free energy and the content of monitored sugars in whole young plants and individual organs (P, R, RoS/S).
The parameters of the PLS model for whole seedlings of ZP 434 maize hybrid (conducted using 30 samples of parts of a young plant: rest of seed + plumule + radicle) were statistically significant, with relatively high values of R2cal and R2CV and a low difference between the RMSEC and RMSECV values (Table 1, Figure 4).

3.5. The multilinear Regression Model for Maize Seedlings

The plot of the measured versus the predicted ΔGo values indicates grouping of samples of P and R on one side of the regression line and samples of the RoS/S on the opposite end (Figure 4A). Additionally, in a group of RoS/S samples, the control samples are distinguished from the treated samples. Figure 4A is proof of source–sink relationships, such as between maize plumules and radicles, as young developing organs and net importers of assimilates (sink organs) and RoS/S, as a net exporter of assimilates (source organ) [64]. For the ZP 434 hybrid model, it is observed that the medium values of the exogenously added 24-EBL had the greatest influence on the source–sink ratios, which coincides with the results of the shoot length in seedlings (Figure 2A). It can be observed that the values of the ΔGo parameter of thermodynamics for the plumule and radicle samples have lower values compared to the RoS/S samples, which indicates that the transfer of energy (and matter) takes place from RoS/S to the plumule and the radicle. This observation is understandable considering the fact that the RoS/S in maize seedlings is the source organ, i.e., the origin of the net production of assimilates (mainly mono- and oligosaccharides) produced by the degradation of the reserve starch, while the plumule and radicle are sink organs, i.e., they consume or are net “importers” of assimilates from the RoS/S [18,64]. Furthermore, the onset of grass seed germination is characterized by the activation of α-Amylase enzyme synthesis (crucial in the process of starch degradation during grass seed germination, which is again controlled by the production of GA in the aleurone layer and scutellum of grass seedlings) [64]. In addition, a positive effect of the examined medium concentrations of 24-EBL on the source-sink relationship for the ZP434 maize hybrid can be seen in Figure 1 and Figure 4A, and based on the results, it can be assumed that the source–sink relations could be driven by BRs.
Glucose has the highest positive impact on ΔGo changes in the whole ZP434 hybrid seedling, followed by arabinose and trehalose (Figure 4B). Glucose is a major substrate of mitochondrial respiration, which, in addition to ATP production, produces the organic acids necessary for transamination during nitrogen fixation [65]. Furthermore, glucose is the major monomer of the cell wall polysaccharide [23] but also one of the constituents of sucrose, the major transport sugar in the vast majority of plants [22]. Bear in mind that BR phytohormones affect some respiration patterns [6] but also the dynamics of cell wall construction [24], which is not surprising because the enzyme XET is induced among other factors by the action of BRs, which explains the observed relationship of the Glu and ΔGo thermodynamic parameter. A positive effect of arabinose on the change in the ΔGo parameter of thermodynamics would be similarly interpreted, since arabinose is one of the important monomers in the arabinogalactan polymers of hemicellulose, an important building block of the cell wall [24].
Furthermore, a positive effect of trehalose on the changes in the ΔGo parameter of whole ZP 434 hybrid seedlings is also observed (Figure 4B). Trehalose has recently been found to be of great importance in plant metabolism, although present in a much smaller amount than other sugars [26]. The significance of this sugar so far stems from its exceptional properties (maybe due to its participation in the formation of glassy state, possibly through the water replacement mechanism [27]), which contribute to the protection of plant proteins under stress [46]. Thus, trehalose complements the effect of other sugars and LEA proteins that are considered as crucial in the formation of glassy state, which strongly contributes to the stability of macromolecular structures in dry seeds [27], but this disaccharide also appears to contribute to their stability during the rehydration and germination of seeds [29]. It is probable that the primary processes affecting the changes in the ΔGo parameter of thermodynamics in whole seedlings of ZP 434 maize hybrid are related to cell wall remodeling, increased respiration of the young plant, and protective processes related to trehalose. The redistribution of assimilates (associated with the sugars fructose and sucrose) does not appear to have the positive effect of changing the ΔGo parameter of thermodynamics in the seedlings. In addition, it is notable that the PLS model of the radicle of seedlings for the ZP 434 maize hybrid is good because of the high values of R2cal (=0.851748) and R2CV (=0.777918) and small differences between the parameters RMSEC (=0.0357414) and RMSECV (=0.0454265), which may indicate different metabolism in these two organs of the ZP 434 hybrids but also of different “sink strength” and different affinity of these organs to assimilates [66,67].
As stated, glucose plays a variety of roles in the seedlings of different plants [23,65]. Since there is a positive effect of arabinose on ΔGo, it is possible that cell wall remodeling processes are of great importance during seed germination and seedling establishment in the ZP 704 maize hybrid.
The statistical parameters of the multilinear regression model for the whole seedlings of the ZP 704 maize hybrid are not the best due to the relatively low values of R2cal (=0.807494) and R2CV (=0.711153), and the difference between the parameters RMSEC (=0.529099) and RMSECV (=0.5680484) is not insignificant. Although the results for the entire seedlings of ZP704 maize hybrid (Figure 5A) are more scattered relative to equivalent data for whole seedlings of the ZP434 maize hybrid (Figure 4A), it is observed (Figure 5A) that the values of the ΔGo parameter of thermodynamics of the ZP 704 hybrid plumule and radicle are lower or the same for the RoS/S of the ZP704 maize hybrid, a trend that is observed for the whole seedlings of the ZP 434 maize hybrids (Figure 4A). This could be associated with the source–sink relationship between plumules and radicles as a sink organ (both young developing organs and net importers of assimilates) and RoS/S as a source organ (net exporter of assimilates) [64].
It is obvious that the changes in the thermodynamic parameter ΔGo (as a measure of the biosynthetic capacity of the system) depend on the source–sink relationships, whereby ΔGo flows from the RoS/S to the plumule and radicle (Figure 4A and Figure 5A). For the ZP 704 model, it is observed that the lower values of the exogenously added 24-EBL have the greatest influence on the source–sink ratios, which is in accordance with the results of the shoot length in seedlings (Figure 2A). Although the vast majority of scientific papers about source–sink relationships are related to fully formed plants, it is clear that due to the simple hydrostatic rules on the transport of matter (and energy) through phloem [68], similar rules can apply to seedlings. Moreover, it has been observed that source–sink transitions of young tissues are determined by the presence of sucrose synthase and invertase enzymes on target cells [67], which leads to the degradation of sucrose as the major transport sugar to its hexose components and thus to the formation of a concentration gradient, which directs the phloem flow to the sink organs. Without going into the question of whether a phloem system has been developed in 7-day-old maize seedlings or whether sugars from the RoS/S are transported to the plumule and the radicle in a simple manner, these processes, and thus the overall source-sink relationships, are highly influenced by plant hormones [69], affecting sucrose-degrading enzymes.
The results (Table 1 and Table 2) indicate the importance of sugar redistribution in different organs (P, R, RoS/S) of maize seedlings and their effect on ΔGo changes within a metabolic network whose activation depends on the genotype, the concentration of exogenously added 24-EBL, and the type of seedling organ. It was mentioned that this diversity depends on the phenomenon of sink strength [66,67]. In addition, the effects of exogenously added 24-EBL on these processes in maize seedlings depend on the presence of BR plasmalemma receptors, the BRI proteins, which can vary in different tissues (root) [70]. Although these phytohormones are thought to be produced in the plant meristem tissue [71], since it is a young plant with very pronounced meristem tissue additionally exposed to the exogenous treatment of 24-EBL, the described experiment corresponds to the concentration of BR-type phytohormones in a wide variety of plant tissues and species [2].
The canalization transport of auxins through the plant, which greatly influences the various processes of plant growth and development [72] and related phenomena of apical dominance and phyllotaxis [73], is the reasons why appropriate cybernetic solutions have been proposed [74]. However, other phytohormones have been reported to be transported through the plants and partially modulate the aforementioned effects of auxin [73]. BRs are one of them [74], which agrees with the intention of Hartwig and Wang [75] to present a “molecular circuit of steroid signaling in plants”. Our contribution to this problem consisted of taking into consideration both the energetics and thermodynamics of the plant, with a presented solution to the problem of source–sink relationships in plant seedlings. The presented study convincingly demonstrates that it is also possible to describe the process of plant seed germination quantitatively by using thermodynamic parameters.

4. Conclusions

The results show that the different concentrations of 24-EBL exhibit varying effects on seed germination. It was found that the highest concentrations of 24-EBL had the greatest impact on seed germination (inhibitory), while lower and medium concentrations have a positive effect on the early stages of seedling development. The enthalpy and entropy correlation affects the redistribution of assimilates in the examined maize seedlings. In addition, the changes in Gibbs free energy influence the growth of seedlings. The changes in the thermodynamic parameter ΔGo (as a measure of the biosynthetic capacity of the system) depend on the source–sink relationships, whereby ΔGo flows from the RoS/S to the plumule and radicle. This study confirms that the flow of important sugars from the RoS/S to the plumule and radicle, which are the sink organs, as well as the changes in Gibbs free energy between these parts are responsible for controlling and regulating the germination and early growth of the maize seedlings. This study suggests that this approach could be used to measure the “germination energy” quantitatively (as ΔGo presents a measure of the chemical potential of living systems), which was previously evaluated qualitatively (as a percent of germinated seeds) in seed science.

Author Contributions

Conceptualization, B.N. and H.W.; methodology, B.B., H.W. and B.N.; validation, B.N., B.B., B.J., J.T., H.W., S.K., M.M. and V.D.; formal analysis, J.T. and B.N.; resources, H.W. and B.J.; writing—original draft preparation, B.N., B.B. and H.W.; writing—review and editing, B.J., M.M., S.K., J.T. and V.D.; visualization, B.J. and J.T.; supervision, V.D. and B.J.; project administration, H.W. and V.D.; funding acquisition, H.W., J.T. and B.N. All authors have read and agreed to the published version of the manuscript.

Funding

Authors would like to acknowledge the financial support of the Ministry of Science, Technological Development and Innovation of the Republic of Serbia, under Contract numbers 451-03-47/2023-01/200017 (“Vinča” Institute of Nuclear Sciences—National Institute of the Republic of Serbia), 451-03-47/2023-01/200051 (Institute of General and Physical Chemistry), and 451-03-47/2023-01 (The Institute for Plant Protection and Environment)), 451-03-47/2023-01/200168 (University of Belgrade—Faculty of Chemistry).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mandava, N.B. Plant growth promoting brassinosteroids. Ann. Rev. Plant Physiol. Plant Mol. Biol. 1988, 39, 23–52. [Google Scholar] [CrossRef]
  2. Li, S.; Zheng, H.; Lin, L.; Wang, F.; Sui, N. Roles of brassinosteroids in plant growth and abiotic stress response. Plant Growth Regul. 2021, 93, 29–38. [Google Scholar] [CrossRef]
  3. Bajguz, A.; Hayat, S. Effects of brassinosteroids on the plant responses to environmental stresses. Plant Physiol. Biochem. 2009, 47, 1–8. [Google Scholar] [CrossRef] [PubMed]
  4. Clouse, S.D. Brassinosteroid Signal Transduction: From Receptor Kinase Activation to Transcriptional Networks Regulating Plant Development. Plant Cell 2011, 23, 1219–1230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Derevyanchuk, M.V.; Grabelnyh, O.I.; Litvinovskaya, R.P.; Voinikov, V.K.; Sauchuk, A.L.; Khripach, V.; Kravets, V. Influence of brassinosteroids on plant cell alternative respiration pathway and antioxidant systems activity under abiotic stress conditions. Biopolym. Cell 2014, 30, 436–442. [Google Scholar] [CrossRef] [Green Version]
  6. Groszyk, J.; Szechyńska-Hebda, M. Effects of 24-Epibrassinolide, Bikinin, and Brassinazole on Barley Growth under Salinity Stress Are Genotype- and Dose-Dependent. Agronomy 2021, 11, 259. [Google Scholar] [CrossRef]
  7. Ahmad, A.; Tola, E.; Alshahrani, T.S.; Seleiman, M.F. Enhancement of Morphological and Physiological Performance of Zea mays L. under Saline Stress Using ZnO Nanoparticles and 24-Epibrassinolide Seed Priming. Agronomy 2023, 13, 771. [Google Scholar] [CrossRef]
  8. Kim, T.-W.; Wang, Z.-Y. Brassinosteroid Signal Transduction from Receptor Kinases to Transcription Factors. Annu. Rev. Plant Biol. 2010, 61, 681–704. [Google Scholar] [CrossRef] [Green Version]
  9. Pokotylo, I.V.; Kretynin, S.V.; Khripach, V.A.; Ruelland, E.; Blume, Y.B.; Kravets, V.S. Influence of 24-epibrassinolide on lipid signaling and metabolism in Brassica napus. Plant Growth Regul. 2014, 73, 9–17. [Google Scholar] [CrossRef]
  10. Gururani, M.A.; Mohanta, T.K.; Bae, H. Current Understanding of the Interplay between Phytohormones and Photosynthesis under Environmental Stress. Int. J. Mol. Sci. 2015, 16, 19055–19085. [Google Scholar] [CrossRef] [Green Version]
  11. Gururani, M.A.; Ganesan, M.; Song, P.-S. Photo-biotechnology as a tool to improve agronomic traits in crops. Biotechnol. Adv. 2015, 33, 53–63. [Google Scholar] [CrossRef] [PubMed]
  12. Marthandan, V.; Geetha, R.; Kumutha, K.; Renganathan, V.G.; Karthikeyan, A.; Ramalingam, J. Seed Priming: A Feasible Strategy to Enhance Drought Tolerance in Crop Plants. Int. J. Mol. Sci. 2020, 21, 8258. [Google Scholar] [CrossRef] [PubMed]
  13. Alhammad, B.A.; Ahmad, A.; Seleiman, M.F.; Tola, E. Seed Priming with Nanoparticles and 24-Epibrassinolide Improved Seed Germination and Enzymatic Performance of Zea mays L. in Salt-Stressed Soil. Plants 2023, 12, 690. [Google Scholar] [CrossRef] [PubMed]
  14. Khan, I.; Awan, S.A.; Ikram, R.; Rizwan, M.; Akhtar, N.; Yasmin, H.; Sayyed, R.Z.; Ali, S.; Ilyas, N. Effects of 24-epibrassinolide on plant growth, antioxidants defense system, and endogenous hormones in two wheat varieties under drought stress. Physiol. Plant. 2021, 172, 696–706. [Google Scholar] [CrossRef] [PubMed]
  15. Xiong, M.; Yu, J.; Wang, J.; Gao, Q.; Huang, L.; Chen, C.; Zhang, C.; Fan, X.; Zhao, D.; Liu, Q.Q.; et al. Brassinosteroids regulate rice seed germination through the BZR1-RAmy3D transcriptional module. Plant Physiol. 2022, 189, 402–418. [Google Scholar] [CrossRef]
  16. Fincher, G.B. Molecular and cellular biology associated with endosperm mobilization in germinating cereal grains. Ann. Rev. Plant Physiol. Plant Mol. Biol. 1989, 40, 305–346. [Google Scholar] [CrossRef]
  17. Komor, E. Source physiology and assimilate transport: The interaction of sucrose metabolism, starch storage and phloem export in source leaves and the effects on sugar status in plants. Aust. J. Plant Physiol. 2000, 27, 497–505. [Google Scholar] [CrossRef]
  18. Roland, F.; Baena-Gonzalez, E.; Sheen, J. Sugar Sensing and Signaling in Plants: Conserved and Novel Mechanisms. Annu. Rev. Plant Biol. 2006, 57, 675–709. [Google Scholar] [CrossRef] [Green Version]
  19. Neto, J.-R.C.F.; de Melo Souza, A.C.; da Silva, M.D.; Benko-Iseppon, A.M.; Pandolfi, V.; da Costa, A.F.; Kido, E.K. The Transcriptional Modulation of Inositol and Raffinose Family Oligosaccharides Pathways in Plants—An (A)biotic Stress, Perspective. In Abiotic and Biotic Stress in Plants: Recent Advances and Future Perspectives; Shanker, A.K., Shanker, C., Eds.; IntechOpen: London, UK, 2016; Volume 4, pp. 81–99. [Google Scholar]
  20. Carpita, N. Structure and biogenesis of the cell walls of grasses. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1996, 47, 445–476. [Google Scholar] [CrossRef]
  21. Wolf, S.; Mravec, J.; Greiner, S.; Mouille, G.; Höfte, H. Plant Cell Wall Homeostasis Is Mediated by Brassinosteroid Feedback Signaling. Curr. Biol. 2012, 22, 1732–1737. [Google Scholar] [CrossRef] [Green Version]
  22. Kosar, F.; Akram, N.A.; Sadiq, M.; Al-Qurainy, F.; Ashraf, M. Trehalose: A key organic osmolyte effectively involved in plant abiotic stress tolerance. J. Plant Growth Reg. 2019, 38, 606–618. [Google Scholar] [CrossRef]
  23. Sarkar, A.K.; Sadhukhan, S. Imperative role of trehalose metabolism and trehalose-6-phosphate signalling on salt stress responses in plants. Physiol. Plant. 2022, 174, e13647. [Google Scholar] [CrossRef] [PubMed]
  24. Iordachescu, M.; Imai, R. Trehalose and Abiotic Stress in Biological Systems. In Abiotic Stress in Plants—Mechanims and Adaptation; Shanker, A., Ed.; InTech: Rijeka, Croatia, 2011; Volume 10, pp. 215–234. [Google Scholar]
  25. Li, J.; Nam, K.H. Regulation of brassinosteroid signaling by a GSK3/SHAGGY-like kinase. Science 2002, 295, 1299–1301. [Google Scholar] [CrossRef] [PubMed]
  26. Sun, W.Q. Methods for the Study of Water Relations Under Desiccation Stress. In Desiccation and Survival in Plants. Drying Without Dying; Black, M., Pritchard, H.W., Eds.; CABI: Walingford, UK, 2002; Volume 2, pp. 47–91. [Google Scholar]
  27. Kranner, I.; Minibayeva, F.V.; Beckett, R.P.; Seal, C.E. What is stress? Concepts, definitions and applications in seed science. New Phytol. 2010, 188, 655–673. [Google Scholar] [CrossRef] [PubMed]
  28. Flock, T.; Weatheritt, R.J.; Latysheva, N.S.; Babu, M.M. Controlling entropy to tune the functions of intrinsically disordered regions. Curr. Opin. Struct. Biol. 2014, 26, 62–72. [Google Scholar] [CrossRef]
  29. Popović, M.; Minceva, M. Standard Thermodynamic Properties, Biosynthesis rates, and the Driving Force of Growth of Five Agricultural Plants. Front. Plant Sci. 2021, 12, 671868. [Google Scholar] [CrossRef]
  30. Dragicevic, V.; Sredojevic, S. Thermodynamics of Seed and Plant Growth. In Thermodynamics—Systems in Equilibrium and Non-Equilibrium; InTech: Rijeka, Croatia, 2011. [Google Scholar] [CrossRef] [Green Version]
  31. Waisi, H.; Nikolic, B.; Jankovic, B. Transformation of Matter and Energy in Crops Under the Influence of Brassinosteroids. In Brassinosteroids: Plant Growth and Development; Hayat, S., Yusuf, M., Bhardwaj, R., Bajguz, A., Eds.; Springer: Singapore, 2019. [Google Scholar] [CrossRef]
  32. Sandler, S. Chemical and Engineering Thermodinamics, 3rd ed.; JohnWiley & Sons: Hoboken, NJ, USA, 1989. [Google Scholar]
  33. Pincus, S.M.; Singer, B.H. Randomness and degrees of irregularity. Proc. Natl. Acad. Sci. USA 1996, 93, 2083–2088. [Google Scholar] [CrossRef] [Green Version]
  34. Lindenmayer, A. Developmental Algorithms for Multieellular Organisms: A Survey of L-Systemst. J. Theor. Biol. 1975, 54, 3–22. [Google Scholar] [CrossRef] [Green Version]
  35. Prusinkiewicz, P.; Lindenmayer, A. The Algorithmic Beauty of Plants; Springer: New York, NY, USA, 1990; p. 228. [Google Scholar]
  36. Prusinkiewicz, P.; Runions, A. Computational models of plant development and form. New Phytol. 2012, 193, 549–569. [Google Scholar] [CrossRef]
  37. Prusinkiewicz, P.; Cieslak, M.; Ferraro, P.; Hanan, J. Modeling plant development with L-systems. In Mathematical Modeling in Plant Biology; Morris, P., Ed.; Springer: New York, NY, USA, 2018; Volume 8, pp. 139–169. [Google Scholar]
  38. Allen, M.T.; Prusinkiewicz, P.; DeJong, T.M. Using L-systems for modeling source-sink interactions, architecture and physiology of growing trees: The L-PEACH model. New Phytol. 2005, 166, 869–880. [Google Scholar] [CrossRef] [Green Version]
  39. Graeff, M.; Rana, S.; Marhava, P.; Moret, B.; Hardtke, C.S. Local and Systemic Effects of Brassinosteroid Perception in Developing Phloem. Curr. Biol. 2020, 30, 1626–1638.e3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Santoro, M.V.; Nievas, F.; Zygadlo, J.; Giordano, W.; Banchio, E. Effects of growth regulators on biomass and the production of secondary metabolites in Peppermint (Mentha piperita) micropropagated in vitro. Am. J. Plant Sci. 2013, 4, 49–55. [Google Scholar] [CrossRef] [Green Version]
  41. Ozga, J.; Dennis, M. Hormonal interactions in fruit development. J. Plant Growth Regul. 2003, 22, 73–81. [Google Scholar] [CrossRef]
  42. Divi, U.K.; Krishna, P. Brassinosteroid: A Biotechnological Target for Enhancing Crop Yield and Stress Tolerance. New Biotechnol. 2009, 26, 131–136. [Google Scholar] [CrossRef]
  43. Deng, Z.; Zhang, X.; Tang, W.; Oses-Prieto, J.A.; Suzuki, N.; Gendron, J.M.; Chen, H.; Guan, S.; Chalkley, R.J.; Peterman, T.K.; et al. A proteomics study of brassinosteroid response in Arabidopsis. Mol. Cell. Proteom. 2007, 6, 2058–2071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Li, Z. Brassinosteroids regulate plant growth, development, and stress responses via members of the GSK3/shaggy-like kinase family. Mol. Plant 2019, 12, 659–671. [Google Scholar]
  45. Steber, C.M.; McCourt, P. A role for brassinosteroids in germination in Arabidopsis. Plant Physiol. 2001, 125, 763–769. [Google Scholar] [CrossRef] [Green Version]
  46. Manghwar, H.; Hussain, A.; Ali, Q.; Liu, F. Brassinosteroids (BRs) role in plant development and coping with different stresses. Int. J. Mol. Sci. 2022, 23, 1012. [Google Scholar] [CrossRef]
  47. Waisi, H.; Janković, B.; Nikolić, B.; Dragičević, V.; Panić, I.; Tosti, T.; Trifković, J. Influence of various concentrations of 24-epibrassinolide on the kinetic parameters during isothermal dehydration of two maize hybrids. S. Afr. J. Bot. 2018, 119, 69–79. [Google Scholar] [CrossRef]
  48. Vidya Vardhini, B.; Seeta Ram Rao, S. Amelioration of osmotic stress by brassinosteroids on seed germination and seedling growth of three varieties of sorghum. Plant Growth Regul. 2003, 41, 25–31. [Google Scholar] [CrossRef]
  49. ISTA, International Seed Testing Association. Seed Testing International (1996). International rules for seed testing. Seed Sci. Technol. 1996, 21, 1–288. [Google Scholar]
  50. Waisi, H.K.; Petković, A.Z.; Nikolić, B.R.; Janković, B.Ž.; Raičević, V.B.; Lalević, B.T.; Giba, Z.S. Influence of 24-epibrassinolide on seedling growth and distribution of mineral elements in two maize hybrids. Chem. Ind. 2017, 71, 201–209. [Google Scholar] [CrossRef] [Green Version]
  51. Waisi, H. The Influence of Brassinosteroid 24-Epibrassinolide on Germination and Early Stages of Growth and Development of Different Maize Hybrids (Zea mays L.). Ph.D. Thesis, University of Belgrade, Faculty of Biology, Belgrade, Serbia, 2016. [Google Scholar]
  52. Matusmoto, T.; Yamada, K.; Yoshizawa, Y.; Oh, K. Comparison of effect of brassinosteroid and gibberellin biosynthesis inhibitors on growth of rice seedlings. Rice Sci. 2016, 23, 51–55. [Google Scholar] [CrossRef] [Green Version]
  53. Zhang, W.; Zhao, J.; Li, X.; Dai, H.; Lei, J. Seed morphology and germination of native Tulipa species. Agriculture 2023, 13, 466. [Google Scholar] [CrossRef]
  54. Gupta, R.; Chakrabarty, S.K. Gibberellic acid in plant: Still a mystery unresolved. Plant Signal. Behav. 2013, 8, e25504. [Google Scholar] [CrossRef] [Green Version]
  55. Olszewski, N.; Sun, T.P.; Gubler, F. Gibberellin signaling: Biosynthesis, catabolism, and response pathways. Plant Cell 2002, 14 (Suppl. S1), S61–S80. [Google Scholar] [CrossRef] [Green Version]
  56. Piazza, F.; Colella, M.; Cinelli, G.; Lopez, F.; Donati, I.; Sacco, P. Effect of α-Amylase on the Structure of Chia Seed Mucilage. Biomimetics 2022, 7, 141. [Google Scholar] [CrossRef]
  57. Yadav, R.K.; Devi, L.L.; Singh, A.P. Brassinosteroids in plant growth and development. In Plant Hormones in Crop Improvement; Academic Press: Cambridge, MA, USA, 2023; pp. 185–203. [Google Scholar]
  58. Bajguz, A.; Orczyk, W.; Gołębiewska, A.; Chmur, M.; Piotrowska-Niczyporuk, A. Occurrence of brassinosteroids and influence of 24-epibrassinolide with brassinazole on their content in the leaves and roots of Hordeum vulgare L. cv. Golden Promise. Planta 2019, 249, 123–137. [Google Scholar] [CrossRef] [Green Version]
  59. Clouse, S.D.; Sasse, J.M. Brassinosteroids: Essential regulators of plant growth and development. Annu. Rev. Plant Biol. 1998, 49, 427–451. [Google Scholar] [CrossRef] [Green Version]
  60. Majda, M.; Robert, S. The role of auxin in cell wall expansion. Int. J. Mol. Sci. 2018, 19, 951. [Google Scholar] [CrossRef] [Green Version]
  61. Waisi, H.; Kosović, A.; Krstić, Đ.; Milojković-Opsenica, D.; Nikolić, B.; Dragićević, V.; Trifković, J. Polyphenolic profile of maize seedlings treated with 24-epibrassinolide. J. Chem. 2015, 2015, 976971. [Google Scholar] [CrossRef] [Green Version]
  62. Buitink, J.; Hoekstra, F.A.; Leprince, O. Biochemistry and Biophysics of Tolerance system. In Desiccation and Survival in Plants. Drying Without Dying; Black, M., Pritchard, H.W., Eds.; CABI: Walingford, UK, 2002; Volume 10, pp. 293–318. [Google Scholar]
  63. Noctor, G.; Foyer, C.H. A re-evaluation of the ATP: NADPH budget dur-ing C3 photosynthesis. A contribution from nitrate assimilation and its associated respiratory activity? J. Exp. Bot. 1998, 49, 1895–1908. [Google Scholar] [CrossRef]
  64. Ho, L.C. Metabolism and Compartmentation of Imported Sugars in Sink Organs in Relation to Sink Strength. Ann. Rev. Plant Physiol. Plant Mol. Biol. 1988, 39, 355–378. [Google Scholar] [CrossRef]
  65. Turgeon, R. The Sink-Source Transition in Leaves. Ann. Rev. Plant Physiol. Plant Mol. Biol. 1989, 40, 119–138. [Google Scholar] [CrossRef]
  66. Kumar, R.; Bishop, E.; Bridges, W.C.; Tharayil, N.; Sekhon, R.S. Sugar partitioning and source–sink interaction are key determinants of leaf senescence in maize. Plant Cell Environ. 2019, 42, 2597–2611. [Google Scholar] [CrossRef]
  67. Ren, H.; Qi, H.; Zhao, M.; Zhou, W.; Wang, X.; Gong, X.; Jiang, Y.; Li, C. Characterization of source–sink traits and carbon translocation in maize hybrids under high plant density. Agronomy 2022, 12, 961. [Google Scholar] [CrossRef]
  68. Van Esse, G.W.; van Mourik, S.; Stigler, H.; ten Hove, C.A.; Molenaar, J.; de Vries, S.C. A Mathematical Model fo BRASSINOSTEROID INSENSITIVE1-Mediated Signaling in Root Growth and Hypocotil Elongation. Plant Physiol. 2012, 160, 523–532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Symons, G.M.; Ross, J.J.; Jager, C.E.; Reid, J.B. Brassinosteroid transport. J. Exp. Bot. 2008, 59, 17–24. [Google Scholar] [CrossRef] [Green Version]
  70. Leyser, O. Auxin, Self-Organisation, and the Colonial Nature of Plants. Curr. Biol. 2011, 21, R331–R337. [Google Scholar] [CrossRef] [Green Version]
  71. Reinhardt, D.; Pesce, E.R.; Stieger, P.; Mandel, T.; Baltensperger, K.; Bennet, M.; Traas, J.; Friml, J.; Kuhlmeier, C. Regulation of phyllotaxis by polar auxin transport. Nature 2003, 426, 255–260. [Google Scholar] [CrossRef]
  72. Smith, R.S.; Guyomarc’h, S.; Mandel, T.; Reinhardt, D.; Kuhlmeier, C.; Prusinkiewic, P. A plausible model of phyllotaxis. Proc. Natl. Acad. Sci. USA 2006, 103, 1301–1306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Crawford, S.; Shinoharabinose, N.; Sieberer, T.; Williamson, L.; George, G.; Hepworth, J.; Müller, D.; Domagalska, M.A.; Leyser, O. Strigolactones enhance competition between shoot branches by dampening auxin transport. Development 2010, 137, 2905–2913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Sankar, M.; Osmont, K.S.; Rolcik, J.; Gujas, B.; Tarkowska, D.; Strnad, M.; Xenarios, I.; Hardtke, C.S. A qualitative continuous model of cellular auxin and brassinosteroid signaling and their crosstalk. Bioinformatics 2011, 27, 1404–1412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Hartwig, T.; Wang, Z.-Y. The molecular circuit of steroid signaling in plants. Essays Biochem. 2015, 58, 71–82. [Google Scholar] [PubMed]
Figure 1. Effect of different concentrations of 24-EBL on germination of 7-day-old seedlings of ZP 434 and ZP 704 maize hybrids. Values indicated by the same letter were not statistically different (p < 0.05).
Figure 1. Effect of different concentrations of 24-EBL on germination of 7-day-old seedlings of ZP 434 and ZP 704 maize hybrids. Values indicated by the same letter were not statistically different (p < 0.05).
Agronomy 13 01673 g001
Figure 2. Effect of different concentrations of 24-EBL on morphometric parameters of 7-day-old seedlings of ZP 434 and ZP 704 maize hybrids. (A) Shoot length; (B) root length. Values indicated by the same letter were not statistically different (p < 0.05).
Figure 2. Effect of different concentrations of 24-EBL on morphometric parameters of 7-day-old seedlings of ZP 434 and ZP 704 maize hybrids. (A) Shoot length; (B) root length. Values indicated by the same letter were not statistically different (p < 0.05).
Agronomy 13 01673 g002
Figure 3. Effect of different concentrations of 24-EBL on total redox status of whole seedlings of ZP 704 and ZP 434 maize hybrids, determined using the EPR spectroscopy method. The results are given in arbitrary units of the EPR signal, double integral values per sample mass. The axis values are displayed in ten thousands. Values indicated by the same letter were not statistically different (p < 0.05).
Figure 3. Effect of different concentrations of 24-EBL on total redox status of whole seedlings of ZP 704 and ZP 434 maize hybrids, determined using the EPR spectroscopy method. The results are given in arbitrary units of the EPR signal, double integral values per sample mass. The axis values are displayed in ten thousands. Values indicated by the same letter were not statistically different (p < 0.05).
Agronomy 13 01673 g003
Figure 4. (A) The plot of the measured versus the predicted ΔGo values obtained from PLS model for whole 7-day-old corn seedlings of the ZP 434 hybrid. “P” represents plumule, “R” represents radicle, and “S” represents rest of seed. K refers to control samples. Numbers from 1K to 30 represent concentrations of exogenously added 24-EBL to seedlings, from higher to lower concentrations, respectively; (B) plot of the coefficients of descriptors (sugars: trehalose, arabinose, glucose, fructose, sucrose, raffinose) in PLS model of the whole 7-day-old maize seedlings of the ZP434 hybrid.
Figure 4. (A) The plot of the measured versus the predicted ΔGo values obtained from PLS model for whole 7-day-old corn seedlings of the ZP 434 hybrid. “P” represents plumule, “R” represents radicle, and “S” represents rest of seed. K refers to control samples. Numbers from 1K to 30 represent concentrations of exogenously added 24-EBL to seedlings, from higher to lower concentrations, respectively; (B) plot of the coefficients of descriptors (sugars: trehalose, arabinose, glucose, fructose, sucrose, raffinose) in PLS model of the whole 7-day-old maize seedlings of the ZP434 hybrid.
Agronomy 13 01673 g004
Figure 5. (A) The plot of the measured versus the predicted ΔGo values obtained from PLS model for whole 7-day-old corn seedlings of the ZP704 maize hybrid. “P” represents plumule, “R” represents radicle, and “S” represents rest of seed. K refers to control samples. Numbers from 1K to 30 represent concentrations of exogenously added 24-EBL to seedlings, from higher to lower concentrations, respectively; (B) plot of the coefficients of descriptors (sugars: trehalose, arabinose, glucose, fructose, sucrose, raffinose) in PLS model of whole 7-day-old maize seedlings of the ZP 704 hybrid.
Figure 5. (A) The plot of the measured versus the predicted ΔGo values obtained from PLS model for whole 7-day-old corn seedlings of the ZP704 maize hybrid. “P” represents plumule, “R” represents radicle, and “S” represents rest of seed. K refers to control samples. Numbers from 1K to 30 represent concentrations of exogenously added 24-EBL to seedlings, from higher to lower concentrations, respectively; (B) plot of the coefficients of descriptors (sugars: trehalose, arabinose, glucose, fructose, sucrose, raffinose) in PLS model of whole 7-day-old maize seedlings of the ZP 704 hybrid.
Agronomy 13 01673 g005
Table 1. Statistical parameters of the models of regression relationship between the content of selected sugars (Tre, Ara, Glu, Fru, Sah, Raf) and Gibbs free energy assessed in 7-day-old maize seedlings (hybrid ZP 434).
Table 1. Statistical parameters of the models of regression relationship between the content of selected sugars (Tre, Ara, Glu, Fru, Sah, Raf) and Gibbs free energy assessed in 7-day-old maize seedlings (hybrid ZP 434).
Thermodynamic ParameterSeedling PartsModel ParametersRegression Coefficients in Model *
EntropyPlumule (P)Low statistical significance model
Radicle (R)RMSEC: 0.0020
RMSECV: 0.0046
R2Cal: 0.7561
R2CV: 0.2554
Tre (+), Sah (+), Ara (+)
Glu (−), Fru (−)
Raf (0)
Rest of seed (RoS/S)Low statistical significance model
Whole seedlingLow statistical significance model
EnthalpyPlumule (P)Low statistical significance model
Radicle (R)RMSEC: 0.3711
RMSECV: 0.8215
R2Cal: 0.7563
R2CV: 0.2823
Fru (+), Glu (+)
Ara (−), Sah (−), Tre (−)
Raf (0)
Rest of seed (RoS/S)Low statistical significance model
Whole seedlingLow statistical significance model
Gibbs Free EnergyPlumule (P)RMSEC: 0.0152
RMSECV: 0.0539
R2Cal: 0.9714
R2CV: 0.8479
Raf (+), Tre (+), Fru (+), Glu (+)
Ara (−), Sah (−)
Radicle (R)RMSEC: 0.0357
RMSECV: 0.0454
R2Cal: 0.8517
R2CV: 0.7779
Glu (+), Fru (+), Tre (+), Sah (+), Raf (+), Ara (+)
Rest of seed (RoS/S)RMSEC: 0.1866
RMSECV: 0.2947
R2Cal: 0.6471
R2CV: 0.3319
Sah (+), Raf (+)
Ara (−), Glu (−), Fru (−), Tre (−)
Whole seedlingRMSEC: 0.5123
RMSECV: 0.5931
R2Cal: 0.7188
R2CV: 0.6328
Glu (+), Ara (+), Tre (+)
Raf (−), Sah (−), Fru (−)
* + (positive influence on the dependent variable, in descending order); − (negative influence on the dependent variable, in descending order).
Table 2. Model parameters of regression relationship between the content of selected sugars (Tre, Ara, Glu, Fru, Sah, Raf) and differential thermodynamic parameters (entropy, enthalpy, and Gibbs free energy) assessed in 7-day-old maize seedlings (hybrid ZP 704).
Table 2. Model parameters of regression relationship between the content of selected sugars (Tre, Ara, Glu, Fru, Sah, Raf) and differential thermodynamic parameters (entropy, enthalpy, and Gibbs free energy) assessed in 7-day-old maize seedlings (hybrid ZP 704).
Thermodynamic ParameterSeedling PartsModel ParametersRegression Coefficients in Model *
EntropyPlumule (P)RMSEC: 0.0072
RMSECV: 0.0098
R2Cal: 0.7339
R2CV: 0.5425
Raf (+)
Glu (−), Sah (−), Ara (−), Tre (−)
Fru (0)
Radicle (R)RMSEC: 0.0087
RMSECV: 0.0097
R2Cal: 0.6509
R2CV: 0.6180
Raf (+), Fru (+), Glu (+), Sah (+)
Ara (−), Tre (−)
Rest of seed (RoS/S)RMSEC: 0.0077
RMSECV: 0.0102
R2Cal: 0.8687
R2CV: 0.8117
Glu (+), Ara (+), Sah (+), Raf (+), Fru (+)
Tre (−)
Whole seedlingLow statistical significance model
EnthalpyPlumule (P)RMSEC: 1.3083
RMSECV: 1.7839
R2Cal: 0.7340
R2CV: 0.5433
Tre (+), Ara (+), Sah (+), Glu (+),
Raf (−)
Fru (0)
Radicle (R)RMSEC: 1.5849
RMSECV: 1.7754
R2Cal: 0.6508
R2CV: 0.5632
Tre (+), Ara (+)
Sah (−), Glu (−), Fru (−), Raf (−)
Rest of seed (RoS/S)RMSEC: 1.4122
RMSECV: 2.0551
R2Cal: 0.8687
R2CV: 0.7782
Tre (+)
Fru (−), Sah (−), Raf (−), Glu (−), Ara (−)
Whole seedlingLow statistical significance model
Gibbs Free EnergyPlumule (P)RMSEC: 0.0709
RMSECV: 0.1692
R2Cal: 0.3456
R2CV: 0.4042
Tre (+), Raf (+), Fru (+), Ara (+), Glu (+)
Sah (−)
Radicle (R)Low statistical significance model
Rest of seed (RoS/S)RMSEC: 0.0904
RMSECV: 0.2579
R2Cal: 0.8282
R2CV: 0.4600
Sah (+), Ara (+), Fru (+), Tre (+)
Glu (−), Raf (−)
Whole seedlingRMSEC: 0.5291
RMSECV: 0.6804
R2Cal: 0.8075
R2CV: 0.7112
Glu (+), Ara (+),
Tre (−), Raf (−), Sah (−), Fru (−)
* + (positive influence on the dependent variable, in descending order); − (negative influence on the dependent variable, in descending order).
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

Božilović, B.; Nikolić, B.; Waisi, H.; Trifković, J.; Dodevski, V.; Janković, B.; Krstić, S.; Mojović, M. Influence of 24-Epibrassinolide on the Energetic Parameters and Early Stages of Growth and Development in Seedlings of Two Maize (Zea mays L.) Genotypes. Agronomy 2023, 13, 1673. https://doi.org/10.3390/agronomy13071673

AMA Style

Božilović B, Nikolić B, Waisi H, Trifković J, Dodevski V, Janković B, Krstić S, Mojović M. Influence of 24-Epibrassinolide on the Energetic Parameters and Early Stages of Growth and Development in Seedlings of Two Maize (Zea mays L.) Genotypes. Agronomy. 2023; 13(7):1673. https://doi.org/10.3390/agronomy13071673

Chicago/Turabian Style

Božilović, Bojana, Bogdan Nikolić, Hadi Waisi, Jelena Trifković, Vladimir Dodevski, Bojan Janković, Sanja Krstić, and Miloš Mojović. 2023. "Influence of 24-Epibrassinolide on the Energetic Parameters and Early Stages of Growth and Development in Seedlings of Two Maize (Zea mays L.) Genotypes" Agronomy 13, no. 7: 1673. https://doi.org/10.3390/agronomy13071673

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