Next Article in Journal
Pulmonary Fibrosis and Progressive Pulmonary Fibrosis in a Prospective Registry of Interstitial Lung Diseases in Eastern Siberia
Next Article in Special Issue
The Role of Heat Acclimation in Thermotolerance of Chickpea Cultivars: Changes in Photochemical and Biochemical Responses
Previous Article in Journal
Informative Censoring—A Cause of Bias in Estimating COVID-19 Mortality Using Hospital Data
Previous Article in Special Issue
Free Radicals Mediated Redox Signaling in Plant Stress Tolerance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulatory Mechanisms of Plant Growth-Promoting Rhizobacteria and Plant Nutrition against Abiotic Stresses in Brassicaceae Family

by
Arshad Jalal
1,
Carlos Eduardo da Silva Oliveira
1,
Fernando Shintate Galindo
2,
Poliana Aparecida Leonel Rosa
1,
Isabela Martins Bueno Gato
1,
Bruno Horschut de Lima
1 and
Marcelo Carvalho Minhoto Teixeira Filho
1,*
1
Department of Plant Health, Rural Engineering, and Soils, Campus of Ilha Solteira, São Paulo State University (UNESP), Av. Brasil, 56- Centro, Ilha Solteira 15385-000, SP, Brazil
2
Faculty of Agricultural and Technological Sciences, Campus of Dracena, São Paulo State University (UNESP), Dracena 17900-000, SP, Brazil
*
Author to whom correspondence should be addressed.
Life 2023, 13(1), 211; https://doi.org/10.3390/life13010211
Submission received: 23 November 2022 / Revised: 29 December 2022 / Accepted: 5 January 2023 / Published: 11 January 2023
(This article belongs to the Special Issue Plant Biotic and Abiotic Stresses)

Abstract

:
Extreme environmental conditions, such as abiotic stresses (drought, salinity, heat, chilling and intense light), offer great opportunities to study how different microorganisms and plant nutrition can influence plant growth and development. The intervention of biological agents such as plant growth-promoting rhizobacteria (PGPRs) coupled with proper plant nutrition can improve the agricultural importance of different plant species. Brassicaceae (Cruciferae) belongs to the monophyletic taxon and consists of around 338 genera and 3709 species worldwide. Brassicaceae is composed of several important species of economical, ornamental and food crops (vegetables, cooking oils, forage, condiments and industrial species). Sustainable production of Brassicas plants has been compromised over the years due to several abiotic stresses and the unbalanced utilization of chemical fertilizers and uncertified chemicals that ultimately affect the environment and human health. This chapter summarized the influence of PGPRs and nutrient management in the Brassicaceae family against abiotic stresses. The use of PGPRs contributed to combating climate-induced change/abiotic factors such as drought, soil and water salinization and heavy metal contamination that limits the general performance of plants. Brassica is widely utilized as an oil and vegetable crop and is harshly affected by abiotic stresses. Therefore, the use of PGPRs along with proper mineral nutrients management is a possible strategy to cope with abiotic stresses by improving biochemical, physiological and growth attributes and the production of brassica in an eco-friendly environment.

1. Introduction

Brassica is one of the most important and economical vegetables of the Brassicaceae family [1] and includes several species (Brassica oleracea, Brassica rapa, Nasturtium officinale, Raphanus sativus, Diplotaxis tenuifolia and Eruca vesicaria), containing secondary metabolites and beneficial contents of putative health-promoting compounds [2]. Brassicaceae are a rich source of primary and secondary metabolites (amino acids, sugars, indoles, phenolics and glucosinolates) that help in the production of antioxidants [3,4] to promote tolerance to biotic and abiotic stresses [5]. Brassicaceae are emergently adapting as a research model crop in plant science due to their interaction with biotic and abiotic stresses as their high defensive mechanisms and a series of alterations in metabolites allow them to survive under climatic extremes [6]. Therefore, proper management practices are needed when encountering extreme environmental conditions (drought, salinity, temperature, heavy metals and nutrients deficiency) and to ensure optimal plant growth and productivity [7].
Abiotic stresses disturb plant physiology and metabolism, which leads to the reduction of plant growth and productivity [8]. The growth, yield and quality of Brassica grown in arid and semi-arid areas were extremely affected by drought conditions [9]. In addition, nutrient limitation is another vulnerable condition that alters plant growth, production and quality. Plants adapt different physiological and biochemical functions to adjust to extreme challenges and avoid injuries under abiotic stresses [10]. Macronutrients mobilize and assimilate along with organic compounds that could improve plant growth and development and mitigate plant abiotic stresses [11]. The absorption of chromium (Cr), zinc (Zn), iron (Fe) and manganese (Mn) was increased with chelating agents of low molecular weight, which led to the improvement of oil content in Brassica juncea up to 35% [12]. The imbalanced utilization of macro and micronutrients may cause metal toxicity in several crop plants [13]. However, Brassica species deal with the hyper-accumulation of these nutrients by improving biochemical processes and the mobilization of nutrients through the roots–shoot system [14]. In addition, the root rhizosphere is influenced by different biotic and abiotic factors including soil and root type and plant species and age. Hence, plant growth-promoting rhizobacteria (PGPR) are classified into several groups on the basis of their capacities and taxonomical status. These bacteria activate several mechanisms that alter soil organic matter to an instantly available form [15], as well as the regularization and transformation of soluble sugars, proline, amino acids and mineral nutrients in the soil above plant parts, thus improving nutrient accumulation in nutrient-deficient soils [16].
The plant and bacteria association promotes nutrient uptake and assimilation, which favors the plants’ tolerance to biotic and abiotic stresses [17]. Plants and microbial communities are the components of similar limited resources with a different relationship. However, plants assist microbial communities with available nutrients from the soil rhizosphere [18] and improve nitrogen mineralization, which can enhance the uptake of other nutrients for a higher performance and yield of plants [19]. The positive association (symbiosis) and negative association (pathogenesis) of the plant rhizosphere microbial community can affect nutrient availability and resource partition, thus increasing or reducing crop production, respectively [18,20]. The positive association of the microbial community increases their activities in the rhizosphere of host plants, which can improve the soil organic matter (SOM) content and nutritional status of the plant [21]. Beneficial bacteria are the first soil-borne communities that alter and re-adjust in stressful environments for their survival; however, their activities and configurations are the first affected factors under stress [22]. The plant growth-promoting rhizobacteria community is vulnerable to stressful conditions of low water potential and nutrient availability that may be reflected in the form of physiological stress in the plants [23].
The eco-physiological and functional activities of nutrients and PGPRs need proper attention and extensive research to improve plant tolerance to abiotic stresses. Therefore, this review highlighted the interaction between plant growth-promoting rhizobacteria and mineral nutrition and their influence on the tolerance to abiotic stresses in the Brassicas plant species.

2. Adverse Effects of Abiotic Stress in Plants

Abiotic stresses are the foremost confining factors for agricultural productivity. Crop plants overcome the drastic external pressure of intrinsic mechanisms caused by environmental and edaphic conditions that affect the growth, development and productivity of plants [24,25]. The sustainable production of vegetables such as Brassicas around the world has been compromised due to several harsh environmental conditions and the unbalanced use of synthetic fertilizers and uncertified chemicals over the years that affect the environment and human health and led to inadequate climatic conditions. Abiotic stresses consist of drought, low/high temperature, salinity, light intensity, flooding, heavy metals toxicity and nutrient starvation. The extensive use of chemicals, macro and micronutrients, non-essential elements and radionuclides are the main sources of metal toxicity in soil [13,25]. Brassicaceae are capable plant species that deal with the hyper-accumulation of heavy metals through their biochemical expression, acquisition and re-mobilization in roots [13,14]. Waterlogging/flooding is an excess of soil water that can reduce oxygen availability in plant root systems and thus negatively affect crop growth and yield [26]. Flooding has negatively affected lipid biosynthesis and the yield of several rapeseed varieties [27].
Cold stress is associated with chilly weather (0–15 °C) and frosty weather (<0 °C) that leads to the disturbance of the photosynthetic process and reduces the primary production of B. oleracea [28]. Cold stress impairs metabolic and enzymatic activities that can disrupt the cell membrane and cause seed rotting in Brassica plants [29,30]. Light radiation (low or high) affects plant morphology and the root–shoot ratio [31]. Exposure of broccoli (B. oleracea) to ultraviolet (UV) light can increase ascorbic acid [32,33]. High light causes photoinhibition of the photosystem and protein degradation in B. rapa plants [34]. In short, abiotic stresses alter several internal functions of plants by disturbing homeostasis, physio-biochemical and molecular attributes, such as water and nutrient use efficiency and assimilation, osmotic adjustment, disruption of membrane integrity and enzymatic activities, as well as reduction in photosynthetic efficiency [29,31,34]. The abiotic stresses and their consequences are summarized in Figure 1.

3. Use of Plant Growth-Promoting Rhizobacteria to Mitigate Adverse Effects of Abiotic Stress

In recent years, the contribution of rhizosphere microorganisms to increasing plant growth and crop productivity as well as tolerance to biotic and abiotic stresses without causing pathogenicity have been discussed in the literature [35]. Several genera of plant growth-promoting rhizobacteria (PGPR) including Azospirillum, Bacillus, Rhizobium, Pseudomonas and Bradyrhizobium showed positive interactions with different vegetables species [36,37]. Several previous studies highlighted the capacity of different PGPRs in biological nitrogen fixation (N2) [38,39], increasing the availability of iron (Fe) [40], phosphorus (P) and zinc (Zn) solubilization and transportation [41,42]. The PGPRs also improved the performance and growth of plants through the production of phytohormones such as gibberellins, ethylene, cytokinin, auxins and salicylic acid [43,44].
The use of PGPRs has contributed to combating climate-induced changes (abiotic factors) such as uneven rainfall (drought), soil and water salinization and heavy metal contamination that limit the general performance of plants [44,45]. These microorganisms improve soil fertility and structure, which contribute to a successful adaptation of the plant under stressful conditions [45]. Researchers have been focused on the use of these microorganisms with emphasis on bacteria of the genera Azospirillum, Bacillus, Pseudomonas, Rhizobium, Bradyrhizobium, Herbaspirillum and Burkholderia [36,38].
PGPRs exist in the rhizosphere and tissues of plants, which may adapt multiple mechanisms including the synthesis and exudation of phytohormones (indole-3-acetic acid (AIA)), cytokinin, ethylene and gibberellins [46]; synthesis of plant growth-regulators including nitric oxide [47]; abscisic acid [48]; polyamines such as spermidine and spermine [49]; increase solubilization and availability of nutrients [50,51]; increase nitrate reductase activity and nutrient use efficiency [38,52]; biocontrol of phytopathogens and diseases [53]; and protection of plants against water and saline stress and toxic chemical elements of the soil [54]. In addition to assisting in biological nitrogen fixation, PGPRs have the ability to enhance cell membrane stability of the leaf and reduce the rate of leaf abscission during drought stress conditions [55]. Several PGPRs improve the tolerance capability of plants by producing certain phytohormones [56] that can be used for heavy metal remediation, mobilization or immobilization from soil into plant tissues [57,58]. These microbes also utilized 1-aminocyclopropane-1-carboxylic acid (ACC) to prevent ethylene production [59] and mitigate stresses by endophytic biota, which were caused due to high radiation and light stress [60]. Plant growth-promoting rhizobacteria adapted several mechanisms to improve the growth and development of the plants of the Brassicaceae family under abiotic stresses (Figure 2).
Plant growth-promoting rhizobacteria (PGPR) promote plant tolerance to abiotic stresses through the adaptation of several mechanisms as well as down- or up-regulating stress genes [61]. The inoculation of rapeseed plants with Pseudomonas sp. and Azospirillum sp. mitigate salt stress [62] by increasing the solubilization and availability of macro- and micronutrients for better uptake in the above-ground part of the host plant [63,64]. PGPRs prominently improved root–shoot fresh and dry weights, leaf area, chlorophyll and several growth-promoting hormones, which ultimately improved the seedling growth of B. oleracea and B. napus [65,66]. Flooding is another abiotic stress that harshly reduces antioxidant activities; however, inoculation with bio-fertilizers (Azotobacter chroococcum, Azospirillum spp. and Pseudomonas spp. and Azospirillum spp., Pseudomonas fluorescens and Basillus subtilis) via seeds and foliar efficiently alleviate flooding affects in canola by increasing growth and yield [67]. In this context, the supply of these rhizobacteria or PGPRs to plants of the Brassicaceae family brought benefits to their cultivation in abiotic conditions (Table 1).

4. Plant Nutrition to Mitigate Adverse Effects of Abiotic Stress on Brassicas

Plants develop extensive adaptive and/or resistance mechanisms to sustain productivity and survival under stressful conditions. However, adequate nutrient application is an imperative tool to meet the Sustainable Development Goals to attain food and nutritious security and promote sustainable productivity under climate extremes [79]. Optimization of nutrient content (macro- and micronutrients, secondary nutrients and heavy metals) in soil and plant systems have been reported to enhance crop adaptation to resilience conditions, as these are structural elements of several co-factors and enzymes. Nutrients assist structures’ stability of protein and alleviate reactive oxygen species (ROS) production. The versatility of nutrient application under severe environmental conditions has significantly improved the yield and quality traits of various crops [80].
Fertilizers are considered the most important and crucial inputs to achieve greater crop growth and production in modern agriculture [81]. Plants require NPK and other essential micronutrients such as iron (Fe), copper (Cu), zinc (Zn), manganese (Mn), molybdenum (Mo), nickel (Ni), chlorine (Cl) and boron (B) in very small quantities for better performance and yield. These elements are collectively considered as essential for humans and animals and their deficiency can affect their metabolic, physical and mental development. Macro- and micronutrients play a critical role in the effectiveness of several biological compounds and enzymes for the proper functioning of different metabolic processes. The relevance of macronutrients’ essentiality for higher yield and nutritional status has been increasing over several decades [82]. Ensuring that plants are well-fed with essential nutrients is a cost-effective strategy with the capacity to mitigate abiotic stresses and enhance productivity [79,81]. The effect of macro- and micronutrients on different functions of Brassicaceae crops promotes plant growth and increases tolerance to abiotic stresses (Figure 3).

4.1. Macronutrients

Macronutrients are considered to be significant drivers for enhancing the yield and quality parameters of crop plants. Traditional fertilizer application in a field may not fulfill the demands of individual plants while over and/or under application causes soil quality degradation, groundwater pollution and reduction in productivity. Leaf nutrition of rapeseeds is an important factor to optimize fertilization and productivity, alongside contributing to commercial and environmental profits [83]. Better management of macronutrient fertilizers can improve plant growth and yield under stressful conditions. The nutrients and their functions in the crop plants are discussed below in detail.

4.1.1. Nitrogen

Nitrogen (N) is the most needed nutrient for most cultivated plants, and it directly affects plant development and yield [84,85,86,87,88,89]. Nitrogen is the main constituent of the atmosphere, but its availability is still one of the main limiting factors for the productivity of terrestrial ecosystems including agro-ecosystems [90]. Nitrogen plays an important role in plant nutrition and development [87], such as the synthesis and production of phytohormones, co-enzymes, nucleic acids, secondary metabolites, chlorophyll and proteins content [91].
Several studies have reported that N fertilization promoted different species of Brassicaceae including oilseed producer crops such as rapeseed (B. napus) [92,93], brown mustard (B. juncea) [94,95] and turnip rape (B. rapa) [96] and horticultural crops such as radish (Raphanus sativus) [97], cauliflower (B. oleracea L. var. botrytis) [98,99], cabbage (B. oleracea L. var. Capitata) [100,101], broccoli (B. oleracea L. (var. italica) [102,103], kale (B. oleracea L. var. sabellica) [104,105] and arugula (Eruca vesicaria subsp. Sativa) [106].
Abiotic stress conditions alter the N metabolism of Brassicaceae plants [94], negatively affecting N uptake and assimilation, N use efficiency (NUE), photosynthetic capacity and plant growth [107], particularly under prolonged (24 h) stress exposure [108]. The interaction of N fertilization and abiotic conditions plays an important role in determining the potential of plant development and abiotic stress tolerance. Stress relief depends on the type of N fertilization; applying ammonium (NH4+) to plants resulted in a stronger tolerance to heat stress as compared to the fertilization with nitrate (NO3) [109]. In addition, N fertilization can compensate for the negative effects of abiotic conditions by facilitating carbon partitioning, cell membrane stability, osmoregulation and antioxidative mechanisms that could improve plant growth and development as well reduce leaf senescence under extreme environmental conditions [110].

4.1.2. Phosphorus

Phosphorus (P) is a primary macronutrient with a structural function in plants. It is involved in drivers of metabolic functions including respiration, energy storage and transportation, production of nucleic acid, membrane stability, catalyze enzymes activities, redox reactions and contribution to carbohydrate metabolism [111]. As with other plant families, P is one of the important nutrients for the Brassicaceae family that directly affects its development and productivity [112]. Holzschuh et al. [113] studied different doses of P fertilizer in Brassicas and reported that the species of this family are highly demanding of P availability in the soil, especially broccoli (B. oleracea var. itálica) and cauliflower (B. oleracea var. botrytis). The optimal management of P fertilization in vegetables is essential for their proper growth, development and yield [112]. Phosphorus deficiency in soil and plants directly affects vegetable vigor, establishment and root development, thus disrupting water use efficiency [114]. Several plants of the Brassicas species have the capability to tolerate and respond to various types of stresses through hormonal stimulation, ion exchange, antioxidant enzymes and the activation of signaling flow in their metabolic and genetic boundaries that mitigate stressed conditions [115].
Application and management of appropriate P fertilization has increased water use efficiency against drought stress [116,117]. Jones et al. [118] indicated that adequate soil P contents compensate for the impact of drought stress on the growth and yield of plants. Application of P source fertilizers may reduce the drastic effects of water scarcity during pollen formation or the reproductive stage that could increase flower and pod production, resulting in a greater yield and high protein content in grains [119]. Phosphate fertilizers improved the performance of B. juncea under salt stress by increasing plant dry mass and P uptake while lowering the Na+/K+ ratio [114]. Phosphorus fertilization adapts different mechanisms that immobilize the metal content in soil [120] by reducing their dissolution under the low pH range of soil, hence leading to the reduction of the bioavailability and uptake of metals by plants [121]. Phosphate fertilization increases the pH of soil solution to constrain absorption of heavy metals, as their availability decreases with increasing P fertilization [122].

4.1.3. Potassium

Plants develop a wide range of adaptive and resistive strategies that sustain productivity and survival under stressful conditions. Plant tissues may adjust osmotic potential through the absorption of various compatible osmolytes such as inorganic ions, carbohydrates, organic acids and free amino acids [123,124]. Plants adjust osmotic potential by regulating stomatal conductance, photosynthesis, leaf turgidity and plant growth rate under drought, salt and high-temperature stresses through potassium (K) osmolytes [125]. Potassium is one of the major inorganic osmolytes that enable osmotic regulation and adjustment during stress conditions. Potassium ion absorption protects plants from harmful impacts of different stresses including drought, salinity, metal toxicity and high or cold temperatures by osmotic adjustment and maintenance of stomatal conductance, protecting cell integrity and increasing photosynthesis as well as via the detoxification of reactive oxygen species [123].
In addition, K is a crucial element for the distribution of photo-assimilates in root systems [126] that protects plants against most abiotic stresses including metal toxicity such as Cd-induced oxidative damage [127], Zn toxicity [128], NaCl toxicity [129], drought stress [130] and high radiance incidence [131]. Potassium supplementation increases the adjustment of stomata, which regulates carbohydrate formation and the growth of Nicotiana rustica during stress conditions [125]. Samar-Raza et al. [132] reported that application of K fertilizer under drought stress enhanced the tolerance of wheat (Triticum aestivum L.) by reducing toxic elements’ absorption and enhancing physiological efficiency and yield [133].

4.1.4. Calcium

Calcium (Ca) plays a vital role in the physiological functions of plants and acts as a second messenger element of external signals for the higher performance of plants. It has an essential role in the structure and stabilization of the cell wall and membrane, regulating metabolic, enzymatic and hormonal processes [134]. The alteration in free cytosolic Ca2+ ion contents is validated during naturally occurring abiotic stimulants (low and high temperature/light, tensions, high osmotic and oxidative tensions, also during biotic stimulants (nodulation aspects and fungal drivers)) [135]. It also has an explicit function in the performance and maintenance of plant development and detoxification of heavy metals [136]. The main function of Ca+ ions under heavy metal stress is to maintain the activities of antioxidant enzymes, reducing the peroxidation of lipids in the cell membrane and improving the physio-biochemical processes of plants [127,137].

4.1.5. Magnesium

Magnesium (Mg2+) is an essential nutrient for plant growth [138], regulating cell membrane stability, carbon fixation, chlorophyll synthesis, carbohydrate transport, enzymatic activities and reproductive process [139,140,141]; thus, it helps plants to adapt defensive mechanisms against abiotic stresses [142]. Plants under Mg nutrition improve root growth and root surface area that increase water and nutrient uptake from the rhizosphere and enhance transportation of photo-assimilates and carbohydrate synthesis, which can mitigate drought-stress-induced deleterious changes [143]. Magnesium transports carbohydrates from roots to shoots and helps in the fixation of photosynthetic CO2 during the reproductive growth stage under salt stress. The efficiency of Mg foliar fertilization is right-away associated with the distribution of nutrients within plants [144]. Nutrient solution with Mg fertilization improved the shoot growth of B. rapa L. var. pervirdis under cadmium (Cd) toxicity [145].
Deficiency of Mg is one of the common nutritional syndromes in plants, which may have drastic impacts on agricultural productivity and quality [146] and lead to morphological and physiological abnormalities of plants [147]. Plants produce antioxidants and antioxidative defensive enzyme activities, especially ascorbic acid during the stress of Mg deficiency [148]. The glutathione-producing ascorbate-determined H2O2 scavengers are responsible for ascorbic acid that can enable the plants to detoxify ROS production to protect plants from climate extremes [149]. Glutathione homeostasis can be regulated through the over-production of glyoxalase genes that can help the plants to sustain Mg content during stressful conditions and increase tolerance to metalliferous soil [150,151].
Magnesium transporters are also involved in metal transport. Under the low Mg content, nickel (Ni+) is well-cited for the suppression of electron flow and impairing photosynthesis functions by replacing Mg2+ in chlorophyll fragments. Adequate fertilizer of Mg alleviates the Ni+ effect in the root rhizosphere that may reduce the negative probability of Ni at the outer surface of the plasma membrane by replacing the targeted ionic binding site [152]. The Mg transporter (AtMHX) from Arabidopsis acts as an H+ exchanger with Zn and Mg and is confined to the vacuole membrane [153]. The AtMGT1 protein derived from AtMGT (transporter gene of Arabidopsis) family in the plasma membrane exhibited greater attraction to Mg2+ ion, which helped in the re-distribution of Ni+, Ca2+, Fe3+, Mn2+ and Cu2+, when they are present in high concentrations [154].

4.1.6. Sulfur

Sulfur (S) is among the very active macronutrients in plant metabolism, which is why it is recognized alongside nitrogen (N) as a key nutrient for plant development [155]. Sulfur is used by plants to assimilate with a variety of organic compounds that are essential for the growth, development and mitigation of plant stress [11,156]. It is also responsible for making vegetables softer and adding greater commercial value [157]. Sulfur is predominantly found in the soil and is one of the main nutrients that is absorbed by plants in the form of sulfate anion (SO42−) from organic matter and a small proportion from the atmosphere in the form of sulfuric gas [158].
Kohlrabi (B. oleracea L.) is one of the crucially demanding S vegetables of the Brassicas family, which absorbs 1.5 kg S ton−1 of yield. Sulfur deficiency can inhibit leaf formation and change young leaves’ color from dark green to light green or yellowish. Proper S fertilization in kohlrabi (B. oleracea L.) improves tuber yield and reduces the undesirable nitrate content in consumable parts [159]. Canola (B. napus L. var. Oleifera) is also one of the most demanding S vegetables in reproductive phases as compared to other winter crops, as it exports a large amount of S to the grains [160]. Sulfur is one of the known nutrients that performs an imperative role in the tolerance to heavy metal toxicity [161].
Chromium is actively transported across the plasma membrane and appears to be mediated by transporters, which are primarily responsible for sulphate uptake [162,163]. This suggests the action of this molecule inhibit the absorption of heavy metals that are toxic to plants (Table 2).

4.2. Silicon

Silicon (Si) is the second most abundant chemical element after oxygen in the earth’s crust [199,200]; however, it is still not available directly to plants and is commonly adsorbed with oxides and silicates, affecting plant nutritional status [180,201,202,203]. In addition, the low dissolution of Si in the soil decreases its availability; thus, it occurs in a very low amount [204].
Plants uptake Si mainly from dissoluble mono-silicic acid (H2SiO4), a noncharged molecule which plays a significant role to increase plant resistance to abiotic and biotic conditions [205,206,207]. Silicon is distributed via xylem in the form of hydrated amorphous silica/silica bodies (SiO2.nH2O) and pledged to the epidermis of cell membrane. After deposition to the cell membrane, Si is no longer available for further distribution into the above-soil parts of the plants [208]. The transport of H4SiO4 occurs in a similar direction to transpiration (mass flow). Therefore, drought conditions increase the deposition of Si in the regions of leaf epidermis to protect water from high transpiration [209].
All soil-grown plants had Si constituents ranging from 0.1 to 10% of dry weight of plants [180,210]. However, Si is classified as a beneficial element, with it being an imperative element for several crops, specifically rice (Oryza sativa L.) and sugarcane (Saccharum officinarum L.). Moreover, its role has been well documented for the performance, growth and development of different Gramineae family crops [180,206,207,211,212,213]. This chemical element has been reported to be beneficial in mitigating abiotic stresses including heavy metal toxicity, salinity, high temperature, drought, radiation, aluminum toxicity, lodging, nutrient imbalance, wounding and freezing [214,215]. Rapeseed is one of the most studied plants of the Brassicaceae family regarding Si application to alleviate abiotic stress conditions [211], with the most common improvements reported in plant resistance to cold stress conditions, as well as the formation of larger seeds [216]. Table 2 summarizes the studies with Si fertilization in Brassicaceae plants under abiotic stress in the last decade (2004–2020).

4.3. Micronutrients

Micronutrients (zinc (Zn), iron (Fe), manganese (Mn), molybdenum (Mo), boron (B), copper (Cu) and chlorine (Cl)) improve plant health, water use efficiency, biomass production and provide systemic response against abiotic stresses [217,218,219]. Whereas plant growth-promoting rhizobacteria (PGPR) promote plant growth and tolerance to abiotic stresses by adapting and altering certain mechanisms, the production of ACC (1-aminocyclopropane-1-carboxylate) deaminase reduces ethylene synthesis, as well as alters phytohormones and antioxidative enzymes synthesis, and improves nutrient uptake [115,220].
Micronutrients may influence directly or indirectly the stress affecting plants due to their role in several enzymatic and metabolic activities [221]. Abiotic stress such as drought harshly impairs mineral nutrient translocation from soil to plant parts [222,223]. The Brassicaceae family is one of the most nutrient-demanding plant species, which is highly affected by inadequate nutrients application [224]. Therefore, deficiency of micronutrients disrupts the net-assimilation rate and stomatal conductance, electron transportation in photosynthesis, chlorophyll content, root–shoot ratio and antioxidant activities of cabbages, turnip and canola under abiotic stresses [225,226,227,228,229,230,231]. Salinity is a critical challenge to high production, physiological and biochemical attributes and nutrient uptake in Brassicaceae species [232]. Brassicas adapted certain mechanisms and variations, especially physiological variations to cope with salinity [233]. Salinization in plant systems can be ameliorated with foliar nutrient spray and rhizosphere micronutrient availability and uptake [234]. The accumulation of sodium (Na+) and chlorine (Cl) ions increases osmotic potential and decreases water availability and nutrient uptake through plant roots [235].
Several studies regarding the Brassicaceae family indicated that most of the species grown on contaminated soils with high accumulation of nutrients (Zn and Cu) and non-essential metals (Pd, Cd, Ni and Cr) [236,237,238,239]. Plants of B. juncea have the ability to accumulate high amounts of Cd, Cu, Ni, Cr, Zn, Fe, Co, Pb and Se from metal-contaminated sites [240,241,242]. Rapeseed subjected to early waterlogging stress resulted in higher accumulation of Mn, Fe, Zn and Cu in the leaves and caused toxicity [243]. Zinc is one of the efficient nutrients in the reduction of heat stroke by improving biochemical activities and superoxide dismutase (SOD) content in B. rapa [244,245]. Boron and Mn application in winter rapeseed (B. napus) positively influenced pod production, photosynthetic rate, N-metabolism, antioxidant activities and improved N and Ca contents in seeds [227,246]. High UV-B radiation may alter nutritional status, disturb plant cell metabolism, increase pathogens and disease tolerance [247], whereas light-emitting diodes (short duration blue light) enhanced phytochemical activities and micronutrient (Zn, Mn, Mo, B, Na, Fe and Cu) concentration in Broccoli (B. oleacea var. italica) [248,249] (Table 3).

5. Conclusions

Based on the updated literature, this review highlighted the importance of adequate and balanced nutrition against abiotic stresses in Brassicas species to ensure food and nutritious security. Proper management of macronutrients, micronutrients and silicon under certain conditions of abiotic stress could improve nutritional and physiological status, thus resulting in higher productivity and quality of Brassicas plants. Balanced application of macro- and micronutrients mitigates abiotic-stress-induced changes in Brassicas plant species by stimulating absorption and accumulation mechanisms for better survival.
The use of plant growth-promoting rhizobacteria (PGPR) has a critical role in combating climate-induced changes such as uneven rainfall (drought), soil and water salinization and heavy metal contamination, which limit the general performance of Brassicas plant species. Among the PGPRs, the genera Azospirillum, Bacillus, Pseudomonas, Herbaspirillum and Burkholderia are well studied for increasing plant nutrition, tolerance to pathogens and climate extreme conditions, and hence could improve plant performance and productivity in adverse growing conditions. Therefore, inoculation with PGPRs can increase productivity of Brassicas grown under abiotic stress conditions.
In the future, attention needs to be paid to the response of Mg and micronutrient application on crop resilience under different abiotic stresses. Dose-response management and multiple interactions of nutrients and heavy metals still need further investigation. Bio-fortification via foliar spray of micronutrients is a cost-effective strategy in alleviating global food and nutritious security which requires future advances and intensified research. The intervention of nano-fertilizers on the basis of integrated evidence is required to reduce the gap. The expansion of enhanced detection, tracking and monitoring strategies may be the best early detection technique for abiotic stresses which can also control yield losses and lethal impacts on the nutritional security of crops.

Author Contributions

Conceptualization, A.J. and M.C.M.T.F.; methodology, A.J. and C.E.d.S.O.; software, C.E.d.S.O. and F.S.G.; validation, A.J., F.S.G. and C.E.d.S.O.; formal analysis: A.J., I.M.B.G.; resources, M.C.M.T.F.; data curation, A.J., C.E.d.S.O., P.A.L.R. and B.H.d.L.; writing—original draft preparation, A.J. and F.S.G. writing—review and editing, M.C.M.T.F. and F.S.G.; visualization, B.H.d.L. and I.M.B.G.; supervision, M.C.M.T.F.; project administration, A.J. and M.C.M.T.F.; funding acquisition, A.J. and M.C.M.T.F. All authors have read and agreed to the published version of the manuscript.

Funding

This review received funding from The World Academy of Science (TWAS) and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), for the first author’s doctoral fellowship (CNPq/TWAS grant number: 166331/2018-0), and the productivity research grant (award number 311308/2020-1) of the corresponding author.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data availability is not applicable to this article.

Acknowledgments

The authors thank São Paulo State University (UNESP) for providing the technology and support as well as CNPq for the financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Essoh, A.P.; Monteiro, F.; Pena, A.R.; Pais, M.S.; Moura, M.; Romeiras, M.M. Exploring glucosinolates diversity in Brassicaceae: A genomic and chemical assessment for deciphering abiotic stress tolerance. Plant Physiol. Biochem. 2020, 150, 151–161. [Google Scholar] [CrossRef] [PubMed]
  2. Verkerk, R.; Knol, J.J.; Dekker, M. The effect of steaming on the glucosinolate content in broccoli. Acta Hortic. 2010, 867, 37–46. [Google Scholar] [CrossRef] [Green Version]
  3. Bruce, T.J.; Pickett, J.A. Plant defence signalling induced by biotic attacks. Curr. Opin. Plant Biol. 2007, 10, 387–392. [Google Scholar] [CrossRef]
  4. Jahangir, M.; Abdel-Farid, I.B.; Choi, Y.H.; Verpoorte, R. Metal ion-inducing metabolite accumulation in Brassica rapa. J. Plant Physiol. 2008, 165, 1429–1437. [Google Scholar] [CrossRef]
  5. Sudha, G.; Ravishankar, G.A. Involvement and interaction of various signaling compounds on the plant metabolic events during defense response, resistance to stress factors, formation of secondary metabolites and their molecular aspects. Plant Cell Tissue Organ Cult. 2002, 71, 181–212. [Google Scholar] [CrossRef]
  6. Jahangir, M.; Abdel-Farid, I.B.; Kim, H.K.; Choi, Y.H.; Verpoorte, R. Healthy and unhealthy plants: The effect of stress on the metabolism of Brassicaceae. Environ. Exp. Bot. 2009, 67, 23–33. [Google Scholar] [CrossRef]
  7. Andleeb, T.; Shah, T.; Nawaz, R.; Munir, I.; Munsif, F.; Jalal, A. QTL mapping for drought stress tolerance in plants. In Salt and Drought Stress Tolerance in Plants; Springer: Cham, Switzerland, 2020; pp. 383–403. [Google Scholar]
  8. Saharan, B.S.; Nehra, V. Plant growth promoting rhizobacteria: A critical review. Life Sci. Med. Res. 2011, 21, 30. [Google Scholar]
  9. Zhang, X.K.; Lu, G.Y.; Long, W.H.; Zou, X.L.; Li, F.; Nishio, T. Recent progress in drought and salt tolerance studies in Brassica crops. Breed. Sci. 2014, 64, 60–73. [Google Scholar] [CrossRef] [Green Version]
  10. Gong, Z.; Xiong, L.; Shi, H.; Yang, S.; Herrera-Estrella, L.R.; Xu, G.; Chao, D.Y.; Li, J.; Wang, P.Y.; Qin, F.; et al. Plant abiotic stress response and nutrient use efficiency. Sci. China Life Sci. 2020, 63, 635–674. [Google Scholar] [CrossRef]
  11. Kopriva, S.; Talukdar, D.; Takahashi, H.; Hell, R.; Sirko, A.; D’souza, S.F.; Talukdar, T. Frontiers of sulfur metabolism in plant growth, development, and stress response. Front. Plant. Sci. 2016, 6, 1–3. [Google Scholar] [CrossRef] [Green Version]
  12. Singh, S.; Sinha, S. Accumulation of metals and its effects in Brassica juncea (L.) Czern. (cv. Rohini) grown on various amendments of tannery waste. Ecotoxicol. Environ. Saf. 2005, 62, 118–127. [Google Scholar] [CrossRef] [PubMed]
  13. Mourato, M.P.; Moreira, I.N.; Leitão, I.; Pinto, F.R.; Sales, J.R.; Martins, L.L. Effect of Heavy Metals in Plants of the Genus Brassica. Int. J. Mol. Sci. 2015, 16, 17975–17998. [Google Scholar] [CrossRef] [Green Version]
  14. Sen, A.; Shukla, K.K.; Singh, S.; Tejovathi, G. Impact of heavy Metals on Root and Shoot Length of Indian Mustard: An Initial Approach for Phytoremediation. Sci. Secure J. Biotechol. 2013, 2, 48–55. [Google Scholar]
  15. Rakshapal, S.; Sumit, K.S.; Rajendra, P.P.; Alok, K. Technology for improving essential oil yield of Ocimum basilicum L. (sweet basil) by application of bioinoculant colonized seeds under organic field conditions. Indian Crop 2013, 45, 335–342. [Google Scholar] [CrossRef]
  16. Kumar Singh, A.; Cabral, C.; Kumar, R.; Ganguly, R.; Kumar Rana, H.; Gupta, A.; Pandey, A.K. Beneficial effects of dietary polyphenols on gut microbiota and strategies to improve delivery efficiency. Nutrients 2019, 11, 2216. [Google Scholar] [CrossRef] [Green Version]
  17. Compant, S.; Samad, A.; Faist, H.; Sessitsch, A. review on the plant microbiome: Ecology, functions, and emerging trends in microbial application. J. Adv. Res. 2019, 19, 29–37. [Google Scholar] [CrossRef]
  18. Classen, A.T.; Sundqvist, M.K.; Henning, J.A.; Newman, G.S.; Moore, J.A.M.; Cregger, M.A.; Moorhead, L.C.; Patterson, C.M. Direct and indirect effects of climate change on soil microbial and soil microbial-plant interactions: What lies ahead? Ecosphere 2015, 6, 1–21. [Google Scholar] [CrossRef]
  19. Galindo, F.S.; Pagliari, P.H.; Fernandes, G.C.; Rodrigues, W.L.; Boleta, E.H.M.; Jalal, A.; Céu, E.G.O.; Lima, B.H.D.; Lavres, J.; Teixeira Filho, M.C.M. Improving Sustainable Field-Grown Wheat Production with Azospirillum brasilense Under Tropical Conditions: A Potential Tool for Improving Nitrogen Management. Front. Environ. Sci. 2022, 10, 821628. [Google Scholar] [CrossRef]
  20. Jalal, A.; Oliveira, C.E.D.S.; Fernandes, H.B.; Galindo, F.S.; Silva, E.C.D.; Fernandes, G.C.; Nogueira, T.A.R.; De Carvalho, P.H.G.; Balbino, V.R.; Lima, B.H.D.; et al. Diazotrophic Bacteria Is an Alternative Strategy for Increasing Grain Biofortification, Yield and Zinc Use Efficiency of Maize. Plants 2022, 11, 1125. [Google Scholar] [CrossRef]
  21. Shahzad, T.; Chenu, C.; Genet, P.; Barot, S.; Perveen, N.; Mougin, C.; Fontaine, S. Contribution of exudates, arbuscular mycorrhizal fungi and litter depositions to the rhizosphere priming effect induced by grassland species. Soil Biol. Biochem. 2015, 80, 146–155. [Google Scholar] [CrossRef]
  22. Castro, H.F.; Classen, A.T.; Austin, E.E.; Norby, R.J.; Schadt, C.W. Soil microbial community responses to multiple experimental climate change drivers. Appl. Environ. Microbiol. 2010, 76, 999–1007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Manzoni, S.; Schimel, J.P.; Porporato, A. Responses of soil microbial communities to water stress: Results from a meta-analysis. Ecology 2012, 93, 930–938. [Google Scholar] [CrossRef] [PubMed]
  24. Dushenkov, S. Trends in phytoremediation of radionuclides. Plant Soil 2003, 249, 167–175. [Google Scholar] [CrossRef]
  25. Wuana, R.A.; Okieimen, F.E. Heavy metals in contaminated soils: A review of sources, chemistry, risks and best available strategies for remediation. Int. Sch. Res. Not. 2011, 2011, 402647. [Google Scholar] [CrossRef] [Green Version]
  26. Nabloussi, A.; Bahri, H.; Lakbir, M.; Moukane, H.; Kajji, A.; El Fechtali, M. Assessment of a set of rapeseeds (Brassica napus L.) varieties under waterlogging stress at different plant growth stages. OCL 2019, 26, 1–11. [Google Scholar] [CrossRef] [Green Version]
  27. Xu, M.; Ma, H.; Zeng, L.; Cheng, Y.; Lu, G.; Xu, J.; Zhang, X.; Zou, X. The effect of waterlogging on yield and seed quality at the early flowering stage in Brassica napus L. Field Crops Res. 2015, 180, 238–245. [Google Scholar] [CrossRef]
  28. Rodríguez, V.M.; Soengas, P.; Alonso-Villaverde, V.; Sotelo, T.; Cartea, M.E.; Velasco, P. Effect of temperature stress on the early vegetative development of Brassica oleracea L. BMC Plant Biol. 2015, 15, 145. [Google Scholar] [CrossRef] [Green Version]
  29. Tippmann, H.F.; Schlüter, U.; Collinge, D.B. Common themes in biotic and abiotic stress signalling in plants. In Floriculture, Ornamental and Plant Biotechnology; Tippmann, H.F., Schlüter, U., Collinge, D.B., Eds.; Global Science Books: Ikenobe, Japan, 2006; pp. 52–67. [Google Scholar]
  30. John, R.; Anjum, N.A.; Sopory, S.K.; Akram, N.A.; Ashraf, M. Some key physiological and molecular processes of cold acclimation. Biol. Plant 2016, 60, 603–618. [Google Scholar] [CrossRef]
  31. Ostonen, I.; Puttsepp, U.; Biel, C.; Alberton, O.; Bakker, M.R.; Lohmus, K.; Majdi, H.; Metcalfe, D.; Olsthoorn AF, M.; Pronk, A.; et al. Specific root length as an indicator of environmental change. Plant Biosyst. 2007, 141, 426–442. [Google Scholar] [CrossRef]
  32. Lemoine, M.L.; Civello, P.M.; Martínez, G.A.; Chaves, A.R. Influence of postharvest UV-C treatment on refrigerated storage of minimally processed broccoli (Brassica oleracea var. Italica). J. Sci. Food Agric. 2007, 87, 1132–1139. [Google Scholar] [CrossRef]
  33. Schonhof, I.; Klaring, H.P.; Krumbein, A.; Claussen, W.; Schreiner, M. Effect of temperature increase under low radiation conditions on phytochemicals and ascorbic acid in greenhouse grown broccoli. Agric. Ecosyst. Environ. 2007, 119, 103–111. [Google Scholar] [CrossRef]
  34. Jiao, S.; Emmanuel, H.; Guikema, J.A. High light stress inducing photoinhibition and protein degradation of photosystem I in Brassica rapa. Plant Sci. 2004, 167, 733–741. [Google Scholar] [CrossRef]
  35. Salvo LP, D.; Ferrando, L.; Fernandéz-Scavino, A.; Salamone IE, G. Microorganisms reveal what plants do not: Wheat growth and rhizosphere microbial communities after Azospirillum brasilense inoculation and nitrogen fertilization under field conditions. Plant Soil 2018, 424, 405–417. [Google Scholar] [CrossRef]
  36. Zeffa, D.M.; Fatin, L.H.; Santos, O.J.A.P.; Oliveira, A.L.M.; Canteri, M.G.; Scapim, C.A.; Gonçalves, L.S.A. The influence of topdressing nitrogen on Azospirillum spp. inoculation in maize crops through meta-analysis. Bragantia 2018, 77, 493–500. [Google Scholar] [CrossRef]
  37. Jalal, A.; Galindo, F.S.; Boleta, E.H.M.; Oliveira, C.E.d.S.; Reis, A.R.d.; Nogueira, T.A.R.; Moretti Neto, M.J.; Mortinho, E.S.; Fernandes, G.C.; Teixeira Filho, M.C.M. Common bean yield and zinc use efficiency in association with diazotrophic bacteria co-inoculations. Agronomy 2021, 11, 959. [Google Scholar] [CrossRef]
  38. Teixeira Filho, M.M.; Galindo, F.S. Inoculation of bacteria with a focus on biological nitrogen fixation and promotion of plant growth. In Topics in Soil Science—Volume X, 1st ed.; Severiano, E.C., Moraes, M.F., Paula, A.M., Eds.; Brazilian Society of Soil Science: Brasilia, Brazil, 2019; Chapter 11; pp. 577–648. [Google Scholar]
  39. Galindo, F.S.; Buzetti, S.; Rodrigues, W.L.; Boleta, E.H.M.; Silva, V.M.; Tavanti, R.F.R.; Fernandes, G.C.; Biagini, A.L.C.; Rosa, P.A.L.; Teixeira Filho, M.C.M. Inoculation of Azospirillum brasilense associated with silicon as a liming source to improve nitrogen fertilization in wheat crops. Sci. Rep. 2020, 10, 6160. [Google Scholar] [CrossRef] [Green Version]
  40. Sivasakthi, S.; Usharani, G.; Saranraj, P. Biocontrol potentiality of plant growth promoting bacteria (PGPR)—Pseudomonas fluorescens and Bacillus subtilis: A review. Afr. J. Agric. Res. 2014, 16, 1265–1277. [Google Scholar]
  41. Jalal, A.; da Silva Oliveira, C.E.; Freitas, L.A.; Galindo, F.S.; Lima, B.H.; Boleta, E.H.M.; Da Silva, E.C.; do Nascimento, V.; Nogueira, T.A.R.; Buzetti, S.; et al. Agronomic biofortification and productivity of wheat with soil zinc and diazotrophic bacteria in tropical savannah. Crop Pasture Sci. 2022, 73, 817–830. [Google Scholar] [CrossRef]
  42. Rosa, P.A.L.; Galindo, F.S.; Oliveira, C.E.D.S.; Jalal, A.; Mortinho, E.S.; Fernandes, G.C.; Marega, E.M.R.; Buzetti, S.; Teixeira Filho, M.C.M. Inoculation with plant growth-promoting bacteria to reduce phosphate fertilization requirement and enhance technological quality and yield of sugarcane. Microorganisms 2022, 10, 192. [Google Scholar] [CrossRef]
  43. Vejan, P.; Abdullah, R.; Khadiran, T.; Ismail, S.; Boyce, A.N. Role of plant growth promoting Rhizobacteria in agricultural sustainability—A review. Molecules 2016, 21, 573. [Google Scholar] [CrossRef] [Green Version]
  44. Zaheer, M.S.; Raza MA, S.; Saleem, M.F.; Khan, I.H.; Ahmad, S.; Iqbal, R.; Manevski, K. Investigating the effect of Azospirillum brasilense and Rhizobium pisi on agronomic traits of wheat (Triticum aestivum L.). Arch. Agron. Soil Sci. 2019, 65, 1554–1564. [Google Scholar] [CrossRef]
  45. Alor, E.T.; Dare, M.O.; Babalola, O.O. Microbial inoculants for soil quality and plant health. In Sustainable Agriculture Reviews; Springer: Cham, Switzerland, 2017; pp. 281–307. [Google Scholar] [CrossRef]
  46. Meza, B.; Bashan, L.E.; Bashan, Y. Involvement of indole-3-acetic acid produced by Azospirillum brasilense in accumulating intracellular ammonium in Chlorella vulgaris. Res. Microbiol. 2015, 166, 72–83. [Google Scholar] [CrossRef]
  47. Cohen, A.; Bottini, R.; Piccoli, P. Azospirillum brasilense Sp 245 produces ABA in chemically-defined culture medium and increases ABA content in arabidopsis plants. Plant Growth Regul. 2008, 54, 97–103. [Google Scholar] [CrossRef]
  48. Fibach-Paldi, S.; Burdman, S.; Okon, Y. Key physiological properties contributing to rhizosphere adaptation and plant growth promotion abilities of Azospirillum brasilense. FEMS Microbiol. Lett. 2012, 326, 99–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Cassán, F.; Maiale, S.; Masciarelli, O.; Vidal, A.; Luna, V.; Ruiz, O. Cadaverine production by Azospirillum brasilense and its possible role in plant growth promotion and osmotic stress mitigation. Eur. J. Soil Biol. 2009, 45, 12–19. [Google Scholar] [CrossRef]
  50. Rosa, P.A.L.; Mortinho, E.S.; Jalal, A.; Galindo, F.S.; Buzetti, S.; Fernandes, G.C.; Barco Neto, M.; Pavinato, P.S.; Teixeira Filho, M.C.M. Inoculation with growth-promoting bacteria associated with the reduction of phosphate fertilization in sugarcane. Front. Environ. Sci. 2020, 32. [Google Scholar] [CrossRef]
  51. Jalal, A.; Oliveira, C.E.D.S.; Bastos, A.D.C.; Fernandes, G.C.; de Lima, B.H.; Furlani Junior, E.; De Carvalho, P.H.G.; Galindo, F.S.; Gato, I.M.B.; Teixeira Filho, M.C.M. Nano-zinc and plant growth-promoting bacteria improve biochemical and metabolic attributes of maize in tropical Cerrado. Front. Plant Sci. 2023, 13, 5293. [Google Scholar]
  52. Galindo, F.S.; Rodrigues, W.L.; Fernandes, G.C.; Boleta, E.H.M.; Jalal, A.; Rosa, P.A.L.; Buzetti, S.; Lavres, J.; Teixeira Filho, M.C.M. Enhancing agronomic efficiency and maize grain yield with Azospirillum brasilense inoculation under Brazilian savannah conditions. Eur. J. Agron. 2022, 134, 126471. [Google Scholar] [CrossRef]
  53. Corrêa, E.B.; Bettiol, W.; Sutton, J.C. Biocontrol of root rot (Pythium aphanidermatum) and growth promotion with Pseudomonas chlororaphis 63-28 and Bacillus subtilis GB03 in hydroponic lettuce. Summa Phytopathol. 2010, 36, 275–281. [Google Scholar] [CrossRef]
  54. Cassán, B.; Diaz-Zorita, M. Azospirillum sp. in current agriculture: From the laboratory to the field. Soil Biol. Biochem. 2016, 103, 117–130. [Google Scholar] [CrossRef]
  55. Silva, E.R.; Zoz, J.; Oliveira CE, S.; Zuffo, A.M.; Steiner, F.; Zoz, T.; Vendruscolo, E.P. Can co-inoculation of Bradyrhizobium and Azospirillum alleviate adverse effects of drought stress on soybean (Glycine max L. Merrill.)? Arch. Microbiol. 2019, 201, 325–335. [Google Scholar] [CrossRef]
  56. de Souza, R.; Meyer, J.; Schoenfeld, R.; da Costa, P.B.; Passaglia LM, P. Characterization of plant growth-promoting bacteria associated with rice cropped in iron-stressed soils. Ann. Microbiol. 2015, 65, 951–964. [Google Scholar] [CrossRef]
  57. Mishra, J.; Singh, R.; Arora, N.K. Alleviation of heavy metal stress in plants and remediation of soil by rhizosphere microorganisms. Front. Microbiol. 2017, 8, 1706. [Google Scholar] [CrossRef] [PubMed]
  58. Banerjee, A.; Jhariya, M.K.; Yadav, D.K.; Raj, A. Micro-remediation of metals: A new frontier in bioremediation. In Handbook of Environmental Materials Management; Springer: Cham, Germany, 2018; pp. 1–36. [Google Scholar] [CrossRef]
  59. İpek, M.; Arıkan, Ş.; Pırlak, L.; Eşitken, A. Sustainability of Crop Production by PGPR Under Abiotic Stress Conditions. In Plant Growth Promoting Rhizobacteria for Agricultural Sustainability; Springer: Singapore, 2019; pp. 293–314. [Google Scholar] [CrossRef]
  60. Glick, B.R. Plant growth-promoting bacteria: Mechanisms and applications. Scientifica 2012, 2012, 963401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Ali, S.; Xie, L. Plant growth promoting and stress mitigating abilities of soil born microorganisms. Recent Pat. Food Nutr. Agric. 2020, 11, 96–104. [Google Scholar] [CrossRef]
  62. Bandyopadhyay, P.; Bhuyan, S.K.; Yadava, P.K.; Varma, A.; Tuteja, N. Emergence of plant and rhizospheric microbiota as stable interactomes. Protoplasma 2017, 254, 617–626. [Google Scholar] [CrossRef] [PubMed]
  63. Otieno, N.; Lally, R.D.; Kiwanuka, S.; Lloyd, A.; Ryan, D.; Germaine, K.J.; Dowling, D.N. Plant growth promotion induced by phosphate solubilizing endophytic Pseudomonas isolates. Front. Microbiol. 2015, 6, 745. [Google Scholar] [CrossRef] [Green Version]
  64. Farhangi-Abriz, S.; Tavasolee, A.; Ghassemi-Golezani, K.; Torabian, S.; Monirifar, H.; Rahmani, H.A. Growth-promoting bacteria and natural regulators mitigate salt toxicity and improve rapeseed plant performance. Protoplasma 2020, 257, 1035–1047. [Google Scholar] [CrossRef]
  65. Turan, M.; Ekinci, M.; Yildirim, E.; Güneş, A.; Karagöz, K.; Kotan, R.; Dursun, A. Plant growth-promoting rhizobacteria improved growth, nutrient, and hormone content of cabbage (Brassica oleracea) seedlings. Turk. J. Agric. For. 2014, 38, 327–333. [Google Scholar] [CrossRef]
  66. Szymańska, S.; Dąbrowska, G.B.; Tyburski, J.; Niedojadło, K.; Piernik, A.; Hrynkiewicz, K. Boosting the Brassica napus L. tolerance to salinity by the halotolerant strain Pseudomonas stutzeri ISE12. Environ. Exp. Bot. 2019, 163, 55–68. [Google Scholar] [CrossRef]
  67. Habibzadeh, F.; Sorooshzadeh, A.; Pirdashti, H.; Modarres Sanavy SA, M. A comparison between foliar application and seed inoculation of biofertilizers on canola (Brassica napus L.) grown under waterlogged conditions. Aust. J. Crop Sci. 2012, 6, 1435–1440. [Google Scholar]
  68. Yildrim, E.; Donmez, M.F.; Turan, M. Use of bioinoculants in ameliorative effects on radish plants under salinity stress. J. Plant Nutr. 2008, 31, 2059–2074. [Google Scholar] [CrossRef]
  69. Mohamed, H.I.; Gomaa, E.Z. Effect of plant growth promoting Bacillus subtilis and Pseudomonas fluorescens on growth and pigment composition of radish plants (Raphanus sativus) under NaCl stress. Photosynthetica 2012, 50, 263–272. [Google Scholar] [CrossRef]
  70. Arvin, P.; Vafabakhsh, J.; Mazaheri, D.; Noormohamadi, G.; Abdallah, I.M. Study of drought stress and plant growth promoting rhizobacteria (PGPR) on yield, yield components and seed oil content of different cultivars and species of Brassica oilseed rape. Ann. Biol. Res. 2012, 3, 4444–4451. [Google Scholar]
  71. Rajkumar, M.; Ma, Y.; Freitas, H. Improvement of Ni phytostabilization by inoculation of Ni resistant Bacillus megaterium SR28C. J. Environ. Manag. 2013, 128, 973–980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Pan, F.; Meng, Q.; Luo, S.; Shen, J.; Chen, B.; Yasmin, K.K.; Japenga, J.; Ma, X.; Yang, X.; Feng, Y. Enhanced Cd extraction of oilseed rape (Brassica napus) by plant growth promoting bacteria isolated from Cd hyperaccumulator Sedum alfredii hance. Int. J. Phytoremediation 2016, 19, 281–289. [Google Scholar] [CrossRef]
  73. Ma, Y.; Rajkumar, M.; Rocha, I.; Oliveira, R.S.; Freitas, H. Serpentine bacteria influence metal translocation and bio-concentration of Brassica juncea and Ricinus communis grown in multi-metal polluted soils. Front. Plant Sci. 2015, 757, 1–14. [Google Scholar] [CrossRef] [Green Version]
  74. Hussein, K.A.; Yoo, J.; Joo, J.H. Tolerance to Salt Stress by Plant Growth-Promoting Rhizobacteria on Brassica rapa var. glabrous. Korean J. Soil Sci. Fertil. 2016, 49, 776–782. [Google Scholar] [CrossRef] [Green Version]
  75. Samancioglu, A.; Yildirim, E.; Turan, M.; Kotan, R.; Sahin, U.; Kul, R. Amelioration of drought stress adverse effect and mediating biochemical content of cabbage seedlings by plant growth promoting rhizobacteria. Int. J. Agric. Biol. 2016, 18, 948–956. [Google Scholar] [CrossRef]
  76. Li, H.; Lei, P.; Pang, X.; Li, S.; Xu, H.; Xu, Z.; Feng, X. Enhanced tolerance to salt stress in canola (Brassica napus L.) seedlings inoculated with the halotolerant Enterobacter cloacae HSNJ4. Appl. Soil Ecol. 2017, 119, 26–34. [Google Scholar] [CrossRef]
  77. Qi, W.; Wen-Ji, Z.; Lin-Yan, H.; Xia-Fang, S. Increased biomass and quality and reduced heavy metal accumulation of edible tissues of vegetables in the presence of Cd-tolerant and immobilizing Bacillus megaterium H3. Ecotoxicol. Environ. Saf. 2018, 148, 269–274. [Google Scholar] [CrossRef]
  78. Ma, Y.; Rajkumar, M.; Zhang, C.; Freitas, H. Inoculation of Brassica oxyrrhina with plant growth promoting bacteria for the improvement of heavy metal phytoremediation under drought conditions. J. Hazard Mater. 2016, 320, 36–44. [Google Scholar] [CrossRef] [PubMed]
  79. Noreen, S.; Fatima, Z.; Ahmad, S.; Athar, H.u.R.; Ashraf, M. Foliar Application of Micronutrients in Mitigating Abiotic Stress in Crop Plants. In Plant Nutrients and Abiotic Stress Tolerance; Hasanuzzaman, M., Fujita, M., Oku, H., Nahar, K., Hawrylak-Nowak, B., Eds.; Springer: Singapore, 2018. [Google Scholar] [CrossRef]
  80. Dhaliwala, S.S.; Sharmaa, V.; Shuklab, A.K. Impact of micronutrients in mitigation of abiotic stresses in soils and plants—A progressive step toward crop security and nutritional quality. In Advances in Agronomy; Academic Press: Cambridge, MA, USA, 2022; pp. 1–78. [Google Scholar]
  81. Gondwe, R.L.; Kinoshita, R.; Suminoe, T.; Aiuchi, D.; Palta, J.P.; Tani, M. Available soil nutrients and NPK application impacts on yield, quality, and nutrient composition of potatoes growing during the main season in Japan. Am. J. Potato Res. 2020, 97, 234–245. [Google Scholar] [CrossRef]
  82. Voortman, R.; Bindraban, P.S. Beyond N and P: Toward a Land Resource Ecology Perspective and Impactful Fertilizer Interventions in Sub-Saharan Africa; VFRC Report 2015/1; Virtual Fertilizer Research Center: Washington, DC, USA, 2015; p. 49. [Google Scholar]
  83. Zhang, X.; Liu, F.; He, Y.; Gong, X. Detecting macronutrients content and distribution in oilseed rape leaves based on hyperspectral imaging. Biosyst. Eng. 2013, 115, 56–65. [Google Scholar] [CrossRef]
  84. Good, A. Toward nitrogen-fixing plants. Science 2018, 359, 869–870. [Google Scholar] [CrossRef]
  85. Galindo, F.S.; Teixeira Filho, M.C.M.; Buzetti, S.; Pagliari, P.H.; Santini, J.M.K.; Alves, C.J.; Megda, M.M.; Nogueira, T.A.R.; Andreotti, M.; Arf, O. Maize yield response to nitrogen rates and sources associated with Azospirillum brasilense. Agron. J. 2019, 111, 1985–1997. [Google Scholar] [CrossRef] [Green Version]
  86. Ichami, S.M.; Sheperd, K.D.; Sila, A.M.; Stoorvogel, J.J.; Hoffland, E. Fertilizer response and nitrogen use efficiency in African smallholder maize farms. Nutr. Cycl. Agroecosyst. 2019, 113, 1–19. [Google Scholar] [CrossRef] [Green Version]
  87. Lollato, R.P.; Figueired, B.M.; Dhillon, J.S.; Arnall, D.B.; Raun, W.R. Wheat grain yield and grain-nitrogen relationships as affected by N, P, and K fertilization: A synthesis of long-term experiments. Field Crops Res. 2019, 236, 42–57. [Google Scholar] [CrossRef]
  88. Tatsumi, K.; Abiko, T.; Kinose, Y.; Inagaki, S.; Izuta, T. Effects of ozone on the growth and yield of rice (Oryza sativa L.) under different nitrogen fertilization regimes. Environ. Sci. Poll. Res. 2019, 26, 32103–32113. [Google Scholar] [CrossRef]
  89. Omara, P.; Aula, L.; Dhillon, J.S.; Oyebiyi, F.; Eickhoff, E.M.; Nambi, E.; Fornah, A.; Carpenter, J.; Raun, W. Variability in winter wheat (Triticum aestivum L.) grain yield response to nitrogen fertilization in long-term experiments. Comm. Soil Sci. Plant Anal. 2020, 51, 403–412. [Google Scholar] [CrossRef]
  90. Rütting, T.; Aronsson, H.; Delin, S. Efficient use of nitrogen in agriculture. Nutr. Cycl. Agroecosyst. 2018, 110, 1–5. [Google Scholar] [CrossRef] [Green Version]
  91. Marschner, P. Marschner’s Mineral Nutrition of Higher Plants, 3rd ed.; Academic Press: New York, NY, USA, 2012; p. 651. [Google Scholar]
  92. Laperche, A.; Aigu, Y.; Jubault, M.; Ollier, M.; Guichard, S.; Glory, P.; Strelkov, S.E.; Gravot, A.; Manzanares-Dauleux, M.J. Clubroot resistance QTL are modulated by nitrogen input in Brassica napus. Theor. Appl. Genet. 2017, 130, 669–684. [Google Scholar] [CrossRef]
  93. Liu, T.; Ren, T.; White, P.J.; Cong, R.; Lu, J. Storage nitrogen co-ordinates leaf expansion and photosynthetic capacity in winter oilseed rape. J. Exp. Bot. 2018, 69, 2995–3007. [Google Scholar] [CrossRef] [PubMed]
  94. Iqbal, N.; Umar, S.; Khan, N.A. Nitrogen availability regulates proline and ethylene production and alleviates salinity stress in mustard (Brassica juncea). J. Plant Physiol. 2015, 178, 84–91. [Google Scholar] [CrossRef]
  95. Gupta, S.; Akhatar, J.; Kaur, P.; Sharma, A.; Sharma, P.; Mittal, M.; Bharti, B.; Banga, S.S. Genetic analyses of nitrogen assimilation enzymes in Brassica juncea (L.) Czern & Coss. Mol. Biol. Rep. 2019, 46, 4235–4244. [Google Scholar] [CrossRef]
  96. Nawaz, M.Q.; Ahmed, K.; Qadir, G.; Rizwan, M.; Nawaz, M.F.; Sarfraz, M. Growth and yield of turnip (Brassica rapa L.) in response to different sowing methods and nitrogen levels in salt-affected soils. Pakistan J. Agric. Res. 2020, 33, 126–134. [Google Scholar] [CrossRef]
  97. Baloch, P.A.; Uddin, R.; Nizamani, F.K.; Solangi, A.H.; Siddiqui, A.A. Effect of nitrogen, phosphorus and potassium fertilizers on growth and yield characteristics of radish (Raphinus sativus L.). Am.-Eurasian J. Agric. Environ. Sci. 2013, 14, 565–569. [Google Scholar]
  98. Bozkurt, S.; Uygur, V.; Agca, N. Yield responses of cauliflower (Brassica oleracea L. var. Botrytis) to different water and nitrogen levels in a Mediterranean coastal area. Acta Agric. Scand. Sect. B Soil Plant Sci. 2011, 61, 183–194. [Google Scholar] [CrossRef]
  99. Xie, Y.; Kristensen, H.L. Overwintering grass-clover as intercrop and moderately reduced nitrogen fertilization maintain yield and reduce the risk of nitrate leaching in an organic cauliflower (Brassica oleracea L. var. botrytis) agroecosystem. Sci. Hortic. 2016, 206, 71–79. [Google Scholar] [CrossRef]
  100. Erley, G.S.; Dewi, E.R.; Nikus, O.; Horst, W.J. Genotypic differences in nitrogen efficiency of white cabbage (Brassica oleracea L.). Plant Soil 2010, 328, 313–325. [Google Scholar] [CrossRef]
  101. Shan, L.; He, Y.; Chen, J.; Huang, Q.; Wang, H. Ammonia volatilization from a Chinese cabbage field under different nitrogen treatments in the Taihu Lake Basin, China. J. Environ. Sci. 2015, 38, 14–23. [Google Scholar] [CrossRef] [PubMed]
  102. Erdem, T.; Arın, L.; Erdem, Y.; Polat, S.; Deveci, M.; Okursoy, H.; Gültaş, H.T. Yield and quality response of drip irrigated broccoli (Brassica oleracea L. var. italica) under different irrigation regimes, nitrogen applications and cultivation periods. Agric. Water Manag. 2010, 97, 681–688. [Google Scholar] [CrossRef]
  103. Elwan, M.W.M.; El-Hamed, K.E. Influence of nitrogen form, growing season and sulfur fertilization on yield and the content of nitrate and vitamin C of broccoli. Sci. Hortic. 2011, 127, 181–187. [Google Scholar] [CrossRef]
  104. Groenbaek, M.; Jensen, S.; Neugart, S.; Schreiner, M.; Kidmose, U.; Kristensen, H.L. Nitrogen split dose fertilization, plant age and frost effects on phytochemical content and sensory properties of curly kale (Brassica oleracea L. var. sabellica). Food Chem. 2016, 197, 530–538. [Google Scholar] [CrossRef]
  105. Zhu, Y.; Li, G.; Liu, H.; Sun, G.; Chen, R.; Song, S. Effects of partial replacement of nitrate with different nitrogen forms on the yield, quality and nitrate content of Chinese kale. Comm. Soil Sci. Plant Anal. 2018, 49, 1384–1393. [Google Scholar] [CrossRef]
  106. Benett, K.S.S.; Xavier, R.C.; Benett, C.G.S.; Salomão, L.C.; Seleguini, A.; Cantuario, F.S.; Martins, A.S. Nitrogen application in arugula culture. J. Agric. Sci. 2019, 11, 2. [Google Scholar] [CrossRef]
  107. Albert, B.; Le Cahérec, F.; Niogret, M.F.; Faes, P.; Avice, J.C.; Leport, L.; Bouchereau, A. Nitrogen availability impacts oilseed rape (Brassica napus L.) plant water status and proline production efficiency under water-limited conditions. Planta 2012, 236, 659–676. [Google Scholar] [CrossRef] [Green Version]
  108. Goel, P.; Singh, A.K. Abiotic stresses downregulate key genes involved in nitrogen uptake and assimilation in Brassica juncea L. PLoS ONE 2015, 10, e0143645. [Google Scholar] [CrossRef] [Green Version]
  109. Zhu, Z.; Gerendas, J.; Bendixen, R.; Schinner, K.; Tabrizi, H.; Sattelmacher, B.; Hansen, U.P. Different tolerance to light stress in N03--and NH4+-grown Phaseolus vulgaris L. Plant Biol. 2000, 2, 558–570. [Google Scholar] [CrossRef]
  110. Saud, S.; Fahad, S.; Yajun, C.; Ihsan, M.Z.; Hammad, H.M.; Nasim, W.; Amanullah, J.; Arif, M.; Alharby, H. Effects of Nitrogen Supply on Water Stress and Recovery Mechanisms in Kentucky Bluegrass Plants. Front. Plant Sci. 2017, 8, 983. [Google Scholar] [CrossRef] [Green Version]
  111. Prado, R.M. Plant Nutrition; UNESP: São Paulo, Brazil, 2008; p. 347. [Google Scholar]
  112. Filgueira, F.A.R. Novo Manual de Olericultura: Agrotecnologia Moderna na Produção e Comercialização de Hortaliças; UFV: Viçosa, Brazil, 2008; p. 242. [Google Scholar]
  113. Holzschuh, M.J.; Bartz, H.R.; Trevisan, J.N.; Martins GA, K. Parâmetros de rendimento de brassicáceas e a disponibilidade de fósforo em Planossolo Hidromórfico distrófico arênico estimado pelo extrator de Mehlich-1. In Proceedings of the Reunião Regional da Sociedade Brasileira para o Progresso da Ciência no Rio Grande do Sul, Santa Maria, Brazil, 23–26 May 2004. [Google Scholar]
  114. Zaman, B.; Ali, A.; Huma, M.K.; Arshadullah, M.; Mahmood, I.A. Growth responses of Brassica juncea to phosphorus application from different sources of fertilizer under salt stress. Songklanakarin J. Sci. Technol. 2015, 37, 631–634. [Google Scholar]
  115. Numan, M.; Bashir, S.; Khan, Y.; Mumtaz, R.; Shinwari, Z.K.; Khan, A.L.; Khan, A.; AL-Harrasi, A. Plant growth promoting bacteria as an alternative strategy for salt tolerance in plants: A review. Microbiol. Res. 2018, 209, 21–32. [Google Scholar] [CrossRef] [PubMed]
  116. Payne, W.A.; Drew, M.C.; Hossner, L.R.; Lascano, R.J.; Onken, A.B.; Wendt, C.W. Soil phosphorus availability and pearl millet water-use efficiency. Crop Sci. 1992, 32, 1010–1015. [Google Scholar] [CrossRef]
  117. Singh, D.K.; Sale PW, G. Growth and potential conductivity of white clover roots in dry soil with increasing phosphorus supply and defoliation frequency. Agron. J. 2000, 92, 868–874. [Google Scholar] [CrossRef]
  118. Jones, C.A.; Jacobsen, J.S.; Wraith, J.M. The effects of P fertilization on drought tolerance of malt barley. Nutr. Manag. Conf. 2003, 5, 88–93. [Google Scholar]
  119. Jin, J.; Wang, G.; Liu, X.; Pan, X.; Herbert, S.J.; Tang, C. Interaction Between Phosphorus Nutrition and Drought on Grain Yield, and Assimilation of phosphorus and nitrogen in two soybean cultivars differing in protein concentration in grains. J. Plant Nutr. 2006, 29, 1433–1449. [Google Scholar] [CrossRef]
  120. Arshad, M.; Ali, S.; Noman, A.; Ali, Q.; Rizwan, M.; Farid, M.; Irshad, M.K. Phosphorus amendment decreased cadmium (Cd) uptake and ameliorates chlorophyll contents, gas exchange attributes, antioxidants, and mineral nutrients in wheat (Triticum aestivum L.) under Cd stress. Arch. Agron. Soil Sci. 2016, 62, 533–546. [Google Scholar] [CrossRef]
  121. Wang, B.; Xie, Z.; Chen, J.; Jiang, J.; Su, Q. Effects of field application of phosphate fertilizers on the availability and uptake of lead, zinc and cadmium by cabbage Brassica chinensis L. in a mining tailing contaminated soil. J. Environ. Sci. 2008, 20, 1109–1117. [Google Scholar] [CrossRef]
  122. Siebers, N.; Siangliw, M.; Tongcumpou, C. Cadmium uptake and subcellular distribution in rice plants as affected by phosphorus: Soil and hydroponic experiments. J. Soil Sci. Plant Nutr. 2013, 3, 833–844. [Google Scholar] [CrossRef] [Green Version]
  123. Wang, M.; Zheng, Q.; Shen, Q.; Guo, S. The critical role of potassium in plant stress response. Int. J. Mol. Sci. 2013, 14, 7370–7390. [Google Scholar] [CrossRef] [Green Version]
  124. Basu, S.; Ramegowda, V.; Kumar, A.; Pereira, A. Plant Adaptation to Drought Stress. F1000Research 2016, 5, F1000. [Google Scholar] [CrossRef] [PubMed]
  125. Bahrami-Radb, S.; Hajiboland, R. Effect of potassium application in drought-stressed tobacco (Nicotiana rustica L.) plants: Comparison of root with foliar application. Ann. Agric. Sci. 2017, 62, 121–130. [Google Scholar] [CrossRef]
  126. Hasanuzzaman, M.; Bhuyan, M.H.M.B.; Nahar, K.; Hossain, M.D.; Mohmud, J.A.; Hossen, M.S.; Masud, A.A.C.; Moumita, M. Potassium: A vital regulator of plant responses and tolerance to abiotic stresses. Agronomy 2018, 8, 31. [Google Scholar] [CrossRef] [Green Version]
  127. Siddiqui, M.H.; Al-Whaibi, M.H.; Sakran, A.M.; Basalah, M.O.; Ali, H.M. Effect of calcium and potassium on antioxidant system of Vicia faba L. under cadmium stress. Int. J. Mol. Sci. 2012, 13, 6604–6619. [Google Scholar] [CrossRef] [Green Version]
  128. Song, Z.Z.; Duan, C.L.; Guo, S.L.; Yang, Y.; Feng, Y.F.; Ma, R.J.; Yu, M.L. Potassium contributes to zinc stress tolerance in peach (Prunus persica) seedlings by enhancing photosynthesis and the antioxidant defense system. Genet. Mol. Res. 2015, 14, 8338–8351. [Google Scholar] [CrossRef] [PubMed]
  129. Yang, Y.; Zheng, Q.; Liu, M.; Long, X.; Liu, Z.; Shen, Q.; Guo, S. Difference in sodium spatial distribution in the shoot of two canola cultivars under saline stress. Plant Cell Physiol. 2012, 53, 1083–1092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Ardestani, H.G.; Shirani, A.H.; Zandi, P. Effect of drought stress on some agronomic traits of two rapeseed varieties grown under different potassium rates. Aust. J. Basic Appl. Sci. 2011, 5, 2875–2882. [Google Scholar]
  131. Schumann, T.; Paul, S.; Melzer, M.; Dörmann, P.; Jahns, P. Plant growth under natural light conditions provides highly flexible short-term acclimation properties toward high light stress. Front. Plant Sci. 2017, 8, 681. [Google Scholar] [CrossRef] [Green Version]
  132. Samar-Raza, M.A. Potassium applied under drought improves physiological and nutrient uptake performances of wheat (Triticum aestivun L.). J. Soil Sci. Plant Nutr. 2013, 13, 175–185. [Google Scholar] [CrossRef] [Green Version]
  133. Shah, S.H.; Houborg, R.; McCabe, M.F. Response of chlorophyll, carotenoid and SPAD-502 measurement to salinity and nutrient stress in wheat (Triticum aestivum L.). Agronomy 2017, 7, 61. [Google Scholar] [CrossRef] [Green Version]
  134. Rivas-Sendra, A.; Calabiug-Serna, A.; Seguí-Simarro, J.M. Dynamics of calcium during in vitro microspore embryogenesis and in vivo microspore development in Brassica napus and Solanum melongena. Front. Plant Sci. 2017, 7, 1177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Sanders, D.; Pelloux, J.; Brownlee, C.; Harper, J.F. Calcium at the crossroads of signaling. Plant Cell 2002, 14, 401–417. [Google Scholar] [CrossRef] [Green Version]
  136. Suzuki, N. Alleviation by calcium of cadmium-induced root growth inhibition in Arabidopsis seedlings. Plant Biotechnol. 2005, 22, 19–25. [Google Scholar] [CrossRef] [Green Version]
  137. Ahmad, P.; Sarwat, M.; Bhat, N.A.; Wani, M.R.; Kazi, A.G.; Tran, L.S. Alleviation of cadmium toxicity in Brassica juncea L. (Czern. & Coss.) by calcium application involves various physiological and biochemical strategies. PLoS ONE 2015, 10, e0114571. [Google Scholar] [CrossRef] [Green Version]
  138. Williams, L.; Salt, D.E. The plant ionome coming into focus. Curr. Opin. Plant Biol. 2009, 12, 247–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Hortensteiner, S. Stay-green regulates chlorophyll and chlorophyll-binding protein degradation during senescence. Trends Plant Sci. 2009, 14, 155–162. [Google Scholar] [CrossRef]
  140. Cakmak, I.; Yazici, A.M. Magnesium: A forgotten element in crop production. Better Crops 2010, 94, 23–25. [Google Scholar]
  141. Tang, R.J.; Luan, S. Regulation of calcium and magnesium homeostasis in plants: From transporters to signaling network. Curr. Opin. Plant Biol. 2017, 39, 97–105. [Google Scholar] [CrossRef]
  142. Senbayram, M.; Gransee, A.; Wahle, V.; Thiel, H. Role of magnesium fertilisers in agriculture: Plant–soil continuum. Crop Pasture Sci. 2016, 66, 1219–1229. [Google Scholar] [CrossRef]
  143. Waraich, E.A.; Ahmad, R.; Ashraf, M.Y. Role of mineral nutrition in alleviation of drought stress in plants. Aust. J. Crop Sci. 2011, 5, 764–777. [Google Scholar]
  144. Ashraf, M.; Athar, H.R.; Harris, P.J.C.; Kwon, T.R. Some prospective strategies for improving crop salt tolerance. Adv. Agron. 2008, 97, 45–110. [Google Scholar] [CrossRef]
  145. Kashem, M.A.; Kawai, S. Alleviation of cadmium phytotoxicity by magnesium in Japanese mustard spinach. Soil Sci. Plant Nutr. 2007, 53, 246–251. [Google Scholar] [CrossRef]
  146. Guo, W.; Nazim, H.; Liang, Z.; Yang, D. Magnesium deficiency in plants: An urgent problem. Crop J. 2016, 4, 83–91. [Google Scholar] [CrossRef] [Green Version]
  147. Verbruggen, N.; Hermans, C. Physiological and molecular responses to magnesium nutritional imbalance in plants. Plant Soil 2013, 368, 87–99. [Google Scholar] [CrossRef]
  148. Cakmak, I.; Kirkby, E.A. Role of magnesium in carbon partitioning and alleviating photooxidative damage. Physiol. Plant. 2008, 133, 692–704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Foyer, C.H.; Noctor, G. Redox homeostasis and antioxidant signaling: A metabolic interface between stress perception and physiological responses. Plant Cell 2005, 17, 1866–1875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Singla-Pareek, S.L.; Yadav, S.K.; Pareek, A.; Reddy, M.K.; Sopory, S.K. Transgenic tobacco overexpressing glyoxalase pathway enzymes grow and set viable seeds in zinc-spiked soils. Plant Physiol. 2006, 140, 613–623. [Google Scholar] [CrossRef] [Green Version]
  151. Yadav, S.K. Heavy metals toxicity in plants: An overview on the role of glutathione and phytochelatins in heavy metal stress tolerance of plants. S. Afr. J. Bot. 2010, 76, 167–179. [Google Scholar] [CrossRef]
  152. Rengel, Z.; Bose, J.; Chen, Q.; Tripathi, B.N. Magnesium alleviates plant toxicity of aluminium and heavy metals. Crop Pasture Sci. 2015, 66, 1298–1307. [Google Scholar] [CrossRef]
  153. Hall, J.Á.; Williams, L.E. Transition metal transporters in plants. J. Exp. Bot. 2003, 54, 2601–2613. [Google Scholar] [CrossRef] [Green Version]
  154. Li, L.; Tutone, A.F.; Drummond, R.S.; Gardner, R.C.; Luan, S. A novel family of magnesium transport genes in Arabidopsis. Plant Cell 2001, 13, 2761–2775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Stipp, R.S.; Casarin, V. A importância do enxofre na agricultura brasileira. In Informações Agronômicas; n. 129; International Plant Nutrition Institute: Norcross, GA, USA, 2010. [Google Scholar]
  156. Capaldi, F.R.; Gratão, P.L.; Reis, A.R.; Lima, L.W.; Azevedo, R.A. Sulfur metabolism and stress defense responses in plants. Trop. Plant Biol. 2015, 8, 60–73. [Google Scholar] [CrossRef] [Green Version]
  157. Malavolta, E.; Morais, M.F. Fundamentos do nitrogênio e do enxofre na nutrição mineral das plantas cultivadas. In Nitrogênio e Enxofre na Agricultura Brasileira; Yamada, T., Abdalla, S.R.S., Vitti, G.C., Eds.; International Plant Nutrition Institute: Piracicaba, Brazil, 2007; pp. 189–249. [Google Scholar]
  158. Raij, B.V. Fertilidade do Solo e Manejo de Nutrientes; International Plant Nutrition Institute: Piracicaba, Brazil, 2011. [Google Scholar]
  159. Losák, T.; Hlušek, J.; Kráěmar, S.; Varga, L. The effect of nitrogen and sulphur fertilization on yield and quality of kohlrabi (Brassica oleracea, L.). Rev. Bras. Ciênc. Solo 2008, 32, 697–703. [Google Scholar] [CrossRef] [Green Version]
  160. Santos, R.F.; Borsoi, A.; Secco, D.; Souza SN, M.; Frigo, E.P. Nitrogen and sulfur sources in the culture of Brassica napus L. var. oleifera. J. Food Agric. Environ. 2012, 10, 516–518. [Google Scholar]
  161. Bashir, H.; Ahmad, J.; Bagheri, R.; Nauman, M.; Qureshi, M.I. Limited sulfur resource forces Arabidopsis thaliana to shift towards non-sulfur tolerance under cadmium stress. Environ. Exp. Bot. 2013, 94, 19–32. [Google Scholar] [CrossRef]
  162. Skeffington, R.A.; Shewry, P.R.; Peterson, P.J. Chromium uptake and transport in barley seedlings (Hordeum vulgare L.). Planta 1976, 132, 209–214. [Google Scholar] [CrossRef] [PubMed]
  163. Schiavon, M.; Pilon-Smits, E.A.; Wirtz, M.; Hell, R.; Malagoli, M. Interactions between chromium and sulfur metabolism in Brassica juncea. J. Environ. Qual. 2008, 37, 1536–1545. [Google Scholar] [CrossRef] [Green Version]
  164. Pedler, J.F.; Kinraide, T.B.; Parker, D.R. Zinc rhizotoxicity in wheat and radish is alleviated by micromolar levels of magnesium and potassium in solution culture. Plant and Soil 2004, 259, 191–199. [Google Scholar] [CrossRef]
  165. Siddiqui, M.H.; Mohammad, F.; Khan, M.N.; Al-Whaibi, M.H.; Bahkali AH, A. Nitrogen in relation to photosynthetic capacity and accumulation of osmoprotectant and nutrients in Brassica genotypes grown under salt stress. Agric. Sci. China 2010, 9, 671–680. [Google Scholar] [CrossRef]
  166. Vatehová, Z.; Kollárová, K.; Zelko, I.; Richterová-Kučerová, D.; Bujdoš, M.; Lišková, D. Interaction of silicon and cadmium in Brassica juncea and Brassica napus. Biologia 2012, 67, 498–504. [Google Scholar] [CrossRef]
  167. Rais, L.; Masood, A.; Inam, A.; Khan, N. Sulfur and nitrogen co-ordinately improve photosynthetic efficiency, growth and proline accumulation in two cultivars of mustard under salt stress. J. Plant Biochem. Physiol. 2013, 1, 1–6. [Google Scholar] [CrossRef] [Green Version]
  168. Augusto, A.S.; Bertoli, A.C.; Cannata, M.G.; Carvalho, R.; Bastos, A.R.R. Bioacumulação de metais pesados em Brassica juncea: Relação de toxicidade com elementos essenciais. Rev. Virtual Quím. 2014, 6, 1221–1236. [Google Scholar] [CrossRef]
  169. Pandey, C.; Khan, E.; Panthri, M.; Tripathi, R.D.; Gupta, M. Impact of silicon on Indian mustard (Brassica juncea L.) root traits by regulating growth parameters, cellular antioxidants and stress modulators under arsenic stress. Plant Physiol. Biochem. 2016, 104, 216–225. [Google Scholar] [CrossRef] [PubMed]
  170. Ashfaque, F.; Inam, A.; Inam, A.; Iqbal, S.; Sahay, S. Response of silicon on metal accumulation, photosynthetic inhibition and oxidative stress in chromium-induced mustard (Brassica juncea L.). S. Afr. J. Bot. 2017, 111, 153–160. [Google Scholar] [CrossRef]
  171. Siddiqui, H.; Yusuf, M.; Faraz, A.; Faizan, M.; Sami, F.; Hayat, S. 24-Epibrassinolide supplemented with silicon enhances the photosynthetic efficiency of Brassica juncea under salt stress. S. Afr. J. Bot. 2018, 118, 120–128. [Google Scholar] [CrossRef]
  172. Guo, D.; Ren, C.; Ali, A.; Du, J.; Zhang, Z.; Li, R.; Zhang, Z. Streptomyces pactum and sulfur mediated the antioxidant enzymes in plant and phytoextraction of potentially toxic elements from a smelter-contaminated soils. Environ. Pollut. 2019, 251, 37–44. [Google Scholar] [CrossRef]
  173. Fanaei, H.R.; Galavi, M.; Kafi, M.; Bonjar, A.G. Amelioration of water stress by potassium fertilizer in two oilseed species. Int. J. Plant Prod. 2009, 3, 41–54. [Google Scholar] [CrossRef]
  174. Hashemi, A.; Abdolzadeh, A.; Sadeghipour, H.R. Beneficial effects of silicon nutrition in alleviating salinity stress in hydroponically grown canola, Brassica napus L. plants. Comm. Soil Sci. Plant Nutr. 2010, 56, 244–253. [Google Scholar] [CrossRef] [Green Version]
  175. Anjum, N.A.; Gill, S.S.; Umar, S.; Ahmad, I.; Duarte, A.C.; Pereira, E. Improving growth and productivity of oleiferous Brassicas under changing environment: Significance of nitrogen and sulphur nutrition, and underlying mechanisms. Sci. World J. 2012, 2012, 657808. [Google Scholar] [CrossRef]
  176. Farshidi, M.; Abdolzadeh, A.; Sadeghipour, H.R. Silicon nutrition alleviates physiological disorders imposed by salinity in hydroponically grown canola (Brassica napus L.) plants. Acta Physiol. Plant. 2012, 34, 1779–1788. [Google Scholar] [CrossRef]
  177. Habibi, G. Contrastive response of Brassica napus L. to exogenous salicylic acid, selenium and silicon supplementation under water stress. Arch. Biol. Sci. 2015, 67, 397–404. [Google Scholar] [CrossRef]
  178. Al-Solaimani, S.G.; Alghabari, F.; Ihsan, M.Z. Effect of different rates of nitrogen fertilizer on growth, seed yield, yield components and quality of canola (Brassica napus L.) under arid environment of Saudi Arabia. Int. J. Agron. Agric. Res. 2015, 6, 268–274. [Google Scholar]
  179. Hasanuzzaman, M.; Nahar, K.; Anee, T.I.; Fujita, M. Exogenous silicon attenuates cadmium-induced oxidative stress in Brassica napus L. by modeling AsA-GSH pathway and glyoxylate system. Front. Plant Sci. 2017, 8, 1061. [Google Scholar] [CrossRef]
  180. Haddad, C.; Arkoun, M.; Jamois, F.; Schwarzenberg, A.; Yvin, J.C.; Etienne, P.; Laine, P. Silicon promotes growth of Brassica napus L. and delays leaf senescence induced by nitrogen starvation. Front. Plant Sci. 2018, 9, 516. [Google Scholar] [CrossRef]
  181. Cong, R.; Liu, T.; Lu, P.; Ren, T.; Li, X.; Lu, J. Nitrogen fertilization compensation the weak photosynthesis of Oilseed rape (Brassca napus L.) under haze weather. Sci. Rep. 2020, 10, 4047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Naveed, M.; Sajid, H.; Mustafa, A.; Niamat, B.; Ahmad, Z.; Yaseen, M.; Kamran, M.; Ahmar, S.; Chen, J.T. Alleviation of Salinity-Induced Oxidative Stress, Improvement in Growth, Physiology and Mineral Nutrition of Canola (Brassica napus L.) through Calcium-Fortified Composted Animal Manure. Sustainability 2020, 12, 846. [Google Scholar] [CrossRef] [Green Version]
  183. Bybordi, A.; Ebrahimian, E. Effect of salinity stress on activity of enzymes involved in nitrogen and phosphorous metabolism case study: Canola (Brassica napus L.). Asian J. Agric. Res. 2011, 5, 208–214. [Google Scholar] [CrossRef]
  184. Terzi, H.; Yıldız, M. Interactive effects of sulfur and chromium on antioxidative defense systems and BnMP1 gene expression in canola (Brassica napus L.) cultivars differing in Cr(VI) tolerance. Ecotoxicology 2015, 24, 1171–1182. [Google Scholar] [CrossRef] [PubMed]
  185. Rezayian, M.; Niknam, V.; Ebrahimzadeh, H. Penconazole and calcium improves drought stress tolerance and oil quality in canola. Soil Sci. Plant Nutr. 2018, 64, 1–10. [Google Scholar] [CrossRef]
  186. Farahani, S.; Heravan, E.M.; Rad, A.H.S.; Noormohammadi, G. Effect of potassium sulfate on quantitative and qualitative characteristics of canola cultivars upon late-season drought stress conditions. J. Plant Nutr. 2019, 42, 1543–1555. [Google Scholar] [CrossRef]
  187. Katroschan, K.W.; Uptmoor, R.; Stützel, H. Nitrogen use efficiency of organically fertilized white cabbage and residual effects on subsequent beetroot. Plant Soil 2014, 382, 237–251. [Google Scholar] [CrossRef]
  188. Wu, Z.; Liu, S.; Zhao, J.; Wang, F.; Du, Y.; Zou, S.; Li, H.; Wen, D.; Huang, Y. Comparative responses to silicon and selenium in relation to antioxidant enzyme system and the glutathione-ascorbate cycle in flowering Chinese cabbage (Brassica campestris L. ssp. chinensis var. utilis) under cadmium stress. Environ. Exp. Bot. 2017, 133, 1–11. [Google Scholar] [CrossRef]
  189. Ahmad, W.; Ayyub, C.M.; Shehzad, M.A.; Ziaf, K.; Ijaz, M.; Sher, A.; Abbas, T.; Shafi, J. Supplemental potassium mediates antioxidant metabolism, physiological processes, and osmoregulation to confer salt stress tolerance in cabbage (Brassica oleracea L.). Hortic. Environ. Biotechnol. 2019, 60, 853–869. [Google Scholar] [CrossRef]
  190. Yang, R.; Guo, L.; Zhou, Y.; Shen, C.; Gu, Z. Calcium mitigates the stress caused by ZnSO4 as a sulfur fertilizer and enhances the sulforaphane formation of broccoli sprouts. RSC Adv. 2015, 5, 12563–12570. [Google Scholar] [CrossRef]
  191. Zaghdoud, C.; Carvajal, M.; Moreno, D.A.; Ferchichi, A.; Martínez-Ballesta, M.C. Health-promoting compounds of broccoli (Brassica oleracea L. var. italica) plants as affected by nitrogen fertilisation in projected future climatic change environments. J. Sci. Food Agric. 2016, 96, 392–403. [Google Scholar] [CrossRef]
  192. Barreto, R.F.; Schiavon Júnior, A.A.; Maggio, M.A.; Prado, R.M. ilicon alleviates ammonium toxicity in cauliflower and in broccoli. Sci. Hortic. 2017, 225, 743–750. [Google Scholar] [CrossRef] [Green Version]
  193. Kałużewicz, A.; Bosiacki, M.; Spiżewski, T. Influence of biostimulants on the content of macro- and micronutrients in broccoli plants exposed to drought stress. J. Elem. 2018, 23, 287–297. [Google Scholar] [CrossRef]
  194. Samuolienė, G.; Brazaitytė, A.; Viršilė, A.; Miliauskienė, J.; Vaštakaitė-Kairienė, V.; Duchovskis, P. Nutrient Levels in Brassicaceae Microgreens Increase Under Tailored Light-Emitting Diode Spectra. Front. Plant Sci. 2019, 10, 1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Freitas, E.M.; Giovanelli, L.B.; Delazari, F.T.; Santos, M.L.; Pereira, S.B.; Silva, D.J.H. Arugula production as a function of irrigation depths and potassium fertilization. Rev. Bras. Eng. Agríc. Ambient. 2017, 21, p197–p202. [Google Scholar] [CrossRef]
  196. Jesus, E.G.; Fatima, R.T.; Guerrero, A.C.; Araújo, J.L.; Brito ME, B. Growth and gas exchanges of arugula plants under silicon fertilization and water restriction. Rev. Bras. Eng. Agríc Ambient. 2018, 22, 119–124. [Google Scholar] [CrossRef]
  197. Souza, J.Z.; Prado, R.M.; Silva SL, O.; Farias, T.P.; Garcia Neto, J.; Souza Junior, J.P. Silicon leaf fertilization promotes biofortification and increases dry matter, ascorbate content, and decreases post-harvest leaf water loss of Chard and Kale. Comm. Soil Sci. Plant Anal. 2018, 50, 164–172. [Google Scholar] [CrossRef]
  198. Chen, H.; Shu, F.; Yang, S.; Li, Y.; Wang, S. Competitive Inhibitory Effect of Calcium Polypeptides on Cd Enrichment of Brassia campestris L. Int. J. Environ. Res. Public Health 2019, 16, 4472. [Google Scholar] [CrossRef] [Green Version]
  199. Tubana, B.S.; Tapasya, B.; Datnoff, L.E. A review of silicon in soils and plants and its role in US agriculture: History and future perspectives. Soil Sci. 2016, 181, 393–411. [Google Scholar] [CrossRef] [Green Version]
  200. Neu, S.; Schaller, J.; Dudel, E.G. Silicon availability modifies nutrient use efficiency and content, C:N:P stoichiometry, and productivity of winter wheat (Triticum aestivum L.). Sci. Rep. 2017, 7, 40829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Haynes, R.J. Significance and role of Si in crop production. Adv. Agron. 2017, 146, 83–166. [Google Scholar]
  202. Linden, C.H.; Delvaux, B. The weathering stage of tropical soils affects the soil-plant cycle of silicon, but depending on land use. Geoderma 2019, 351, 209–220. [Google Scholar] [CrossRef]
  203. Coskun, D.; Britto, D.T.; Huynh, W.Q.; Kronzucker, H.J. The role of silicon in higher plants under salinity and drought stress. Front. Plant Sci. 2016, 7, 1072. [Google Scholar] [CrossRef] [Green Version]
  204. Pati, S.; Pal, B.; Badole, S.; Hazra, G.C.; Mandal, B. Effect of silicon fertilization on growth, yield, and nutrient uptake of rice. Comm. Soil Sci. Plant Anal. 2016, 47, 284–290. [Google Scholar] [CrossRef]
  205. Crusciol, C.A.C.; Arruda, D.P.; Fernandes, A.M.; Antonangelo, J.A.; Alleoni, L.R.F.; Nascimento, C.A.C.; Rossato, O.B.; Mccray, J.M. Methods and extractants to evaluate silicon availability for sugarcane. Sci. Rep. 2018, 8, 1–14. [Google Scholar] [CrossRef] [Green Version]
  206. Ramírez-Olvera, S.M.; Trejo-Téllez, L.I.; Pérez-Sato, J.A.; Gómez-Merino, F.C. Silicon stimulates initial growth and chlorophyll a/b ratio in rice seedlings, and alters the concentrations of Ca, B, and Zn in plant tissues. J. Plant Nutr. 2019, 42, 1928–1940. [Google Scholar] [CrossRef]
  207. Vega, I.; Nikolic, M.; Pontigo, S.; Godoy, K.; Mora ML, L.; Cartes, P. Silicon improves the production of high antioxidant or structural phenolic compounds in barley under aluminum stress cultivars. Agronomy 2019, 9, 388. [Google Scholar] [CrossRef]
  208. Guntzer, F.; Keller, C.; Meunier, J. Benefits of plant silicon for crops: A review. Agron. Sustain. Dev. 2012, 32, 201–213. [Google Scholar] [CrossRef] [Green Version]
  209. Isa, M.; Bai, S.; Yokoyama, T.; Ma, J.F.; Ishibashi, Y.; Yuasa, T.; Iwaya-Inoue, M. Silicon enhances growth independent of silica deposition in a low-silica rice mutant, lsi1. Plant Soil 2010, 331, 361–375. [Google Scholar] [CrossRef]
  210. Amin, M.; Ahmad, R.; Ali, A.; Hussain, I.; Mahmood, R.; Aslam, M.; Lee, D.J. Influence of silicon fertilization on maize performance under limited water supply. Silicon 2018, 10, 177–183. [Google Scholar] [CrossRef]
  211. Artyszak, A. Effect of silicon fertilization on crop yield quantity and quality—A literature review in Europe. Plants 2018, 7, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Baggio, G.; Dupas, E.; Galindo, F.S.; Megda, M.M.; Pereira, N.C.M.; Luchetta, M.O.; Tritapepe, C.A.; da Silva, M.R.; Jalal, A.; Teixeira Filho, M.C.M. Silicon Application Induced Alleviation of Aluminum Toxicity in Xaraés Palisadegrass. Agronomy 2021, 11, 1938. [Google Scholar] [CrossRef]
  213. Galindo, F.S.; Pagliari, P.H.; Buzetti, S.; Rodrigues, W.L.; Fernandes, G.C.; Biagini, A.L.C.; Marega, E.M.R.; Tavanti, R.F.R.; Jalal, A.; Teixeira Filho, M.C.M. Corn shoot and grain nutrient uptake affected by silicon application combined with Azospirillum brasilense inoculation and nitrogen rates. J. Plant Nutr. 2021, 45, 168–184. [Google Scholar] [CrossRef]
  214. Etesami, H.; Jeong, B.R. Review and future prospects on the action mechanisms in alleviating biotic and abiotic stresses in plants. Ecotoxicol. Environ. Saf. 2018, 147, 881–896. [Google Scholar] [CrossRef] [PubMed]
  215. Malhotra, C.; Kapoor, R.T. Silicon: A sustainable tool in abiotic stress tolerance in plants. In Plant Abiotic Stress Tolerance; Hasanuzzaman, M., Hakeem, K., Nahar, K., Alharby, H., Eds.; Springer: Berlin/Heidelberg, Germany, 2019; pp. 333–356. [Google Scholar] [CrossRef]
  216. Artyszak, A.; Kucinska, K. Silicon nutrition and crop improvement: Recent advances and future perspective. In Silicon in Plants; Tripathi, D.K., Singh, V.P., Ahmad, P., Chauhan, D.K., Prasad, S.M., Eds.; CRC Press: London, UK; New York, NY, USA, 2016; pp. 297–319. [Google Scholar]
  217. Angle, J.S.; Singh, U.; Dimpka, C.O.; Hellums, D.; Bindraban, P.S. Role of fertilisers for climate-resilient agriculture. In Proceedings of the International Fertiliser Society; International Fertiliser Society: Colchester, UK, 2017; pp. 1–44. [Google Scholar]
  218. Dimkpa, C.O.; Singh, U.; Bindraban, P.S.; Adisa, I.O.; Elmer, W.H.; Gardea-Torresdey, J.L.; White, J.C. Addition-omission of zinc, copper, and boron nano and bulk oxide particles demonstrate element and size-specific response of soybean to micronutrients exposure. Sci. Total Environ. 2019, 665, 606–616. [Google Scholar] [CrossRef]
  219. Iqbal, A.; Raza, H.; Zaman, M.; Khan, R.; Adnan, M.; Khan, A.; Gillani, S.W.; Khalil, S.K. Impact of Nitrogen, Zinc and Humic Acid Application on Wheat Growth, Morphological Traits, Yield and Yield Components. J. Soil Plant Environ. 2022, 1, 50–71. [Google Scholar] [CrossRef]
  220. Etesami, H.; Maheshwari, D.K. Use of plant growth promoting rhizobacteria (PGPRs) with multiple plant growth promoting traits in stress agriculture: Action mechanisms and future prospects. Ecotoxicol. Environ. Saf. 2018, 156, 225–246. [Google Scholar] [CrossRef]
  221. Hajiboland, R. Effect of Micronutrient Deficiencies on Plants Stress Responses. In Abiotic Stress Responses in Plants; Ahmad, P., Prasad, M.N.V., Eds.; Springer: New York, NY, USA, 2012; pp. 281–330. [Google Scholar]
  222. Silva, E.C.; Nogueira, R.J.M.C.; Silva, M.A.; Albuquerque, M. Drought stress and plant nutrition. Plant Stress 2011, 5, 32–41. [Google Scholar]
  223. Ahanger, M.A.; MoradTalab, N.; Fathi Abd-Allah, E.; Ahmad, P.; Hajiboland, R. Plant growth under drought stress: Significance of mineral nutrients. Water Stress Crop Plants Sustain. Approach 2016, 2, 649–668. [Google Scholar] [CrossRef]
  224. Prasad BV, G.; Chakravorty, S. Performance of Mulches and Micronutrients on Water Use of Broccoli (Brassica oleracea L var. Italica Plenck). Int. J. Curr. Microbiol. App. Sci. 2019, 8, 102–108. [Google Scholar] [CrossRef]
  225. Mohapatra, S.K.; Munsi, P.S.; Mahapatra, P.N. Effect of integrated nutrient management on growth, yield and economics of broccoli (Brassica oleracea L. Var. italica plenck.). Veg. Sci. 2013, 40, 69–72. [Google Scholar]
  226. Hajiboland, R.; Amirazad, H. Drought tolerance in Zn-deficient red cabbage (Brassica oleracea L. var. capitata f. rubra) plants. Hort. Sci. 2010, 37, 88–98. [Google Scholar] [CrossRef] [Green Version]
  227. Han, S.; Tang, N.; Jiang, H.X.; Yang, L.T.; Li, Y.; Chen, L.S. CO2 assimilation, photosystem II photochemistry, carbohydrate metabolism and antioxidant system of citrus leaves in response to boron stress. Plant Sci. 2009, 176, 143–153. [Google Scholar] [CrossRef]
  228. Hajiboland, R.; Farhanghi, F. Effect of low boron supply in turnip plants under drought stress. Biol. Plant. 2011, 55, 775–778. [Google Scholar] [CrossRef]
  229. Abid, M.; Khan, M.M.H.; Kanwal, M.; Sarfraz, M. Boron application mitigates salinity effects in canola (Brassica napus) under calcareous soil conditions. Int. J. Agric. Biol. 2014, 16, 6. [Google Scholar]
  230. Nawaz, F.; Ahmad, R.; Ashraf, M.Y.; Waraich, E.A.; Khan, S.Z. Effect of selenium foliar spray on physiological and biochemical processes and chemical constituents of wheat under drought stress. Ecotoxicol. Environ. Saf. 2015, 113, 191–200. [Google Scholar] [CrossRef]
  231. Pourjafar, L.; Zahedi, H.; Sharghi, Y. Effect of foliar application of nano iron and manganese chelated on yield and yield component of canola (Brassica napus L.) under water deficit stress at different plant growth stages. Agric. Sci. Dig. Res. J. 2016, 36, 172–178. [Google Scholar] [CrossRef] [Green Version]
  232. Siddiqui, M.H.; Mohammad, F.; Khan, M.N. Morphological and physio-biochemical characterization of Brassica juncea L. Czern. & Coss. genotypes under salt stress. J. Plant Interact. 2009, 4, 67–80. [Google Scholar] [CrossRef]
  233. Ashraf, M.P.J.C.; Harris, P.J.C. Potential biochemical indicators of salinity tolerance in plants. Plant Sci. 2004, 166, 3–16. [Google Scholar] [CrossRef]
  234. Hussain, S.; Khalid, M.F.; Hussain, M.; Ali, M.A.; Nawaz, A.; Zakir, I.; Fatima, Z.; Ahmad, S. Role of micronutrients in salt stress tolerance to plants. In Plant Nutrients and Abiotic Stress Tolerance; Springer: Singapore, 2018; pp. 363–376. [Google Scholar] [CrossRef]
  235. Acosta-Motos, J.R.; Ortuño, M.F.; Bernal-Vicente, A.; Diaz-Vivancos, P.; Sanchez-Blanco, M.J.; Hernandez, J.A. Plant responses to salt stress: Adaptive mechanisms. Agronomy 2017, 7, 18. [Google Scholar] [CrossRef] [Green Version]
  236. Feigl, G.; Kumar, D.; Lehotai, N.; Tugyi, N.; Molnár, Á.; Ördög, A.; Szepesi, Á.; Gémes, K.; Laskay, G.; Erdei, L.; et al. Physiological and morphological responses of the root system of Indian mustard (Brassica juncea L. Czern.) and rapeseed (Brassica napus L.) to copper stress. Ecotoxicol. Environ. Saf. 2013, 94, 179–189. [Google Scholar] [CrossRef] [Green Version]
  237. Brunetti, G.; Farrag, K.; Rovira, P.S.; Nigro, F.; Senesi, N. Greenhouse and field studies on Cr, Cu, Pb and Zn phytoextraction by Brassica napus from contaminated soils in the Apulia region, Southern Italy. Geoderma 2011, 160, 517–523. [Google Scholar] [CrossRef]
  238. Purakayastha, T.J.; Viswanath, T.; Bhadraray, S.; Chhonkar, P.K.; Adhikari, P.P.; Suribabu, K. Phytoextraction of zinc, copper, nickel and lead from a contaminated soil by different species of brassica. Int. J. Phytoremediation 2008, 10, 61–72. [Google Scholar] [CrossRef] [PubMed]
  239. Marchiol, L.; Assolari, S.; Sacco, P.; Zerbi, G. Phytoextraction of heavy metals by canola (Brassica napus) and radish (Raphanus sativus) grown on multicontaminated soil. Environ. Pollut. 2004, 132, 21–27. [Google Scholar] [CrossRef]
  240. Weerakoon, S.; Somaratne, S. Phytoextractive potential among mustard (Brassica juncea) genotypes in Sri Lanka. Ceylon J. Sci. (Biol. Sci.) 2010, 38, 85–93. [Google Scholar] [CrossRef] [Green Version]
  241. Jinal, H.N.; Gopi, K.; Prittesh, P.; Kartik, V.P.; Amaresan, N. Phytoextraction of iron from contaminated soils by inoculation of iron-tolerant plant growth-promoting bacteria in Brassica juncea L. Czern. Environ. Sci. Pollut. Res. 2019, 26, 32815–32823. [Google Scholar] [CrossRef]
  242. Hassan, T.U.; Bano, A.; Naz, I. Alleviation of heavy metals toxicity by the application of plant growth promoting rhizobacteria and effects on wheat grown in saline sodic field. Int. J. Phytoremediation 2017, 19, 522–529. [Google Scholar] [CrossRef]
  243. Wollmer, A.C.; Pitann, B.; Mühling, K.H. Timing of waterlogging is crucial for the development of micronutrient deficiencies or toxicities in winter wheat and rapeseed. J. Plant Growth Regul. 2019, 38, 824–830. [Google Scholar] [CrossRef]
  244. Coolong, T.W.; Randle, W.M.; Toler, H.D.; Sams, C.E. Zinc availability in hydroponic culture influences glucosinolate concentrations in Brassica rapa. Hortscience 2004, 39, 84–86. [Google Scholar] [CrossRef] [Green Version]
  245. Bybordi, A.; Mamedov, G. Evaluation of application methods efficiency of zinc and iron for canola (Brassica napus L.). Not. Sci. Biol. 2010, 2, 94–103. [Google Scholar] [CrossRef] [Green Version]
  246. Jankowski, K.J.; Sokólski, M.; Dubis, B.; Krzebietke, S.; Żarczyński, P.; Hulanicki, P.; Hulanicki, P.S. Yield and quality of winter oilseed rape (Brassica napus L.) seeds in response to foliar application of boron. Agric. Food Sci. 2016, 25, 164–176. [Google Scholar] [CrossRef]
  247. Jansen, M.A.; Klem, K.; Robson, T.M.; Urban, O. UV-B-induced morphological changes-an enigma. In UV-B Radiation and Plant Life: Molecular Biology to Ecology; CABI: Boston, MA, USA, 2017; pp. 58–71. [Google Scholar]
  248. Kopsell, D.A.; Sams, C.E. Increases in shoot tissue pigments, glucosinolates, and mineral elements in sprouting broccoli after exposure to short-duration blue light from light emitting diodes. J. Am. Soc. Hortic. Sci. 2013, 138, 31–37. [Google Scholar] [CrossRef] [Green Version]
  249. Craver, J.K.; Gerovac, J.R.; Lopez, R.G.; Kopsell, D.A. Light intensity and quality from sole-source light-emitting diodes impact growth, morphology, and nutrient content of Brassica microgreens. HortScience 2017, 51, 497–503. [Google Scholar] [CrossRef]
  250. Huang, L.; Ye, Z.; Bell, R.W.; Dell, B. Boron nutrition and chilling tolerance of warm climate crop species. Ann. Bot. 2005, 96, 755–767. [Google Scholar] [CrossRef]
  251. Abedi, T.; Pakniyat, H. Antioxidant enzymes changes in response to drought stress in ten cultivars of oilseed rape (Brassica napus L.). Czech J. Genet. Plant Breed. 2010, 46, 27–34. [Google Scholar] [CrossRef] [Green Version]
  252. Olama, V.; Ronaghi, A.; Karimian, N.; Yasrebi, J.; Hamidi, R.; Tavajjoh, M.; Kazemi, M.R. Seed quality and micronutrient contents and translocations in rapeseed (Brassica napus L.) as affected by nitrogen and zinc fertilizers. Arch. Agron. Soil Sci. 2014, 60, 423–435. [Google Scholar] [CrossRef]
  253. Wollmer, A.C.; Pitann, B.; Mühling, K.H. Waterlogging events during stem elongation or flowering affect yield of oilseed rape (Brassica napus L.) but not seed quality. J. Agron. Crop Sci. 2018, 204, 165–174. [Google Scholar] [CrossRef]
  254. Tunçtürk, M.; Tunçtürk, R.; Yildirim, B.; Çiftçi, V. Changes of micronutrients, dry weight and plant development in canola (Brassica napus L.) cultivars under salt stress. Afr. J. Biotechnol. 2011, 10, 3726–3730. [Google Scholar]
  255. Sikka, R.; Nayyar, V. Cadmium accumulation and its effects on uptake of micronutrients in Indian mustard [Brassica juncea (L.) czern.] grown in a loamy sand soil artificially contaminated with cadmium. Commun. Soil Sci. Plant Anal. 2012, 43, 672–688. [Google Scholar] [CrossRef]
  256. Smith, T.E.; Grattan, S.R.; Grieve, C.M.; Poss, J.A.; Läuchli, A.E.; Suarez, D.L. pH dependent salinity-boron interactions impact yield, biomass, evapotranspiration and boron uptake in broccoli (Brassica oleracea L.). Plant Soil 2013, 370, 541–554. [Google Scholar] [CrossRef]
  257. Dimovska, D.; Bogevska, Z.; Iljovski, I.; Zdravkovska, M.; Kunguloski, D.; Atanasova-Pancevska, N. Quantitative and chemical traits in broccoli (Brassica oleracea L. var. italica) grown with the use of microbiological fertilizer. In Proceedings of the 2nd International Balkan Agriculture Congress, Tekirdag, Turkey, 16–18 May 2017; p. 200. [Google Scholar]
  258. Nie, Z.J.; Hu, C.X.; Sun, X.C.; Tan, Q.L.; Liu, H.E. Effects of molybdenum on ascorbate-glutathione cycle metabolism in Chinese cabbage (Brassica campestris L. ssp. pekinensis). Plant Soil 2007, 295, 13–21. [Google Scholar] [CrossRef]
  259. Millaleo, R.; Reyes-Dıaz, M.; Alberdi, M.; Ivanov, A.G.; Krol, M.; Huner NP, A. Excess manganese differentially inhibits photosystem I versus II in Arabidopsis thaliana. J. Exp. Bot. 2013, 64, 343–354. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Effects of abiotic stresses and their consequences on Brassicaceae.
Figure 1. Effects of abiotic stresses and their consequences on Brassicaceae.
Life 13 00211 g001
Figure 2. Role of plant growth-promoting rhizobacteria in Brassica species against abiotic stresses.
Figure 2. Role of plant growth-promoting rhizobacteria in Brassica species against abiotic stresses.
Life 13 00211 g002
Figure 3. Effect of macronutrients (indicated in blue color), silicon (orange color) and micronutrients (green color) on different functions of Brassicaceae crops.
Figure 3. Effect of macronutrients (indicated in blue color), silicon (orange color) and micronutrients (green color) on different functions of Brassicaceae crops.
Life 13 00211 g003
Table 1. Summary of the positive effects of PGPR in mitigating unfavorable abiotic stress conditions in Brassicas (2008–2020).
Table 1. Summary of the positive effects of PGPR in mitigating unfavorable abiotic stress conditions in Brassicas (2008–2020).
CropsAbiotic Stresses Positive Effect of PGPRReference
RadishSalinity Bacillus subtilis, B. atrophaeus and B. spharicus reduced osmotic effects of salinity to improve production.[68]
RadishSalinity Bacillus subtilis and Pseudomonas fluorescens improved morphological and biochemical attributes as well as hormonal levels of plants.[69]
RapeseedDrought Inoculation of rapeseeds with Pseudomonas fluorescens or P. putida improved yield, 1000-grain weight, grains/pod, pods and branches/plant.[70]
RapeseedHeavy metals Use of Bacillus megaterium reduced soil Ni concentrations through the activity of IAA and solubilization of P.[71]
RapeseedSalinity Pseudomonas sp. and Azospirillum brasilense mitigated harmful effects of salinity by increasing leaf water content, activity of antioxidant enzymes, leaf area, osmolyte production, productivity and leaf nutrient concentrations. [72]
RapeseedHeavy metals Use of Pseudomonas sp. A3R3 and Psychrobacter sp. SRS8 reduced Zn toxicity in the soil due to the production of hormones and siderophore activity.[73]
RapeseedHeavy metals Use of Arthrobacter sp., Bacillus altitudinis SrN9, B. megatherium reduced soils cadmium contamination by producing IAA and siderophores.[64]
CabbageSalinity Azotobacter chroococcum minimized salt stress by increasing root development and IAA. [74]
CabbageDrought Inoculation with Bacillus megaterium, Pantoea agglomerans and Brevibacillus choshiensis improved physiology of membrane integrity and increased accumulation of osmolytes, antioxidant enzymes, hormonal production, decreased electrolyte leakage and production of ROS-eliminating enzymes[75]
CanolaSalinityE. cloacae improved tolerance to saline soils by promoting root–shoot growth and increasing production of phytohormones.[76]
TurnipHeavy metals B. megaterium reduced soil contamination with cadmium and lead by the synthesis of IAA and siderophore activity.[77]
TurnipDrought and phytotoxicity of Zn and CuInoculation with Pseudomonas libanensis TR1 and Pseudomonas reactans Ph3R3 reduced phytoremediation of metals polluted soils and increased relative water content by the synthesis of IAA, 1-aminocyclopropane-1-carboxylate deaminase, and siderophore.[78]
Table 2. Summary of the positive effects of N, P, K, S, Mg, Ca and Si fertilization in mitigating abiotic stress conditions in Brassicas (2004–2020).
Table 2. Summary of the positive effects of N, P, K, S, Mg, Ca and Si fertilization in mitigating abiotic stress conditions in Brassicas (2004–2020).
CropsAbiotic StressesMacronutrients and Si Positive EffectsReference
RadishZn toxicity Mg2+ enhanced uptake and translocation of Zn, as well as alleviated Zn toxicity.[164]
MustardSalinity Nitrogen maintains synthesis of proline and ethylene to combat drastic impacts of salinity on photosystem.[94]
Cd toxicity Nutrient solution of Mg fertilization improved shoot growth of B. rapa under cadmium (Cd) toxicity.[145]
Salinity Nitrogen ameliorates salinity effects by improving growth attributes, physio-biochemical attributes (total chlorophyll, water content, stomatal conductance, K / Na ratio, carbonic anhydrase activity and malondialdehyde) and yield attributes (seeds pod−1, pod and yield plant−1).[165]
Cd toxicity Silicon increased photosynthetic pigments and reduced inhibitory effects of Cd on root elongation.[166]
SalinityHigher proline accumulation and photosynthetic efficiency increased plant growth with S fertilization.[167]
Heavy metalsCadmium and lead have negative effects on P, Ca, Mn and Fe content root and leaves dry mass. [168]
Heavy metals Application of Ca increases tolerance to Cd in mustard plants by restoring morphological and biochemical attributes. [137]
As toxicity Silicon modulated root elongation with development of both primary and lateral roots.[169]
Cr toxicity Silicon reduced transportation of Cr from root to shoot and photosynthetic activity by increasing net photosynthetic rate, chlorophyll, and carotenoid content.[170]
Salinity Silicon increased plant growth, antioxidant activity (catalase, peroxidase and superoxide dismutase) and proline content.[171]
Heavy metalsApplication of S mediated antioxidant enzymes in the plant, contributing to phytoextraction of potentially toxic elements (cadmium and zinc) from contaminated soils, helping in phytoremediation process of the soil.[172]
RapeseedDrought Potassium fertilization improved relative water content, stomatal conductance, relative chlorophyll index, and productivity.[173]
Salinity Silicon nutrition ameliorated the lethal impacts of salinization in canola by lowering Na absorption, maintaining root cell integrity, reduced lipid peroxidation, enhancing the scavenging capability of ROS and decreased lignification.[174]
Drought Fertilization with K2SO4 alleviated deleterious effects of water stress by stimulating productive characteristics (pods plant−1, seeds pod−1 and grain yield).[130]
DroughtNitrogen improved proline production to maintain water balance and integrity of proteins, enzymes and cell membranes.[107]
Oxidative stress Nitrogen and S reducing reactive oxygen species production.[175]
Salinity Silicon application prevented toxic ions (Na and Cl) accumulation while maintaining K, P and Fe content in plants. [176]
DroughtSilicon increased shoot–root biomass, total chlorophyll content, activity of superoxide dismutase and catalase while reducing lipid peroxidation.[177]
Salinity Phosphate fertilizers improved plant performance under salt stress by lowering Na+/K+ ratio and increasing P uptake.[114]
DroughtNitrogen increased plant height, number of branches, number of fruits per plant, thousand seed weights and crude protein. [178]
Cd toxicity Silicon reduced oxidative damage in plants by increasing antioxidant components and methylglyoxal detoxification system that enhance tolerance to Cd stress.[179]
Drought Silicon improved antioxidants enzymes, ascorbate and glutathione pool, glyoxalase systems and proline by increasing protective role and maintaining redox status of plants.[126]
Oxidative stress Silicon improved biomass, N uptake and chlorophyll content. Also, decreased oxidative stress by reducing hydrogen peroxide and malondialdehyde production.[180]
Haze conditionNitrogen increased shoot biomass and photosynthetic productivity.[181]
CanolaSalinityCalcium-fortified composted animal manure alleviate oxidative stress, improvement in growth, physiology and mineral nutrition [182]
Salinity Increased activity of phosphatase enzymes and reduced phosphate levels in plants.[183]
SalinityPotassium fertilization mitigates the effects of salinity by confining Na absorption, activating cellular compartmentalization of excess Na+ in cell vacuole.[129]
Heavy metals Sulfur application increases lipid peroxidation and activities of antioxidant enzymes.[184]
Drought Fertilization of Ca allows plants to resist drought by improving antioxidant capacity, oil quality and essential fatty acids (linolenic acid and linolenic acid) in seeds.[185]
Drought Potassium mitigated the effect of water deficiency by increasing water and nitrogen use efficiency, improving chlorophyll index, leaf area index, cell membrane integrity and productivity.[186]
CabbageDrought Nitrogen increased harvest index and dry matter production.[187]
Cd toxicitySilicon alleviated Cd toxicity by increasing activities of antioxidant enzymes and shoot and root biomass.[188]
Salinity Potassium fertilization improved absorption of total soluble free amino acids and proteins, proline content, regulated activities of antioxidant and improved gas exchange traits. [189]
BroccoliHeavy metals Calcium fertilization mitigates ZnSO4 toxicity by increasing total phenolic content and antioxidant capacity of sprouts.[190]
Salinity Nitrogen increased photosynthetic capacity and vitamin C content.[191]
NH4+ toxicity Silicon alleviated NH4+ toxicity in cauliflower by increasing physical integrity of membranes while increasing water use efficiency in broccoli.[192]
Drought Co-application of macro- and micronutrient and biostimulants increased nutritional status of broccoli plants in water deficient conditions.[193]
High luminosityA positive correlation between Fe, Mg and Ca, and high light was observed.[194]
ArugulaDrought Potassium mitigated the effect of water deficiency by increasing water and nitrogen use efficiency, improving chlorophyll index, leaf area index, cell membrane integrity and productivity.[195]
DroughtSilicon improved gas exchanges capacity.[196]
KaleDrought Silicon reduced water loss, increased shoot biomass and plant height.[197]
TurnipHeavy metals Polypeptide Ca has a dual function in competitive inhibition in cadmium-contaminated agricultural land.[198]
Table 3. Summary of the positive effects of micronutrients fertilization in mitigating unfavorable abiotic stress conditions in Brassicas (2005–2020).
Table 3. Summary of the positive effects of micronutrients fertilization in mitigating unfavorable abiotic stress conditions in Brassicas (2005–2020).
Crops Abiotic Stresses Micronutrients EffectsReference
RapeseedCold temperature and high light radiationBoron removal, mobilization and partitioning into root–shoot and younger leaves of plant were imperatively reduced in chilling temperature and intensive light.[250]
Drought Drought drastically reduced Zn contents and led to photosynthetic damages with alternative reduction in transpiration, net assimilation and stomatal conductance and act as Cu-Zn SOD enzyme.[251]
Salinity Protein and micronutrients were improved in order of Mn > Zn > Cu > Fe in aerial parts with application of N and Zn.[252]
WaterloggingWaterlogging severely impaired growth and nutrients accumulation and ATP synthesis in plants.[253]
WaterloggingEarly waterlogging stress resulted in higher accumulation of Mn, Fe, Zn and Cu in the leaves and causes toxicity.[243]
TurnipDrought Boron deficiency is increased under drought stress and led to the disturbance of electron transportation in photosynthesis, lowering chlorophyll content and root–shoot ratio.[228]
Drought Drought drastically reduced Zn contents and led to photosynthetic damages with alternative reduction in transpiration, net assimilation and stomatal conductance.[226]
CanolaSalinity Micronutrients (Fe, Mn and Cu) contents in plant aerial parts were improved under salt stress [254]
Drought Boron deficiency is increased under drought stress and led to the disturbance of electron transportation in photosynthesis, lowering chlorophyll content and root–shoot ratio.[229]
Drought Yield and yield components were improved with lower dose of foliar Fe and Mn.[231]
MustardHeavy metals Accumulate high amount of Cd, Cu, Ni, Cr, Zn, Fe, Co, Pb and Se from contaminated sites.[241]
Heavy metals High levels of Cd decrease micronutrients (Mn, Fe, Cu and Zn) content. [255]
BroccoliHigh light radiationShort duration of blue light-emitting diodes (LED) prominently improves phytochemical components, essential micronutrient (B, Fe, Zn, Mn, Mo, Na and Cu) and macronutrients (Ca, P, K, Mg and S).[248]
SalinitySalinity reduced yield and boron accumulation in aerial parts of plant.[256]
Hight light radiationShort duration blue light enhanced different phytochemical activities, micronutrients (Zn, Mn, Mo, B, Na, Fe and Cu) concentration and also macronutrients (Ca, P, K, S and Mg) concentration in plants.[248,249]
Salinity Biofertilizers improve Fe availability and also Ca and Mg content in plants.[257]
Chinese cabbageCold temperatureMolybdenum promotes antioxidant and phytochemical activities, improve growth, quality and yield.[258]
A. thalianaHigh light radiationZinc prevents photo-inhibitory damages to photosynthetic apparatus by producing ROS and enhancing carotenoid contents plant leaves.[259]
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

Jalal, A.; da Silva Oliveira, C.E.; Galindo, F.S.; Rosa, P.A.L.; Gato, I.M.B.; de Lima, B.H.; Teixeira Filho, M.C.M. Regulatory Mechanisms of Plant Growth-Promoting Rhizobacteria and Plant Nutrition against Abiotic Stresses in Brassicaceae Family. Life 2023, 13, 211. https://doi.org/10.3390/life13010211

AMA Style

Jalal A, da Silva Oliveira CE, Galindo FS, Rosa PAL, Gato IMB, de Lima BH, Teixeira Filho MCM. Regulatory Mechanisms of Plant Growth-Promoting Rhizobacteria and Plant Nutrition against Abiotic Stresses in Brassicaceae Family. Life. 2023; 13(1):211. https://doi.org/10.3390/life13010211

Chicago/Turabian Style

Jalal, Arshad, Carlos Eduardo da Silva Oliveira, Fernando Shintate Galindo, Poliana Aparecida Leonel Rosa, Isabela Martins Bueno Gato, Bruno Horschut de Lima, and Marcelo Carvalho Minhoto Teixeira Filho. 2023. "Regulatory Mechanisms of Plant Growth-Promoting Rhizobacteria and Plant Nutrition against Abiotic Stresses in Brassicaceae Family" Life 13, no. 1: 211. https://doi.org/10.3390/life13010211

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