Next Article in Journal
Deep Soil Water Availability Regulates the Transpiration of Afforested Apple Trees (Malus pumila Mill.) in a Sub-Humid Loess Region
Previous Article in Journal
Prediction Models of Growth Characteristics and Yield for Chinese Winter Wheat Based on Machine Learning
Previous Article in Special Issue
The Effects of Brassinosteroids on Nitrogen Utilization in Rice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Brassinosteroids: Relevant Evidence Related to Mitigation of Abiotic and Biotic Stresses in Plants

by
Aminthia Pombo Sudré da Silva
1,
Antônio André da Silva Alencar
2,
Cláudia Pombo Sudré
2,
Maria do Socorro Bezerra de Araújo
3,* and
Allan Klynger da Silva Lobato
4
1
Instituto Federal Fluminense—IFF Cabo Frio, Cabo Frio 28909-971, RJ, Brazil
2
Laboratório de Melhoramento Genético Vegetal, Universidade Estadual do Norte Fluminense Darcy Ribeiro, Campos dos Goytacazes 28013-602, RJ, Brazil
3
Faculdade de Ciências Biológicas e Agrárias, Universidade do Estado de Mato Grosso—Cidade Universitária, Av. Santos Dumont, s/n- Santos Dumont, Cáceres 78211-332, MT, Brazil
4
Núcleo de Pesquisa Vegetal Básica e Aplicada, Universidade Federal Rural da Amazônia, Paragominas 68627-451, PA, Brazil
*
Author to whom correspondence should be addressed.
Agronomy 2024, 14(4), 840; https://doi.org/10.3390/agronomy14040840
Submission received: 1 February 2024 / Revised: 3 April 2024 / Accepted: 4 April 2024 / Published: 17 April 2024

Abstract

:
Extreme events of climate change are increasing, such as droughts and heat waves, causing limitations on growth and yield in relevant food crops, as well as threatening global food security. Brassinosteroids (BRs) are natural or synthetic steroids with significant properties that promote plant growth and development. In the current world scenario, research and solutions that can improve plant tolerance to climate change are strategic to ensure food security. The distinctiveness and novelty of this review lie in its comprehensive and detailed approach to the role of BRs in plants under biotic and abiotic stresses. We consolidate information on the action mechanisms on specific organs, providing detailed experimental conclusions of these plant growth regulators, including also commercial products and concentrations tested aiming to mitigate the adverse effects of the stresses. This practical approach highlights the potential of BRs in agriculture and plant protection against stresses. Additionally, our review presents results with plant models and essential food crops, focusing on multidisciplinary approaches and using physiological, biochemical, nutritional, anatomical and agronomic tools to explain the mechanisms of action of brassinosteroids in plants exposed to abiotic and biotic stresses.

1. Introduction

On one of the first pages of the book Phytohormones [1], the authors cite an old lullaby: Oats, peas, beans and barley grow, Oats, peas, beans and barley grow, Can you, or I, or anyone know how oats, peas, beans and barley grow? How plants grow is a question that permeates the imagination of children and researchers. Much progress has been made since the knowledge that phytohormones were involved in several physiological processes related to growth, such as etiolation, elongation, germination, etc.
Plant growth regulators (PGRs) are signalling molecules involved in plant growth and development. The regulation of cellular and developmental processes can be determined by one or a combination of molecules [2]. The most studied and relevant PGRs in growth and development processes are auxin, cytokinin, ethylene, abscisic acid, gibberellin, brassinosteroid, salicylic acid, jasmonic acid, strigolactone and polyamines [3]. Robust equipment and new software for detecting and identifying molecules increase the accuracy, with new PGRs being studied in terms of plant × molecule × environment interaction. These molecules can be produced in plant organs and influence a specific physiological process when transferred to other parts. Thus, they are characterised as chemical messengers responsible for coordinating certain physiological activities [1]. Among the main activities of PGRs are the regulation of developmental processes and signalling networks involved in plant responses to various biotic and abiotic stresses [4,5,6], such as diseases caused by fungi, bacteria, viruses and nematodes [7] and drought, thermal stress, salt stress and potentially toxic elements [8]. This review presents a comprehensive and detailed approach to the role of brassinosteroids (BRs) in plants under biotic and abiotic stresses. We consolidate information on the action mechanisms of specific organs, providing detailed experimental conclusions of these PGRs, including commercial products and concentrations tested that are aiming to mitigate the adverse effects of the stresses. This practical approach highlights the potential of brassinosteroids in agriculture and plant protection against stresses.

2. Brassinosteroids

BRs are considered PGRs. In 1970, Mitchell and collaborators published in Nature the paper “Brassins—a New Family of Plant Hormones from Rape Pollen”. In this research, when testing an oily extract, “Brassin” of Brassica Napus pollen, on bean seedlings at deficient concentrations (10 μg plant−1), researchers observed the superior elongation of the internodes compared to the seedlings treated with controls, gibberellin and fractionated lanolin. The authors named these substances Brassins. Seedlings treated with these substances were histologically differentiated from those treated with the control. According to the chemical analyses, the authors suggested that Brassins would be a new family of PGRs, since they had a glyceride structure and differentiated action from gibberellin [9].
In 1979, the first brassinosteroid was isolated, brassinolide (BL), a lactone steroid, which is the most active [10]. In the 1990s, mutants that did not biosynthesise BRs in some species were identified. The isolation of the Brassinosteroid Insensitive 1 (BRI1) gene, associated with molecular and biochemical studies, allowed the demonstration that BRI1 is a ubiquitously expressed leucine-rich receptor kinase, which, through the phosphorylation of serine and threonine, acts in the signalling of BRs [11]. This receptor detects extracellular BRs and, as needed, triggers an intracellular signalling cascade [12]. These studies recognised these compounds as a new class of PGR. These substances gained attention from the scientific community, and many metabolic processes were understood. By 2017, 81 natural BRs, five conjugates, 137 analogues and eight metabolites isolated from 64 species, including angiosperms, gymnosperms, pteridophytes, bryophytes and algae, had already been characterised [13,14], whose division is, according to the B ring and the orientation of the hydroxyl, ketone and epoxide groups, at the C-1, C-2, C-3 and C-6 position on the A ring [15]. In addition to their chemical structure, these brassinosteroids differ in physiological activity [16].
BRs are found in small amounts in all plant organs. The concentration is higher in young tissues, varying between 1 and 100 ng g−1 of fresh mass. However, mature tissues have a marked reduction between 0.01 and 0.1 ng g−1 [17]. These hormones can act on metabolism, resulting in a greater tolerance to biotic and abiotic stresses, cell expansion and elongation, redox signalling and several other functions.
BRs are polyhydroxylated phytosterols synthesised from the sterols campesterol, sitosterol and cholesterol, which are metabolised in many intermediates in plant cells. Still, few of these metabolites have biological activity [15], BRs were defined as steroids with an oxygen fraction at C-3 and additional ones at one or more carbon atoms C-2, C-6, C-22 and C-23. All BRs can fall into three categories according to the carbon number of each steroid molecule: C27, C28 and C29. The biosynthesis of BRs (C27, C28 or C29) occurs through three biochemical pathways. The first phase, which results in the synthesis of sterols, can be the mevalonate or non-mevalonate pathway. Subsequently, follow cycloartenol and cycloartenol-dependent pathways. Various pathways lead to the synthesis of cholesterol and campesterol from cycloartenol, while sitosterol also can be synthesised from cycloartenol [18]. Campesterol and 24-β-methylcholesterol are precursors of C28 BRs, C27 BR’s cholesterol and C29 BR’s sitosterol [19]. Not all the biosynthetic compounds of all BRs have been deciphered, but there are already commercially available synthetic BRs. BRs can act on metabolism, resulting in greater tolerance to biotic and abiotic stresses, cell expansion and elongation, redox signalling and several other functions (Figure 1).

3. Abiotic Stresses

Similar metabolic processes may result from different stresses, causing equal cellular damage. Osmotic stress disrupts homeostasis and results in cellular electrolyte leakage (EL), whereas oxidative stress can denature functional and structural proteins. Both stresses generally activate similar signalling pathways and cellular responses, such as the production of stress proteins. When subjected to stress, the plant reacts with cellular biochemical changes to adapt to the new environmental condition, including new metabolic pathways and changes in the number of phytohormones [20].
Reactive oxygen species (ROS) play a key role in plant growth, development and interaction with biotic and abiotic stresses. ROS, constantly produced during photosynthesis and respiration, are essential regulatory and signalling elements in various cellular processes. Disruption of the protective mechanisms of redox homeostasis in the cell can cause oxidative stress, resulting in the increased production of ROS (O2, H2O2, −OH, NO), leading to oxidative damage and cell death. Water deficit, flooding, desiccation, salinisation, cooling, thermal shock, heavy metal poisoning, ultraviolet radiation, ozone, mechanical damage, nutritional deficit, the action of phytopathogenic microorganisms and excessive luminosity are stresses that cause disturbances in cellular homeostasis [21].
Antioxidant enzymes like superoxide dismutase (SOD), catalase (CAT), glutathione peroxidase (GPX) and ascorbate peroxidase (APX) and non-enzymatic antioxidants like ascorbic acid, tocopherols, carotenoids, glutathione, etc. in plants under different stress conditions are modified by exogenous application of BRs. ROS interact directly or via enzymatic catalysis with vitamins C, E and glutathione, whereas carotenes inhibit intermediate forms of ROS [22,23,24,25,26].
The increased tolerance of plants under abiotic stresses (heat, cold, drought, salinity, metal toxicity, etc.) with the action of BRs has been proven using several phenotypic, anatomical, physiological, biochemical and molecular studies, which demonstrate how these stresses cause damage and how BRs help plants in the dynamic defence response to such pressures (Table 1).
For each type of response, factors such as time, BR concentration, plant species, application method and type of stress may vary. Biochemical reactions are faster than phenotypic responses, for example. BRs respond to all stresses. However, they are susceptible to injuries and wounds, implying application methods without mechanical damage, such as seed soaking, substrate application, root or fruit immersion and foliar spraying (Table 1). According to [94], providing BRS between 14 and 21 days before exposure to stress via roots is the method that generates the most reproducible results.

3.1. Thermal Stress

With the effects of climate change, the use of substances that increase stress tolerance to extreme temperatures is increasingly necessary. Temperature stress comprises extreme values of both low and high temperatures. BRs most studied in increasing tolerance to heat stress are 24-epibrassinolide (24-epiBL), 28-Homobrassinolide (28-homoBL) and BL (Table 1). The concentrations of the tested BR solutions vary between 0.01 and 10 µM. Different modes of application of BRs are reported, for example, needleless foliar injection, foliar spraying and fruit immersion, in at least six studied species (Camellia sinensis, B. napus, Solanum lycopersicum, Capsicum annuum, Vigna radiata and Cucumis sativus). Some of the described effects of BRs on thermotolerance are increased carboxylation efficiency and improved antioxidant enzyme systems in leaves [62], increase in the endogenous concentration of abscisic acid (ABA) in leaves of B. napus [61], protective effect against membrane leakage and degradation of photosynthetic pigments [66]; increased levels of partial free amino acids (proline, arginine, aspartic acid and glycine) [37]; increased nitrate reductase (NR), glutamine synthase (GS), glutamate synthase (GOGAT) and glutamate dehydrogenase (GDH) activities in sweet pepper seedling leaves and reduced H2O2 e O2· accumulation [37]; reduced electrolyte leakage (EL) and malondialdehyde (MDA) leakage content and increased proline content [63] (Table 1).
BRs increase thermotolerance by reducing, for example, the effects of cold stress, such as photoinhibition caused by damage to chloroplast thylakoid membranes and the inactivation of enzymes involved in CO2 assimilation [95], which indirectly leads to oxidative damage to DNA, proteins and lipids [96]. In addition, BRs alter the physicochemical properties of the membrane, regulate the expression of stress-responsive genes, and have indirect effects on metabolism through other hormones or signalling molecules (such as H2O2) [16].
Cold and drought stress response genes (COR47 and COR78) and heat (HSP83, HSP70, HSF3, Hsc70-3 and Hsc70-G7), related to oxidative stress that encodes monodehydroascorbate reductase and thioredoxin, were identified in mutant plants of BR-deficient Arabidopsis treated with BR [97]. The BR-biosynthetic gene DWARF4 from transgenic Arabidopsis seeds evidences the role of BRs in cold stress tolerance [98]. Photoprotective pathways and hormone regulation can be regulated by H2O2 in the apoplast acting as a signalling molecule, and photoinhibition was probably reduced in tomato due to the abundance of D1 protein, derived from the application of EBR (0.2 μM) [95]. The photosystem II reaction centre (PSII) comprises proteins D1 and D2, because the damage caused by ROS on D1 and D2 proteins reduce carbon fixation and oxygen evolution and disrupts the flow of electrons [95,99].
In A. thaliana, cold tolerance and freezing mediated by BRs have already been studied in cell suspension [100] and seedling spraying before exposure to low temperatures [101,102], as well as the immersion of eggplant (Solanum melongena) fruits in BR solution before cold storage [103]. Grape seedlings sprayed with 0.5 μM 24-epiBL and subsequently subjected to cold (0 °C/36 h) had lower relative EL and higher maximum quantum yield of photosystem II—PSII (Fv/Fm) [67]. Seedlings of B. napus and A. thaliana were cultured in Murashige and Skoog (MS) medium containing one μmol/L of 24-epiBL for 14 or 21 days before being subjected to heat stress (45 °C) [98]. Pre-treatment with BR increased the relative survival rate, the accumulation of heat shock proteins (HSPs) in B. napus [104] and the level of ABA compared to untreated A. thaliana seedlings [105]. Positive effects on damage caused by stress at high temperatures by the application of BRs have already been detected in grass (Bromus inermis) subjected to 40–45 °C [106], in C. sinensis at 35 °C [106], in B. napus at 45 °C [61] and tomato in a 12 h photoperiod at 40–30 °C [62]. In tomatoes under high temperatures (40 °C), applying 24-epiBL on leaf discs increased the activity of CAT, peroxidase—POD and SOD [107]. The benefits of using BRs under cold stress have also been proven for Oryza sativa [108,109], Z. mays [110], C. sativus [111] and Musa paradisiaca [112].

3.2. Drought

Water, and consequently the hydric balance, is essential for the survival and development of the plant since more than 90% of its green mass is made up of water [113]. Gas exchange, photosynthesis, absorption, and transport of nutrients are some factors that water mediates. Water stress causes biochemical, anatomical and physiological changes. BRs 24-epiBL, 28-homoBL, BL, and Spirostanic analogue of brassinosteroid (BB16) showed positive responses when applied at doses ranging from 0.01 to 5 μmol/L in cultures of S. lycopersicum, Phaseolus vulgaris, B. napus, C. annuum, C. sativus, Raphanus sativus, Carthamus tinctorius, Glycine max, Triticum aestivum, Vitis vinifera, Brassica juncea, Arabidopsis, Sorghum vulgare, Arachis hypogaea, Robinia pseudoacacia, Carica papaya and Xanthoceras sorbifolia. Among the effects of BRs that alleviate water stress are greater CO2 assimilation, reduced photoinhibition, increased efficiency in light utilisation and dissipation of excitation energy in PSII in leaves, increased concentration of ABA and antioxidant enzymes, carotenoids and proline (Table 1). Lower net photosynthetic rates and lower accumulation of osmoprotectants were detected in barley allelic mutants (Hordeum vulgare) with deficient enzymatic activity (BR-biosynthetic C6-oxidase) related to drought tolerance [114]. The exogenous use of BL increased the activities of SOD, CAT, APX, and the levels of vitamin C and carotenoids in corn seedlings under water deficit [22].
Waterlogging stress limits gas exchange and subjects the root system to low oxygen (O2) conditions. These define the functions and growth of the roots and consequently affect the entire development of all organs, which can lead to the death of the plant. Corn seedlings subjected to waterlogging stress, previously treated with BL, showed an improvement in the antioxidant system, regulating the formation of aerenchyma, with consequent maintenance of the general vital system of the plant [59,115] induced the condition of hypoxia in mutants of A. thaliana det2, blocked the synthesis of BRs, and observed anatomical changes, such as thicker layers of the cuticle; however, when adding BL, the anatomical characteristics were similar to the control (Cv. Columbia). The authors presume that BRs can mediate plant developmental responses under hypoxic conditions because BRs, in their biosynthesis, require molecular oxygen in several steps.

3.3. Salt Stress

Salt stress is one of the most severe and essential stresses. In 2002, FAO estimated that 1 to 2% of the arable land on the planet would be unfeasible for agriculture per year due to salinity or sodicity [116]. Approximately half of the irrigated area in the world has or has a problem with salinisation and soil saturation [117]. The exogenous application or endogenous stimulation of BRs minimises the effects of salinity on plants in these soils and their respective consequences, including the prevention of secondary problems such as diseases (Table 1).
Most articles focused on studying the effect of BRs on salinity tolerance, but the induction of salt stress is mainly due to sodium chloride (NaCl) (Table 1). However, MgSO4 and CaCl2 have also been reported. In the literature, exogenous applications of BRs (24-epiBL; 28-homoBL and BL) are reported to mitigate salt stress with concentrations of 10−11 to 10−1 M. The main effects of the application of these hormones by different methods, such as seed soaking, foliar spraying and substrate application, were: higher levels of nutrients in the tissues [74,87,92], higher levels of photosynthetic pigments and photochemical efficiency [74,81]; more excellent activity of enzymes and antioxidant compounds (CAT, POD, SOD) and non-toxic osmoprotectors, such as proline, glycine, betaine, mannitol and trehalose [35,74,81]; increased activity of the ROS elimination system; reduction in MDA content [70]; improvement in anatomical attributes [71]; removal of electrolyte leakage [118,119], among others. It is worth mentioning that different effects were observed for the maturation stages of the plant, for the cultivars of the same species, for the salt concentration and for the attention and form of application of the BRs. These data were analysed for 23 plant species (Table 1). CAT, SOD and glutathione reductase (GR) activities increased in rice seedlings subjected to salt stress (NaCl) and treated with the BB16 analogue [23]. Seedlings of salinity-sensitive varieties treated with 24-epiBL had partial and short-term improvement in salinity tolerance due to a sustained higher activity of APX under these conditions [24].

3.4. Potentially Toxic Elements

Potentially toxic elements (PTEs) can cause stress in the metabolic pathways of plants, although some cultivars are used for bioremediation. High concentrations of PTM can cause changes in the chemical structure of molecules, such as the blocking of functional groups (−OH, −COOH, −C=O) or displacement or replacement of essential ions, which can lead to regulatory changes such as chlorosis and nutritional imbalance [120].
The BRs most used in studies of attenuation of the effects of intoxication by PTM are 24-epiBL, 28-homoBL and epiBL with concentrations from 0.01 nM to 5 μM. The most studied metal regarding the interaction with BRs was cadmium. Other metals are mentioned, such as Al, Cu, Fe, Ni, Pb and Zn, in 11 plant species (R. sativus, Solanum nigrum, S. lycopersicum, C. sativus, O. sativa, Triticum aestivum, Vigna radiata, B. juncea, B. napus, Cicer arietinum, Phaseolus vulgaris) (Table 1). In these studies, the BRs were applied in a culture medium by foliar spraying, seed imbibition and application in the root whose effects were: 24-epiBL and 28-homoBL caused additional stimulation of antioxidant enzymes and proline content, elevation of the proline level in association with the antioxidant system that, at least in part, was responsible for the improvement of Al stress in mung bean (V. radiata) seedlings [27]. Antioxidant enzyme activities (CAT, POD and SOD) and proline contents increased in control, regardless of treatment [121]. The analogue 24-epiBL reduced the toxic effect of Cd on photochemical processes, decreasing damage to reaction centres and O2 evolving photochemical centres and maintaining efficient photosynthetic electron transport [32].

3.5. Abiotic Disorders

One of the central abiotic diseases is blossom-end rot (BER), which affects several crops, mainly Solanaceae. The low Ca2+ content in the distal tissue of the fruit is the main trigger for this physiological disorder. The imbalance of this ion leads to membrane rupture, accumulation of ROS and cell death, serving as a gateway to secondary diseases. In the ‘BRS Montese’ tomato, the application of 24-epiBL (0.01 µM) reduced the incidence of BER by 44.2% and increased fruit mass and size. Fruits showed physiological improvement by increasing the concentration of ascorbic acid, the activity of antioxidant enzymes (APX, CAT and SOD) and the concentration of soluble Ca2+, necessary for the structure and stability of the membrane, resulting in increased tolerance to BER [26]. Calcium is also essential in the defence induced by BRs.

3.6. Crosstalk between Brassinosteroids and Other Plant Growth Regulators

The regulation of plant development against biotic and abiotic factors is made possible by the interaction between PGRs and biochemical pathways triggered by external stimulation. The connections between the signalling pathways are integrated by a network whose products are BRs, auxin, ethylene, jasmonic acid (JA) and salicylic acid (SA). There are synergistic effects of BRs with other PGRs, such as increased ethylene production in mung bean epicotyl segments [122] and with auxin on stem segment elongation [123,124]. Brassinosteroids enhanced expression proteins (BEE1, BEE2 and BEE3), and the opposite effect was observed; when BEE1 was overexpressed, there was a reduction in ABA accumulation, suggesting that this protein may function as a signalling intermediary in multiple pathways. In sorghum (Sorghum vulgare), ABA and BR synergistic effects on increasing drought tolerance were observed [125]. There is a potential interaction between BR and JA biosynthesis, reinforcing the role of BRs in mitigating the effects caused by abiotic and biotic stresses [59,115].
The synergistic effects of BRs and the BR-INSENSITIVE 2 (BIN2) inhibitor bikinin lead to increased physiological responses mediated by SA, such as resistance to Pseudomonas syringae pv. tomato DC3000 in Arabidopsis. SA is an elicitor that regulates several immune responses in plants, including pathogen-associated molecular pattern (PAMP)—triggered immunity (PTI), effector-triggered immunity (ETI) and systemic acquired resistance (SAR). Brassinosteroids enhance SA-mediated immune responses in Arabidopsis by inactivating BIN2, which inhibits clade I TGACG motif-binding transcription factors (TGAs). Only TGA1 and TGA4 interact with BIN2, resulting in changes. In TGA4, BIN2 phosphorylates serine at position 202, suppressing the redox-dependent interaction between the non-expressor of pathogenesis-related genes (NPRs) and TGA4 and destabilising the latter. Plant defence responses are regulated by BRs, not exclusively via crosstalk with the PTI signature but also with hormonal interaction with SA [126]. Gibberellic acid inhibitor proteins (DELLA) and the BR-regulated transcription factor BZR1 are central hubs for crosstalk and signal integration, playing ambivalent roles in plant innate immunity, microbial virulence, and disease resistance [127].

4. Biotic Stresses

Phytopathogenic microorganisms penetrate plants through natural openings (stomata, lenticels and hydathodes) or lesions (artificial, natural and by the action of fungi). Phytopathogenic bacteria and viruses do not have active structures that allow penetration into plants by mechanical force. Insects and mites usually spread viruses by pores that extend along the cell wall. Phytopathogenic fungi may have different structures that also allow direct penetration into the host, including direct penetration at the subcuticular level and the cellular level (with haustoria, intercellular mycelium, intercellular mycelium with haustoria or appressorium and intracellular mycelium) [128].
In the interaction process between plant and pathogen, there is bidirectional communication. Plants can recognise foreign organisms and activate their defence mechanisms. On the other hand, the pathogen can manipulate the plant’s biology to create an ideal environment for its growth and development by avoiding a plant defence response. In this relationship of attack and defence, the pathogen and the host can produce a series of toxic chemical compounds. The symptoms resulting from the interference of these compounds in the host physiology are indicators of its susceptibility or resistance. Structural and biochemical mechanisms, both pre-formed (passive or constitutive) and post-formed (active or induced), are the means of plant defence against the attempted infection and colonisation of the pathogen [129].
Structural mechanisms are physical barriers that inhibit or hinder the pathogen’s penetration and/or colonisation. Constitutive physical barriers include cuticles (waxes and cutin), cell walls, trichomes, stomata, etc. Induced physical barriers include cellular defence structures, lignification, papillae, halos, cork layers, abscission layers, tyloses and glycoproteins rich in the amino acids hydroxyproline and glycine. Biochemical processes involve substances that can inhibit the development of the pathogen or generate adverse conditions for its survival in the tissues of the survivor. The constitutive biochemical apparatus includes phenolic compounds, saponins, unsaturated lactones, glycosides, phytotoxins, antimicrobial proteins and peptides. The activated biochemical substances include phytoalexins, ROS and proteins related to pathogenesis (PR), which can also pre-exist in the plant [130]. In recent decades, phytohormones have stood out in plant–pathogen interactions as a biochemical defence. Plant responses to pathogen infection are coordinated through integrated signalling, communication and response systems. The transmission of signals between cells is essential for the life of an organism, as it provides the control by which cells perceive, recognise and respond to external stimuli [131].
The main chemical mechanisms of pathogen attacks on the host are enzymes, toxins, and hormones. The pathogen secretes these chemical weapons to penetrate, colonise and/or provide nutrients in the host cell. Enzymes are essential in all these stages of pathogenesis and act in the disintegration of cellular components and substances present in host cells. Toxins are highly mobile molecules, active at deficient concentrations (10−6 to 10−8 M), capable of causing physiological, metabolic or structural changes. Their sites of action are the plasma membrane, mitochondria, chloroplasts and enzymes, which are essential for host defence. Some phytopathogenic agents synthesise hormones in the same way as those of the host and other different hormones. This pathogen attack mechanism causes a hormonal imbalance, resulting in irregular growth and development and accelerating plant tissue senescence [132].
Salicylic and jasmonic acids and ethylene are the main hormones involved in the response of plants under biotic stress [133]. Abscisic acid, auxin, gibberellic acid, cytokinin, brassinosteroids and peptide hormones also act in plant defence signalling pathways in different pathosystems [134]. Depending on the type of plant–pathogen interaction, various hormones play positive roles (when the hormone can act against the disease), negative (when the occurrence of the disease is favoured) or neutral (when there is no influence of the hormone). In mango fruits, it was observed that resistance to Colletotrichum gloeosporioides resulted from the involvement of the up-regulation of ET, BRs and phenylpropanoid pathways [135].
The reprogramming of host metabolism, gene expression, and modulation of plant defence responses against pathogen attacks can occur through treatment with plant hormones. BRs regulate other hormones and are involved in plant–pathogen interactions, regardless of whether they are biotrophic, hemibiotrophic or necrotrophic [134]. When applied exogenously, they play an essential role in the innate immunity of plants [127].

4.1. Oomycetes, Bacteria, Fungi and Virus

BR concentrations, modes of application, and plant development stages must be well-defined to control diseases efficiently using these hormones. Inadequate concentrations, for example, can stimulate the growth of the pathogen and, consequently, the progression of the disease [136]. The root oomycete Pythium graminicola, which causes root rot in rice, uses BRs as virulence factors to cause disease; in this case, BRs favour the pathogen, causing greater plant susceptibility [137]
Treatments with 24-epiBL and a formulation based on a brassinosteroid analogue DI-31 (BB16) activated the innate immunity of strawberry plants (Fragaria spp.) in response to infection by Colletotrichum acutatum. Both steroids were more effective in inducing the resistance response at the lowest concentration (0.1 mg/L), with the BB16 analogue causing a more substantial effect than 24-epiBL. Defence biochemical markers showed an increase in the production of H2O2, O2·, NO, calcium oxalate crystals and greater callose and lignin deposition in plants treated with 24-epiBL and BB16 compared to control plants. It was also observed that in plants treated with BB16, 40% of the stomata closed six hours after treatment, while 24-epiBL did not cause any change in stomata opening [138].
The ability of BRs to induce resistance to more than one disease has been observed in rice and tobacco plants. In rice plants, BL treatment induced resistance to rice blast, caused by the fungus Magnaporthe grisea, and bacterial blight, caused by Xanthomonas oryzae pv. oryzae. BL provided excellent resistance to the tobacco mosaic virus (TMV) and Pseudomonas syringae pv. tabaci in tobacco (bacteria) and Oidium sp. (fungus). Resistance triggered by BL occurred differently from systemic acquired resistance and resistance to disease induced by wounds; there was no induction of SA biosynthesis or acidic or basic pathogenesis-related (PR) gene expression [27].
BRs have also been effective pretreatments for postharvest diseases. Grape bunches, submitted to exogenous application of 24-EpiBR (0.4 and 0.8 mg L−1), stored at 0 ± 1 °C, without B. cinerea inoculation, evaluated in 15, 30 and 60 days, greater firmness and less deterioration, fall and weight loss of the berries were observed. In the fungus inoculation treatment, bunches were previously immersed in 1L of solution with 24-EpiBR and, after six h, artificially inoculated with B. cinerea and stored at 25 ± 2 °C and 95% relative humidity. There was an increase in resistance to grey mould with a reduction of oxidising substances (OO−2 and MDA) and an increase in the activities of defence-related enzymes (SOD, POD, CAT and PAL) in berries pretreated with BRs [139]. Similar results were observed in the Rosa hybridaBotrytis cinerea pathosystem. Potential defence elicitors against grey mould in cut roses indicated that BR at a concentration of 1 µmol/L provided disease inhibition and increased anthocyanin and water content, the vase life of cut flowers (~four days), of SOD and CAT activities and relative fresh mass [140].
In addition to increasing thermotolerance, BRs also increased the resistance of C. sinensis seedlings to C. gloeosporioides and significantly suppressed the spread of this fungus. The use of exogenous 24-epiBL reduced H2O2 accumulation in tea leaves. It markedly increased the PAL activity of leaves from 48.74% 24 h after inoculation with C. gloeosporioides to 71.21% when leaves received the combined treatment (24-epiBL and fungus suspension). This response is correlated with increased levels of transcription of genes involved in the phenylpropanoid pathway CsPAL, CsC4H and Cs4CL. There was an increase in the relative expression of the pathogenesis-related gene 1 (CsPR1), indicating that 24-epiBL activates the systemic defence response. However, there was a reduction in the face of the lipoxygenase 1 (CsLOX1) gene with this BR [141]. Failure to activate the lipoxygenase pathway can suppress JA biosynthesis, which is related to defence against pathogens [142].
In mango fruits, resistance against anthracnose (C. gloeosporioides) was improved by the exogenous use of 24-epiBL (ideal concentration of one μmol/L). The reduction in the severity and incidence of anthracnose in mango fruits treated with 24-epiBL was mainly due to the activation of defence-related enzymes (PAL, chitinase—CHT, cinnamate-4-hydroxylase—C4H, β-1,3-glucanase—GLU and 4-hydroxycinnamate CoA ligase—4CL), increased levels of total phenolics, flavonoids, lignin and prospecting, suppression of pectin hydrolases and regulation of ROS levels and antioxidant enzyme activities (CAT, APX, SOD and POD). There was no direct action of this BR to inhibit the growth of C. gloeosporioides in vitro [143]. On the other hand, fungistatic activity against Helminthosporium teres, which causes net blotch in barley (H. vulgare), was observed in vitro experiments, with 50% of the growth suppressed with the mechanical mixture of 24-epicastasterone and succinic acid. In the field, this combination (25 to 50 mg/ha) increased resistance to leaf spot and productivity with a 34% increase in seed mass per ear [144].

4.2. Nematodes

BRs can induce favourable or unfavourable host responses. In rice, we are applying epibrassinolide against root-knot nematode (Meloidogyne graminicola)-induced root susceptibility, favouring the nematodes. The mutual antagonistic effect on the signalling pathways between BR and JA is a critical factor in the interaction between Oryza sativaM. graminicola. Jasmonic acid activates PR genes that result in the expression of resistance. However, low concentrations of BR inhibit the expression of OsAOS2 (JA biosynthesis) and OsJAmyb (JA response) genes [145]. On the other hand, the exogenous application of BR in Solanum lycopersicum resulted in resistance to M. incognita. BR induced H2O2 production in the apoplast and activated MPK1/2/3 (mitogen-activated protein kinase 1/2/3), associated with induced systemic resistance (ISR) against root-knot nematodes. Plants with BR-deficient roots decreased the accumulation of apoplastic ROS, the transcription of Respiratory Burst Oxidase Homolog1 (RBOH1) and Whitefly Induced1 (WFI1), and the activation of MPK1/2 and MPK3. BR-induced activation of MPK1/2 is dependent on RBOH1. In this case, BR-induced systemic resistance is probably against the JA (ISR) and SA (SAR) independent nematodes [146].

4.3. Pests

In B. napus mutants, beneficial pleiotropic effects of the BR biosynthetic gene AtDWF4 were observed, resulting in antifungal action (Leptosphaeria maculans and Sclerotinia sclerotiorum), increased branching, silica and seed formation, more remarkable sprouting and root system, increased tolerance to drought and basal thermotolerance of the seedlings (at 45 °C/3–4 h, recovered at 20 °C, with the accumulation of heat shock protein, Hsp101 and Hsp90). These effects may result from three main mechanisms, in which BRs act as modulators of cell structure and functions, acting mainly on protein synthesis, increasing stress responses and under stress conditions, mobilising resources for growth regulation. Although the transgenic lines AtDWF4, BL35 and BL16 showed more excellent resistance to fungi, they were 50% and 100% more attractive to spider mites than wild-type plants [147].
The negative interaction of BR with the JA pathway has been associated with the formation of anti-herbivory characteristics (glandular trichomes, synthesis of allelochemicals and expression of enzymes, such as proteinase and polyphenol oxidase inhibitors) in tomatoes. Trichome density and allelochemical content are directly affected by BRs and JA. However, on the contrary, BR deficiency increases pubescence, whereas JA deficiency favours the reduction of trichomes. The double mutant with the genes that cause a lack of both hormones (dpy × jai-1) demonstrates an epistatic effect between these genes, indicating the action of BR in the JA signalling pathway. The interaction of these hormones in the defence against Spodoptera frugiperda and Tuta absoluta was confirmed through the density of trichomes since plants with the dpy gene had high densities of trichomes and plants with the dpy and jai-1 genes had an intermediate or higher density of trichomes compared to plants that contained only one of the genes. Plants with dpy, deficient in BRs, express proteinase inhibitor I (PI-I) and produce 4.5 times the content of zingiberene (zgb), an allelochemical with an anti-herbivory effect [148].
The plant–BR–insect interaction is little studied, but it could contribute to the control of insects in plantations. Rice plants’ susceptibility to brown plant hopper (BPH, Nilaparvata lugens) was increased when treated with 24-epiBL, and there is a suppression of the BRs’ pathway when there is BPH infestation and activation of the SA and JA pathways. This is confirmed with BR-deficient mutants, which increase resistance to this insect [149].

5. Commercial Products of BRs

In the early 1980s, the usefulness of BRs as agrochemicals was confirmed by studies carried out in Japan and the former Soviet Union [150,151,152]. Since then, numerous reports worldwide have appeared, and many potential practical uses have been patented [153]. Despite the vast importance of BRs, the amount present in plants could be higher, limiting their practical application in crops. One way to overcome this obstacle is using synthetic compounds called brassinosteroids analogues (Table 2), which play a physiological role similar to the natural one. These analogues have certain structural groups in natural BRs [154].
In soybean cultivars under water stress, a single foliar application of the functional analogue of BR DI31 (BB16) was able to promote stability in the canopy and efficient use of water from the first days of stress. There was also an improvement in the response antioxidant, nitrogen homeostasis, and attenuation of nodular senescence in roots. Finally, there was a 9% increase in productivity when applications were spaced at 21 days [160].
Foliar application of 1 μM of 24-epiBL in strawberries increased fruit growth rate, fruit mass, water, phenolic and anthocyanin contents, and carbohydrate metabolism [161]. The 24-epiBL steroid is one of the most active and studied forms of BRs, as it is organic, biodegradable and regulates several biochemical and physiological processes in plants [162].
BRs can influence the quantity produced and the quality of what is produced. The exogenous application of BRs at doses of 10 and 20 mg.ha−1 in potato plants promoted a 20% increase in productivity, reduced nitrate content in tubers and increased starch and vitamin C content [163]. The application of 28-homoBL directly to the roots of tomato seedlings 20 days after germination resulted in increased production of fruits with higher lycopene and ß-carotene contents compared to the control treatment [72].
Horticultural products (fruits, vegetables and greens) are highly perishable, requiring various post-harvest care to prolong the shelf life of these products. The BRs induce the ripening of the fruits; a solution for this is inhibiting their synthesis. Brassinazole (BZ), a BR inhibitor, increased the shelf life of fruits and vegetables [164]. The application of 5 µM of BZ in persimmon fruits (immersion for 30 min) delayed ripening, increased fruit firmness and cellulose content, decreased ethylene biosynthesis, the activity of enzymes that act in the degradation of the cell wall and the expression of genes related to ethylene biosynthesis [158]. The application of BL can accelerate fruit ripening by up to five days. With the application of BZ in strawberries, ripening was delayed by seven days compared to the control treatment [165].
In addition to the benefits promoted in crop yield and quality, BR’s analogues are environmentally friendly, as they are non-toxic and are a sustainable and low-cost tool for use in agriculture. Current research has been directed towards improving the understanding of the fundamental mechanisms of action of exogenously applied BRs and investigating the ideal stage of growth and development of plants so that their use is more efficient in the production and improvement of crops [153].

6. Conclusions

With global warming, the recurrence of abiotic and biotic stresses and the infeasibility of arable soils due to salinity or contamination by toxic metals, the search for new sustainable strategies is necessary. BRs are promising PGRs that act against biotic and abiotic stresses. However, not all BR biosynthesis pathways have been fully elucidated, which leads to a low number of synthetic BRs available on the market. Furthermore, endogenous BR synthesis is little explored and limited information is found in the literature.
BRs are involved in several processes in plant metabolism, including those associated with expressing epistatic and pleiotropic genes, increasing tolerance and/or resistance to abiotic and biotic stresses, and improvement of agronomical characteristics. Thus, the studies available in the literature (molecular, anatomical, biochemical, physiological, nutritional, morpho-agronomical, etc.) can generate opportunities to utilize these PGRs in agriculture and plant biology, maximizing the tolerance to stress. Additionally, interactions (synergistic and antagonistic) and crosstalk between BRs and other phytohormones also can be exciting strategies to control pests and diseases, being environmentally friendly and durable in various agricultural ecosystems [127].
In this review, we described several studies with abiotic stress, mainly focused on the action mechanisms of BRs in situations where plants are exposed to drought, salinity, and potentially toxic elements. With biotic stresses, our review provided recent research demonstrating the possibilities of BRs to mitigate adverse impacts occasioned by oomycetes, bacteria, fungi and viruses. However, nematodes and pests are critical biotic stresses whose interaction with BRs has not been explored.

Author Contributions

A.P.S.d.S., A.A.d.S.A., C.P.S. and M.d.S.B.d.A. wrote and edited the manuscript. A.K.d.S.L. critically revised it. All authors read and approved the final version of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received financial support from Fundação Amazônia de Amparo a Estudos e Pesquisas (FAPESPA/Brazil), Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq/Brazil) and Universidade Federal Rural da Amazônia (UFRA/Brazil) to A.K.S.L.

Data Availability Statement

Data are available upon request to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Went, F.W.; Went, F.W.; Thimann, K.V. Phytohormones; The Macmillan Company: New York, NY, USA, 1937. [Google Scholar]
  2. Gray, W.M. Hormonal Regulation of Plant Growth and Development. PLoS Biol. 2004, 2, e311. [Google Scholar] [CrossRef] [PubMed]
  3. Alcantara-Cortes, J.S.; Acero Godoy, J.; Alcántara Cortés, J.D.; Sánchez Mora, R.M. Principales Reguladores Hormonales y Sus Interacciones En El Crecimiento Vegetal. Nova 2019, 109–129. [Google Scholar] [CrossRef]
  4. Bücker-Neto, L.; Paiva, A.L.S.; Machado, R.D.; Arenhart, R.A.; Margis-Pinheiro, M. Interactions between Plant Hormones and Heavy Metals Responses. Genet. Mol. Biol. 2017, 40, 373–386. [Google Scholar] [CrossRef]
  5. Nishiyama, R.; Watanabe, Y.; Fujita, Y.; Le, D.T.; Kojima, M.; Werner, T.; Vankova, R.; Yamaguchi-Shinozaki, K.; Shinozaki, K.; Kakimoto, T.; et al. Analysis of Cytokinin Mutants and Regulation of Cytokinin Metabolic Genes Reveals Important Regulatory Roles of Cytokinins in Drought, Salt and Abscisic Acid Responses, and Abscisic Acid Biosynthesis. Plant Cell 2011, 23, 2169–2183. [Google Scholar] [CrossRef] [PubMed]
  6. Colebrook, E.H.; Thomas, S.G.; Phillips, A.L.; Hedden, P. The Role of Gibberellin Signalling in Plant Responses to Abiotic Stress. J. Exp. Biol. 2014, 217, 67–75. [Google Scholar] [CrossRef]
  7. Miao, R.; Li, C.; Liu, Z.; Zhou, X.; Chen, S.; Zhang, D.; Luo, J.; Tang, W.; Wang, C.; Wu, J.; et al. The Role of Endogenous Brassinosteroids in the Mechanisms Regulating Plant Reactions to Various Abiotic Stresses. Agronomy 2024, 14, 356. [Google Scholar] [CrossRef]
  8. Yao, T.; Xie, R.; Zhou, C.; Wu, X.; Li, D. Roles of Brossinosteroids Signaling in Biotic and Abiotic Stresses. J. Agric. Food Chem. 2023, 71, 7947–7960. [Google Scholar] [CrossRef] [PubMed]
  9. Mitchell, J.W.; Mandava, N.; Worley, J.F.; Plimmer, J.R.; Smith, M.V. Brassins-a New Family of Plant Hormones from Rape Pollen. Nature 1970, 225, 1065–1066. [Google Scholar] [CrossRef] [PubMed]
  10. Grove, M.D.; Spencer, G.F.; Rohwedder, W.K.; Mandava, N.; Worley, J.F.; Warthen, J.D.; Steffens, G.L.; Flippen-Anderson, J.L.; Cook, J.C. Brassinolide, a Plant Growth-Promoting Steroid Isolated from Brassica Napus Pollen. Nature 1979, 281, 216–217. [Google Scholar] [CrossRef]
  11. Friedrichsen, D.M.; Joazeiro, C.A.P.; Li, J.; Hunter, T.; Chory, J. Brassinosteroid-Insensitive-1 Is a Ubiquitously Expressed Leucine-Rich Repeat Receptor Serine/Threonine Kinase. Plant Physiol. 2000, 123, 1247–1256. [Google Scholar] [CrossRef]
  12. Manghwar, H.; Hussain, A.; Ali, Q.; Liu, F. Brassinosteroids (BRs) Role in Plant Development and Coping with Different Stresses. Int. J. Mol. Sci. 2022, 23, 1012. [Google Scholar] [CrossRef] [PubMed]
  13. Bajguz, A. Brassinosteroids – Occurence and Chemical Structures in Plants. In Brassinosteroids: A Class of Plant Hormone; Springer: Dordrecht, The Netherlands, 2011; pp. 1–27. [Google Scholar]
  14. Liu, J.; Zhang, D.; Sun, X.; Ding, T.; Lei, B.; Zhang, C. Structure-Activity Relationship of Brassinosteroids and Their Agricultural Practical Usages. Steroids 2017, 124, 1–17. [Google Scholar] [CrossRef] [PubMed]
  15. Bajguz, A.; Tretyn, A. The Chemical Characteristic and Distribution of Brassinosteroids in Plants. Phytochemistry 2003, 62, 1027–1046. [Google Scholar] [CrossRef] [PubMed]
  16. Sadura, I.; Janeczko, A. Physiological and Molecular Mechanisms of Brassinosteroid-Induced Tolerance to High and Low Temperature in Plants. Biol. Plant 2018, 62, 601–616. [Google Scholar] [CrossRef]
  17. Hayat, S.; Ahmad, A.; Fariduddin, Q. Brassinosteroids: A Regulator of 21st Century. In Brassinosteroids; Springer: Dordrecht, The Netherlands, 2003; pp. 231–246. [Google Scholar]
  18. Fujioka, S.; Yokota, T. Biosynthesis and Metabolism of Brassinosteroids. Annu. Rev. Plant Biol. 2003, 54, 137–164. [Google Scholar] [CrossRef] [PubMed]
  19. Bajguz, A.; Chmur, M.; Gruszka, D. Comprehensive Overview of the Brassinosteroid Biosynthesis Pathways: Substrates, Products, Inhibitors, and Connections. Front. Plant Sci. 2020, 11, 1–9. [Google Scholar] [CrossRef] [PubMed]
  20. Bajguz, A.; Hayat, S. Effects of Brassinosteroids on the Plant Responses to Environmental Stresses. Plant Physiol. Biochem. 2009, 47, 1–8. [Google Scholar] [CrossRef] [PubMed]
  21. Gapper, C.; Dolan, L. Control of Plant Development by Reactive Oxygen Species. Plant Physiol. 2006, 141, 341–345. [Google Scholar] [CrossRef] [PubMed]
  22. Li, L.; van Staden, J. Effects of Plant Growth Regulators on Drought Resistance of Two Maize Cultivars. South. Afr. J. Bot. 1998, 64, 116–120. [Google Scholar] [CrossRef]
  23. Núñes, M.; Mazzafera, P.; Mazorra, L.M.; Siqueira, W.J.; Zullo, M.A.T. Influence of a Brassinosteroid Analogue on Antioxidant Enzymes in Rice Grown in Culture Medium with NaCl. Biol. Plant 2003, 1, 67–70. [Google Scholar] [CrossRef]
  24. Özdemir, F.; Bor, M.; Demiral, T.; Türkan, İ. Effects of 24-Epibrassinolide on Seed Germination, Seedling Growth, Lipid Peroxidation, Proline Content and Antioxidative System of Rice (Oryza sativa L.) under Salinity Stress. Plant Growth Regul. 2004, 42, 203–211. [Google Scholar] [CrossRef]
  25. Vardhini, B.V.; Seeta, S.; Rao, R. Amelioration of Osmotic Stress by Brassinosteroids on Seed Germination and Seedling Growth of Three Varieties of Sorghum. Plant Growth Regul. 2003, 25–31. [Google Scholar] [CrossRef]
  26. Riboldi, L.B.; Gaziola, S.A.; Azevedo, R.A.; de Freitas, S.T.; de Camargo e Castro, P.R. 24-Epibrassinolide Mechanisms Regulating Blossom-End Rot Development in Tomato Fruit. J. Plant Growth Regul. 2019, 38, 812–823. [Google Scholar] [CrossRef]
  27. Ali, B.; Hasan, S.A.; Hayat, S.; Hayat, Q.; Yadav, S.; Fariduddin, Q.; Ahmad, A. A Role for Brassinosteroids in the Amelioration of Aluminium Stress through Antioxidant System in Mung Bean (Vigna Radiata L. Wilczek). Environ. Exp. Bot. 2008, 62, 153–159. [Google Scholar] [CrossRef]
  28. Zhao, Y.; Liang, Z.-Y.; Yang, Y.-J. Effects of Exogenous Brassinosteroid on Cd Tolerance in Solanum Nigrum Seedlings. Chin. J. Eco-Agric. 2013, 21, 872–876. [Google Scholar] [CrossRef]
  29. Hayat, S.; Hasan, S.A.; Hayat, Q.; Ahmad, A. Brassinosteroids Protect Lycopersicon esculentum from Cadmium Toxicity Applied as Shotgun Approach. Protoplasma 2010, 239, 3–14. [Google Scholar] [CrossRef] [PubMed]
  30. Hasan, S.A.; Hayat, S.; Ahmad, A. Brassinosteroids Protect Photosynthetic Machinery against the Cadmium Induced Oxidative Stress in Two Tomato Cultivars. Chemosphere 2011, 84, 1446–1451. [Google Scholar] [CrossRef] [PubMed]
  31. Sharma, I.; Pati, P.K.; Bhardwaj, R. Regulation of Growth and Antioxidant Enzyme Activities by 28-Homobrassinolide in Seedlings of Raphanus sativus L. under Cadmium Stress. Indian. J. Biochem. Biophys. 2010, 47, 172–177. [Google Scholar] [PubMed]
  32. Janeczko, A.; Koscielniak, J.; Pilipowicz, M.; Szarek-Lukaszewska, G.; Skoczowski, A. Protection of Winter Rape Photosystem 2 by 24-Epibrassinolide under Cadmium Stress. Photosynthetica 2005, 43, 293–298. [Google Scholar] [CrossRef]
  33. Sun, J.Y.; Guo, R.; Jiang, Q.; Chen, C.Z.; Gao, Y.Q.; Jiang, M.M.; Shen, R.F.; Zhu, X.F.; Huang, J. Brassinosteroid Decreases Cadmium Accumulation via Regulating Gibberellic Acid Accumulation and Cd Fixation Capacity of Root Cell Wall in Rice (Oryza sativa). J. Hazard. Mater. 2024, 469, 133862. [Google Scholar] [CrossRef]
  34. Fariduddin, Q.; Yusuf, M.; Hayat, S.; Ahmad, A. Effect of 28-Homobrassinolide on Antioxidant Capacity and Photosynthesis in Brassica juncea Plants Exposed to Different Levels of Copper. Environ. Exp. Bot. 2009, 66, 418–424. [Google Scholar] [CrossRef]
  35. Fariduddin, Q.; Khalil, R.R.A.E.; Mir, B.A.; Yusuf, M.; Ahmad, A. 24-Epibrassinolide Regulates Photosynthesis, Antioxidant Enzyme Activities and Proline Content of Cucumis sativus under Salt and/or Copper Stress. Environ. Monit. Assess. 2013, 185, 7845–7856. [Google Scholar] [CrossRef] [PubMed]
  36. Tadaiesky, L.B.A.; da Silva, B.R.S.; Batista, B.L.; Lobato, A.K. da S. Brassinosteroids Trigger Tolerance to Iron Toxicity in Rice. Physiol. Plant 2021, 171, 371–387. [Google Scholar] [CrossRef] [PubMed]
  37. Yang, P.; Wang, Y.; Li, J.; Bian, Z. Effects of Brassinosteroids on Photosynthetic Performance and Nitrogen Metabolism in Pepper Seedlings under Chilling Stress. Agronomy 2019, 9, 839. [Google Scholar] [CrossRef]
  38. Yusuf, M.; Fariduddin, Q.; Hayat, S.; Hasan, S.A.; Ahmad, A. Protective Response of 28-Homobrassinolide in Cultivars of Triticum aestivum with Different Levels of Nickel. Arch. Environ. Contam. Toxicol. 2011, 60, 68–76. [Google Scholar] [CrossRef] [PubMed]
  39. Guedes, F.R.C.M.; Maia, C.F.; da Silva, B.R.S.; Batista, B.L.; Alyemeni, M.N.; Ahmad, P.; da Silva Lobato, A.K. Exogenous 24-Epibrassinolide Stimulates Root Protection, and Leaf Antioxidant Enzymes in Lead Stressed Rice Plants: Central Roles to Minimize Pb Content and Oxidative Stress. Environ. Pollut. 2021, 280, 116992. [Google Scholar] [CrossRef] [PubMed]
  40. Soares, T.F.S.N.; dos Dias, D.C.F.S.; Oliveira, A.M.S.; Ribeiro, D.M.; dos Dias, L.A.S. Exogenous Brassinosteroids Increase Lead Stress Tolerance in Seed Germination and Seedling Growth of Brassica juncea L. Ecotoxicol. Environ. Saf. 2020, 193, 110296. [Google Scholar] [CrossRef] [PubMed]
  41. Arora, P.; Bhardwaj, R.; Kanwar, M.K. 24-Epibrassinolide Regulated Diminution of Cr Metal Toxicity in Brassica juncea L. Plants. Braz. J. Plant Physiol. 2010, 22, 159–165. [Google Scholar] [CrossRef]
  42. Ramakrishna, B.; Rao, S.S.R. Foliar Application of Brassinosteroids Alleviates Adverse Effects of Zinc Toxicity in Radish (Raphanus sativus L.) Plants. Protoplasma 2015, 252, 665–677. [Google Scholar] [CrossRef] [PubMed]
  43. Mahesh, K.; Balaraju, P.; Ramakrishna, B.; Ram Rao, S.S. Effect of Brassinosteroids on Germination and Seedling Growth of Radish (Raphanus sativus L.) under PEG-6000 Induced Water Stress. Am. J. Plant Sci. 2013, 04, 2305–2313. [Google Scholar] [CrossRef]
  44. Upreti, K.K.; Murti, G.S.R. Effects of Brassmosteroids on Growth, Nodulation, Phytohormone Content and Nitrogenase Activity in French Bean Under Water Stress. Biol. Plant 2004, 48, 407–411. [Google Scholar] [CrossRef]
  45. Hu, W.; Yan, X.; Xiao, Y.; Zeng, J.; Qi, H.; Ogweno, J.O. 24-Epibrassinosteroid Alleviate Drought-Induced Inhibition of Photosynthesis in Capsicum Annuum. Sci. Hortic. 2013, 150, 232–237. [Google Scholar] [CrossRef]
  46. Kang, Y.-Y.; Guo, S.-R.; Li, J.; Duan, J.-J. Effect of Root Applied 24-Epibrassinolide on Carbohydrate Status and Fermentative Enzyme Activities in Cucumber (Cucumis sativus L.) Seedlings under Hypoxia. Plant Growth Regul. 2009, 57, 259–269. [Google Scholar] [CrossRef]
  47. Behnamnia, M.; Kalantari, K.M.; Ziaie, J. The Effects of Brassinosteroid on the Induction of Biochemical Changes in Lycopersicon esculentum under Drought Stress. Turk. J. Bot. 2009, 33, 417–428. [Google Scholar] [CrossRef]
  48. Kaya, C.; Ashraf, M.; Wijaya, L.; Ahmad, P. The Putative Role of Endogenous Nitric Oxide in Brassinosteroid-Induced Antioxidant Defence System in Pepper (Capsicum Annuum L.) Plants under Water Stress. Plant Physiol. Biochem. 2019, 143, 119–128. [Google Scholar] [CrossRef] [PubMed]
  49. Zafari, M.; Ebadi, A.; Sedghi, M.; Jahanbakhsh, S. Alleviating Effect of 24- Epibrassinolide on Seed Oil Content and Fatty Acid Composition under Drought Stress in Safflower. J. Food Compos. Anal. 2020, 92, 103544. [Google Scholar] [CrossRef]
  50. Dos Santos Ribeiro, D.G.; da Silva, B.R.S.; da Silva Lobato, A.K. Brassinosteroids Induce Tolerance to Water Deficit in Soybean Seedlings: Contributions Linked to Root Anatomy and Antioxidant Enzymes. Acta Physiol. Plant 2019, 41, 82. [Google Scholar] [CrossRef]
  51. Avalbaev, A.; Bezrukova, M.; Allagulova, C.; Lubyanova, A.; Kudoyarova, G.; Fedorova, K.; Maslennikova, D.; Yuldashev, R.; Shakirova, F. Wheat Germ Agglutinin Is Involved in the Protective Action of 24-Epibrassinolide on the Roots of Wheat Seedlings under Drought Conditions. Plant Physiol. Biochem. 2020, 146, 420–427. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, Y.-T.; Chen, Z.-Y.; Jiang, Y.; Duan, B.-B.; Xi, Z.-M. Involvement of ABA and Antioxidant System in Brassinosteroid-Induced Water Stress Tolerance of Grapevine (Vitis Vinifera L.). Sci. Hortic. 2019, 256, 108596. [Google Scholar] [CrossRef]
  53. Savaliya, D.D.; Mandavia, C.K.; Mandavia, M.K. Role of Brassinolide on Enzyme Activities in Groundnut under Water Deficit Stress. Indian. J. Agric. Biochem. 2013, 26, 92–96. [Google Scholar]
  54. Zhang, M.; Zhai, Z.; Tian, X.; Duan, L.; Li, Z. Brassinolide Alleviated the Adverse Effect of Water Deficits on Photosynthesis and the Antioxidant of Soybean (Glycine Max L.). Plant Growth Regul. 2008, 56, 257–264. [Google Scholar] [CrossRef]
  55. Li, K.R.; Wang, H.H.; Han, G.; Wang, Q.J.; Fan, J. Effects of Brassinolide on the Survival, Growth and Drought Resistance of Robinia Pseudoacacia Seedlings under Water-Stress. New For. 2008, 35, 255–266. [Google Scholar] [CrossRef]
  56. Li, K.R.; Feng, C.H. Effects of Brassinolide on Drought Resistance of Xanthoceras Sorbifolia Seedlings under Water Stress. Acta Physiol. Plant 2011, 33, 1293–1300. [Google Scholar] [CrossRef]
  57. Gomes, M.d.M.d.A.; Torres Netto, A.; Campostrini, E.; Bressan-Smith, R.; Zullo, M.A.T.; Ferraz, T.M.; Siqueira, L.d.N.; Leal, N.R.; Núñez-Vázquez, M. Brassinosteroid Analogue Affects the Senescence in Two Papaya Genotypes Submitted to Drought Stress. Theor. Exp. Plant Physiol. 2013, 25, 186–195. [Google Scholar] [CrossRef]
  58. Jiménez-Bohórquez, E.F.; Díaz-Arias, M.A.; Balaguera-López, H.E. Exogenous Brassinosteroids Application in Purple Passion Fruit Plants Grafted onto a Sweet Calabash Passion Fruit Rootstock and under Water Stress. Rev. Colomb. De Cienc. Hortícolas 2024, 18, e16514. [Google Scholar]
  59. Salah, A.; Nwafor, C.C.; Han, Y.; Liu, L.; Rashid, M.; Batool, M.; El-Badri, A.M.; Cao, C.; Zhan, M. Spermidine and Brassinosteroid Regulate Root Anatomical Structure, Photosynthetic Traits and Antioxidant Defense Systems to Alleviate Waterlogging Stress in Maize Seedlings. South. Afr. J. Bot. 2022, 144, 389–402. [Google Scholar] [CrossRef]
  60. Li, X.; Wei, J.-P.; Ahammed, G.J.; Zhang, L.; Li, Y.; Yan, P.; Zhang, L.-P.; Han, W.-Y. Brassinosteroids Attenuate Moderate High Temperature-Caused Decline in Tea Quality by Enhancing Theanine Biosynthesis in Camellia sinensis L. Front. Plant Sci. 2018, 9, 1016. [Google Scholar] [CrossRef] [PubMed]
  61. Kurepin, L.V.; Qaderi, M.M.; Back, T.G.; Reid, D.M.; Pharis, R.P. A Rapid Effect of Applied Brassinolide on Abscisic Acid Concentrations in Brassica Napus Leaf Tissue Subjected to Short-Term Heat Stress. Plant Growth Regul. 2008, 55, 165–167. [Google Scholar] [CrossRef]
  62. Ogweno, J.O.; Song, X.S.; Shi, K.; Hu, W.H.; Mao, W.H.; Zhou, Y.H.; Yu, J.Q.; Nogués, S. Brassinosteroids Alleviate Heat-Induced Inhibition of Photosynthesis by Increasing Carboxylation Efficiency and Enhancing Antioxidant Systems in Lycopersicon esculentum. J. Plant Growth Regul. 2008, 27, 49–57. [Google Scholar] [CrossRef]
  63. Aghdam, M.S.; Mohammadkhani, N. Enhancement of Chilling Stress Tolerance of Tomato Fruit by Postharvest Brassinolide Treatment. Food Bioproc Tech. 2014, 7, 909–914. [Google Scholar] [CrossRef]
  64. Jiang, Y.-P.; Huang, L.-F.; Cheng, F.; Zhou, Y.-H.; Xia, X.-J.; Mao, W.-H.; Shi, K.; Yu, J.-Q. Brassinosteroids Accelerate Recovery of Photosynthetic Apparatus from Cold Stress by Balancing the Electron Partitioning, Carboxylation and Redox Homeostasis in Cucumber. Physiol. Plant 2013, 148, 133–145. [Google Scholar] [CrossRef] [PubMed]
  65. Huang, B.; Chu, C.-H.; Chen, S.-L.; Juan, H.-F.; Chen, Y.-M. A Proteomics Study of the Mung Bean Epicotyl Regulated by Brassinosteroids under Conditions of Chilling Stress. Cell Mol. Biol. Lett. 2006, 11, 264–278. [Google Scholar] [CrossRef] [PubMed]
  66. Janeczko, A.; Gullner, G.; Skoczowski, A.; Dubert, F.; Barna, B. Effects of Brassinosteroid Infiltration Prior to Cold Treatment on Ion Leakage and Pigment Contents in Rape Leaves. Biol. Plant 2007, 51, 355–358. [Google Scholar] [CrossRef]
  67. Wang, Y.; Jiang, Q.; Wang, X.; Xi, Z. Brassinosteroid Stimulates Hydrogen Peroxide Biosynthesis and Reduces the Effect of Cold Stress. J. Plant Growth Regul. 2022, 42, 3757–3769. [Google Scholar] [CrossRef]
  68. Qing-mao, S. Exogenous Brassinosteroid Induced the Salt Resistance of Cucumber (Cucumis sativus L.) Seedlings. Zhongguo Nong Ye Ke Xue = Zhongguo Nongye Kexue 2006, 39, 1872–1877. [Google Scholar]
  69. Daur, I.; Tatar, O. Effects of Gypsum and Brassinolide on Soil Properties, and Berseem (Trifolium Alexandrinum L.) Growth, Yield and Chemical Composition Grown on Saline Soil. Legume Res. 2013, 36, 306–311. [Google Scholar]
  70. Sharma, I.; Bhardwaj, R.; Pati, P.K. Stress Modulation Response of 24-Epibrassinolide against Imidacloprid in an Elite Indica Rice Variety Pusa Basmati-1. Pestic. Biochem. Physiol. 2013, 105, 144–153. [Google Scholar] [CrossRef]
  71. Abbas, S.; Latif, H.H.; Elsherbiny, E.A. Effect of 24-Epibrassinolide on the Physiological and Genetic Changes on Two Varieties of Pepper under Salt Stress Conditions. Pak. J. Bot. 2013, 45, 1273–1284. [Google Scholar]
  72. Ali, B.; Hayat, S.; Hasan, S.A.; Ahmad, A. Effect of Root Applied 28-Homobrassinolide on the Performance of Lycopersicon esculentum. Sci. Hortic. 2006, 110, 267–273. [Google Scholar] [CrossRef]
  73. Çoban, Ö.; Göktürk Baydar, N. Brassinosteroid Effects on Some Physical and Biochemical Properties and Secondary Metabolite Accumulation in Peppermint (Mentha Piperita L.) under Salt Stress. Ind. Crops Prod. 2016, 86, 251–258. [Google Scholar] [CrossRef]
  74. De Oliveira, V.P.; Lima, M.D.R.; da Silva, B.R.S.; Batista, B.L.; da Silva Lobato, A.K. Brassinosteroids Confer Tolerance to Salt Stress in Eucalyptus Urophylla Plants Enhancing Homeostasis, Antioxidant Metabolism and Leaf Anatomy. J. Plant Growth Regul. 2019, 38, 557–573. [Google Scholar] [CrossRef]
  75. Tabur, S.; Demir, K. Cytogenetic Response of 24-Epibrassinolide on the Root Meristem Cells of Barley Seeds under Salinity. Plant Growth Regul. 2009, 58, 119–123. [Google Scholar] [CrossRef]
  76. Ekinci, M.; Yildirim, E.; Dursun, A.; Turan, M. Mitigation of Salt Stress in Lettuce (Lactuca Sativa L. Var. Crispa) by Seed and Foliar 24-Epibrassinolide Treatments. HortScience 2012, 47, 631–636. [Google Scholar] [CrossRef]
  77. Rattan, A.; Kapoor, D.; Kapoor, N.; Bhardwaj, R.; Sharma, A. Brassinosteroids Regulate Functional Components of Antioxidative Defense System in Salt Stressed Maize Seedlings. J. Plant Growth Regul. 2020, 39, 1465–1475. [Google Scholar] [CrossRef]
  78. El-Khallal, S.M.; Hathout, T.A.; Ahsour, A.E.R.A.; Kerrit, A.A.A. Brassinolide and Salicylic Acid Induced Antioxidant Enzymes, Hormonal Balance and Protein Profile of Maize Plants Grown under Salt Stress. Res. J. Agric. Biol. Sci. 2009, 5, 391–402. [Google Scholar]
  79. Soliman, M.; Elkelish, A.; Souad, T.; Alhaithloul, H.; Farooq, M. Brassinosteroid Seed Priming with Nitrogen Supplementation Improves Salt Tolerance in Soybean. Physiol. Mol. Biol. Plants 2020, 26, 501–511. [Google Scholar] [CrossRef] [PubMed]
  80. Zhang, S.; Hu, J.; Zhang, Y.; Xie, X.J.; Knapp, A. Seed Priming with Brassinolide Improves Lucerne (Medicago Sativa L.) Seed Germination and Seedling Growth in Relation to Physiological Changes under Salinity Stress. Aust. J. Agric. Res. 2007, 58, 811. [Google Scholar] [CrossRef]
  81. Shahbaz, M.; Ashraf, M.; Athar, H.-R. Does Exogenous Application of 24-Epibrassinolide Ameliorate Salt Induced Growth Inhibition in Wheat (Triticum aestivum L.)? Plant Growth Regul. 2008, 55, 51–64. [Google Scholar] [CrossRef]
  82. Avalbaev, A.M.; Yuldashev, R.A.; Fatkhutdinova, R.A.; Urusov, F.A.; Safutdinova, Y.V.; Shakirova, F.M. The Influence of 24-Epibrassidinolide on the Hormonal Status of Wheat Plants under Sodium Chloride. Appl. Biochem. Microbiol. 2010, 46, 99–102. [Google Scholar] [CrossRef]
  83. Song, S.; Liu, W.; Guo, S.; Shang, Q.; Zhang, Z. Salt Resistance and Its Mechanism of Cucumber under Effects of Exogenous Chemical Activators. Chin. J. Appl. Ecol. 2006, 17, 1871–1876. [Google Scholar]
  84. Wang, B.; Zhang, J.; Xia, X.; Zhang, W.-H. Ameliorative Effect of Brassinosteroid and Ethylene on Germination of Cucumber Seeds in the Presence of Sodium Chloride. Plant Growth Regul. 2011, 65, 407–413. [Google Scholar] [CrossRef]
  85. Arora, N.; Bhardwaj, R.; Sharma, P.; Arora, H.K. Effects of 28-Homobrassinolide on Growth, Lipid Peroxidation and Antioxidative Enzyme Activities in Seedlings of Zea Mays L. under Salinity Stress. Acta Physiol. Plant 2008, 30, 833–839. [Google Scholar] [CrossRef]
  86. El-Mashad, A.A.A.; Mohamed, H.I. Brassinolide Alleviates Salt Stress and Increases Antioxidant Activity of Cowpea Plants (Vigna Sinensis). Protoplasma 2012, 249, 625–635. [Google Scholar] [CrossRef] [PubMed]
  87. Karlidag, H.; Yildirim, E.; Turan, M. Role of 24-Epibrassinolide in Mitigating the Adverse Effects of Salt Stress on Stomatal Conductance, Membrane Permeability, and Leaf Water Content, Ionic Composition in Salt Stressed Strawberry (Fragaria×ananassa). Sci. Hortic. 2011, 130, 133–140. [Google Scholar] [CrossRef]
  88. Dalio, R.J.D.; Pinheiro, H.P.; Sodek, L.; Haddad, C.R.B. 24-Epibrassinolide Restores Nitrogen Metabolism of Pigeon Pea under Saline Stress. Bot. Stud. 2013, 54, 9. [Google Scholar] [CrossRef] [PubMed]
  89. Alyemeni, M.N.; Hayat, S.; Wijaya, L.; Anaji, A. Foliar Application of 28-Homobrassinolide Mitigates Salinity Stress by Increasing the Efficiency of Photosynthesis in Brassica juncea. Acta Bot. Bras. 2013, 27, 502–505. [Google Scholar] [CrossRef]
  90. Fedina, E.O. Effect of 24-Epibrassinolide on Pea Protein Tyrosine Phosphorylation after Salinity Action. Russ. J. Plant Physiol. 2013, 60, 351–358. [Google Scholar] [CrossRef]
  91. Ding, H.D.; Zhu, X.-H.; Zhu, Z.W.; Yang, S.J.; Zha, D.S.; Wu, X.X. Amelioration of Salt-Induced Oxidative Stress in Eggplant by Application of 24-Epibrassinolide. Biol. Plant 2012, 56, 767–770. [Google Scholar] [CrossRef]
  92. Talaat, N.B.; Shawky, B.T. 24-Epibrassinolide Alleviates Salt-Induced Inhibition of Productivity by Increasing Nutrients and Compatible Solutes Accumulation and Enhancing Antioxidant System in Wheat (Triticum aestivum L.). Acta Physiol. Plant 2013, 35, 729–740. [Google Scholar] [CrossRef]
  93. Zeng, L.L.; Song, L.Y.; Wu, X.; Ma, D.N.; Song, S.W.; Wang, X.X.; Zheng, H.L. Brassinosteroid Enhances Salt Tolerance via S-Nitrosoglutathione Reductase and Nitric Oxide Signaling Pathway in Mangrove Kandelia Obovata. Plant Cell Environ. 2024, 47, 511–526. [Google Scholar] [CrossRef] [PubMed]
  94. Krishna, P.; Prasad, B.D.; Rahman, T. Brassinosteroid Action in Plant Abiotic Stress Tolerance. In Methods in Molecular Biology; Humana Press Inc.: Totowa, NJ, USA, 2017; Volume 1564, pp. 193–202. [Google Scholar]
  95. Fang, P.; Yan, M.; Chi, C.; Wang, M.; Zhou, Y.; Zhou, J.; Shi, K.; Xia, X.; Foyer, C.H.; Yu, J. Brassinosteroids Act as a Positive Regulator of Photoprotection in Response to Chilling Stress. Plant Physiol. 2019, 180, 2061–2076. [Google Scholar] [CrossRef] [PubMed]
  96. Gururani, M.A.; Venkatesh, J.; Tran, L.S.P. Regulation of Photosynthesis during Abiotic Stress-Induced Photoinhibition. Mol. Plant 2015, 8, 1304–1320. [Google Scholar] [CrossRef] [PubMed]
  97. Müssig, C.; Fischer, S.; Altmann, T. Brassinosteroid-Regulated Gene Expression. Plant Physiol. 2002, 129, 1241–1251. [Google Scholar] [CrossRef] [PubMed]
  98. Dhaubhadel, S.; Chaudhary, S.; Dobinson, K.F.; Krishna, P. Treatment with 24-Epibrassinolide, a Brassinosteroid, Increases the Basic Thermotolerance of Brassica Napus and Tomato Seedlings. Plant Mol. Biol. 1999, 40, 333–342. [Google Scholar] [CrossRef]
  99. Shevela, D.; Kern, J.F.; Govindjee, G.; Messinger, J. Solar Energy Conversion by Photosystem II: Principles and Structures. Photosynth. Res. 2023, 156, 279–307. [Google Scholar] [CrossRef] [PubMed]
  100. Cao, S.; Xu, Q.; Cao, Y.; Qian, K.; An, K.; Zhu, Y.; Binzeng, H.; Zhao, H.; Kuai, B. Loss-of-Function Mutations in DET2 Gene Lead to an Enhanced Resistance to Oxidative Stress in Arabidopsis. Physiol. Plant 2005, 123, 57–66. [Google Scholar] [CrossRef]
  101. Choe, S. Brassinosteroid Biosynthesis and Inactivation. Physiol. Plant 2006, 126, 539–548. [Google Scholar] [CrossRef]
  102. Cobbett, C.; Goldsbrough, P. Phytochelatins and Metallothioneins: Roles in Heavy Metal Detoxification and Homeostasis. Annu. Rev. Plant Biol. 2002, 53, 159–182. [Google Scholar] [CrossRef] [PubMed]
  103. Gao, H.; Kang, L.; Liu, Q.; Cheng, N.; Wang, B.; Cao, W. Effect of 24-Epibrassinolide Treatment on the Metabolism of Eggplant Fruits in Relation to Development of Pulp Browning under Chilling Stress. J. Food Sci. Technol. 2014, 52, 3394–3401. [Google Scholar] [CrossRef] [PubMed]
  104. Dhaubhadel, S.; Browning, K.S.; Gallie, D.R.; Krishna, P. Brassinosteroid Functions to Protect the Translational Machinery and Heat-shock Protein Synthesis Following Thermal Stress. Plant J. 2002, 29, 681–691. [Google Scholar] [CrossRef] [PubMed]
  105. Divi, U.K.; Rahman, T.; Krishna, P. Gene Expression and Functional Analyses in Brassinosteroid-Mediated Stress Tolerance. Plant Biotechnol. J. 2016, 14, 419–432. [Google Scholar] [CrossRef] [PubMed]
  106. Wilen, R.W.; Sacco, M.; Gusta, L.V.; Krishna, P. Effects of 24-Epibrassinolide on Freezing and Thermotolerance of Bromegrass (Bromus Inermis) Cell Cultures. Physiol. Plant 1995, 95, 195–202. [Google Scholar] [CrossRef]
  107. Mazorra, L.M.; Núñez, M.; Hechavarria, M.; Coll, F.; Sánchez-Blanco, M.J. Influence of Brassinosteroids on Antioxidant Enzymes Activity in Tomato Under Different Temperatures. Biol. Plant 2002, 45, 593–596. [Google Scholar] [CrossRef]
  108. Fujii, S.; Saka, H. The Promotive Effect of Brassinolide on Lamina Joint-Cell Elongation, Germination and Seedling Growth under Low-Temperature Stress in Rice (Oryza sativa L.). Plant Prod. Sci. 2001, 4, 210–214. [Google Scholar] [CrossRef]
  109. Hotta, Y.; Tanaka, T.; Luo, B.; Takeuchi, Y.; Konnai, M. Improvement of Cold Resistance in Rice Seedlings by 5-Aminolevulinic Acid. J. Pestic. Sci. 1998, 23, 29–33. [Google Scholar] [CrossRef]
  110. He, R.; Wang, G.; Wang, X. Effects of Brassinolide on Growth and Chilling Resistance of Maize Seedlings. In Brassinosteroids; American Chemical Society: Washington, DC, USA, 1991; pp. 220–230. [Google Scholar]
  111. Khripach, V. Twenty Years of Brassinosteroids: Steroidal Plant Hormones Warrant Better Crops for the XXI Century. Ann. Bot. 2000, 86, 441–447. [Google Scholar] [CrossRef]
  112. Gonzalez Olmedo, J.L.; Cordova, A.; Aragon, C.E.; Pina, D.; Rivas, M.; Rodriguez, R. Effect of an Analogue of Brassinosteroid on FHIA-18 Plantlets Exposed to Thermal Stress. Infomusa 2005, 14, 18–20. [Google Scholar]
  113. Pimentel, C. A Relação da Planta Com a Água, 1st ed.; Edur: Seropédica, Brazil, 2004; pp. 171–191. ISBN 85-85720-45-X. [Google Scholar]
  114. Janeczko, A.; Gruszka, D.; Pociecha, E.; Dziurka, M.; Filek, M.; Jurczyk, B.; Kalaji, H.M.; Kocurek, M.; Waligórski, P. Physiological and Biochemical Characterisation of Watered and Drought-Stressed Barley Mutants in the HvDWARF Gene Encoding C6-Oxidase Involved in Brassinosteroid Biosynthesis. Plant Physiol. Biochem. 2016, 99, 126–141. [Google Scholar] [CrossRef]
  115. Ramonell, K.M.; Kuang, A.; Porterfield, D.M.; Crispi, M.L.; Xiao, Y.; Mcclure, G.; Musgrave, M.E. Influence of Atmospheric Oxygen on Leaf Structure and Starch Deposition in Arabidopsis Thaliana. Plant Cell Environ. 2001, 24, 419–428. [Google Scholar] [CrossRef] [PubMed]
  116. Ludazi, M.; Lambrou, Y.; Shaw, J.; Monsieur, C. Gender and Sustainable Development in Drylands: An Analysis of Field Experiences; Food and Agriculture Organization of the United Nations: Rome, Italy, 2003. [Google Scholar]
  117. Srivastava, P.; Wu, Q.-S.; Giri, B. Salinity: An Overview. In Microorganisms in Saline Environments: Strategies and Functions; Springer: Cham, Switzerland, 2019; pp. 3–18. [Google Scholar]
  118. Wei, L.-J.; Deng, X.-G.; Zhu, T.; Zheng, T.; Li, P.-X.; Wu, J.-Q.; Zhang, D.-W.; Lin, H.-H. Ethylene Is Involved in Brassinosteroids Induced Alternative Respiratory Pathway in Cucumber (Cucumis sativus L.) Seedlings Response to Abiotic Stress. Front. Plant Sci. 2015, 6, 982. [Google Scholar] [CrossRef] [PubMed]
  119. Zhu, T.; Deng, X.G.; Tan, W.R.; Zhou, X.; Luo, S.S.; Han, X.Y.; Zhang, D.W.; Lin, H.H. Nitric Oxide Is Involved in Brassinosteroid-Induced Alternative Respiratory Pathway in Nicotiana Benthamiana Seedlings’ Response to Salt Stress. Physiol. Plant 2016, 156, 150–163. [Google Scholar] [CrossRef] [PubMed]
  120. Yaashikaa, P.R.; Kumar, P.S.; Jeevanantham, S.; Saravanan, R. A Review on Bioremediation Approach for Heavy Metal Detoxification and Accumulation in Plants. Environ. Pollut. 2022, 301, 119035. [Google Scholar] [CrossRef]
  121. Anuradha, S.; Rao, S.S.R. The Effect of Brassinosteroids on Radish (Raphanus sativus L.) Seedlings Growing under Cadmium Stress. Plant Soil. Environ. 2008, 53, 465–472. [Google Scholar] [CrossRef]
  122. Murashige, T.; Skoog, F. A Revised Medium for Rapid Growth and Bio Assays with Tobacco Tissue Cultures. Physiol. Plant 1962, 15, 473–497. [Google Scholar] [CrossRef]
  123. Cohen, J.D.; Meudt, W.J. Investigations on the Mechanism of the Brassinosteroid Response. Plant Physiol. 1983, 72, 691–694. [Google Scholar] [CrossRef] [PubMed]
  124. Katsumi, M. Physiological Modes of Brassinolide Action in Cucumber Hypocotyl Growth. In Brassinosteroids. Chemistry, Bioactivity and Applications; Cutler, H.G., Yokota, T., Adam, G., Eds.; ACS Symposium Series; American Chemical Society: Washington, DC, USA, 1991; Volume 474, pp. 246–254. ISBN 9780841221260. [Google Scholar]
  125. Xu, H.; Shida, A.; Futatsuya, F.; Kumura, A. Effects of Epibrassinolide and Abscisic Acid on Sorghum Plants Growing under Soil Water Deficit. I. Effects on Growth and Survival. Jpn. J. Crop Sci. 1994, 63, 671–675. [Google Scholar] [CrossRef]
  126. Kim, Y.-W.; Youn, J.-H.; Roh, J.; Kim, J.-M.; Kim, S.-K.; Kim, T.-W. Brassinosteroids Enhance Salicylic Acid-Mediated Immune Responses by Inhibiting BIN2 Phosphorylation of Clade I TGA Transcription Factors in Arabidopsis. Mol. Plant 2022, 15, 991–1007. [Google Scholar] [CrossRef] [PubMed]
  127. De Bruyne, L.; Höfte, M.; de Vleesschauwer, D. Connecting Growth and Defense: The Emerging Roles of Brassinosteroids and Gibberellins in Plant Innate Immunity. Mol. Plant 2014, 7, 943–959. [Google Scholar] [CrossRef] [PubMed]
  128. Michereff, S.J. Fundamentos de Fitopatologia; UFRPE: Recife, Brazil, 2001; pp. 1–145. [Google Scholar]
  129. Taiz, L.; Zeiger, E. Fisiologia Vegetal, 5th ed.; Artemed: Porto Alegre, Brazil, 2013. [Google Scholar]
  130. Agrios, G. Plant Pathology: Fifth Edition; Elsevier: Amsterdam, The Netherlands, 2005; pp. 1–922. [Google Scholar] [CrossRef]
  131. Liu, J.-Z.; Lam, H.-M. Signal Transduction Pathways in Plants for Resistance against Pathogens. Int. J. Mol. Sci. 2019, 20, 2335. [Google Scholar] [CrossRef] [PubMed]
  132. Ma, K.-W.; Ma, W. Phytohormone Pathways as Targets of Pathogens to Facilitate Infection. Plant Mol. Biol. 2016, 91, 713–725. [Google Scholar] [CrossRef] [PubMed]
  133. Svoboda, T.; Thon, M.R.; Strauss, J. The Role of Plant Hormones in the Interaction of Colletotrichum Species with Their Host Plants. Int. J. Mol. Sci. 2021, 22, 12454. [Google Scholar] [CrossRef] [PubMed]
  134. Bari, R.; Jones, J.D.G. Role of Plant Hormones in Plant Defence Responses. Plant Mol. Biol. 2009, 69, 473–488. [Google Scholar] [CrossRef] [PubMed]
  135. Sudheeran, P.K.; Sela, N.; Carmeli-Weissberg, M.; Ovadia, R.; Panda, S.; Feygenberg, O.; Maurer, D.; Oren-Shamir, M.; Aharoni, A.; Alkan, N. Induced Defense Response in Red Mango Fruit against Colletotrichum Gloeosporioides. Hortic. Res. 2021, 8, 17. [Google Scholar] [CrossRef] [PubMed]
  136. Krishna, P. Brassinosteroid-Mediated Stress Responses. J. Plant Growth Regul. 2003, 22, 289–297. [Google Scholar] [CrossRef] [PubMed]
  137. De Vleesschauwer, D.; Van Buyten, E.; Satoh, K.; Balidion, J.; Mauleon, R.; Choi, I.-R.; Vera-Cruz, C.; Kikuchi, S.; Höfte, M. Brassinosteroids Antagonize Gibberellin- and Salicylate-Mediated Root Immunity in Rice. Plant Physiol. 2012, 158, 1833–1846. [Google Scholar] [CrossRef] [PubMed]
  138. Furio, R.N.; Albornoz, P.L.; Coll, Y.; Martínez Zamora, G.M.; Salazar, S.M.; Martos, G.G.; Díaz Ricci, J.C. Effect of Natural and Synthetic Brassinosteroids on Strawberry Immune Response against Colletotrichum Acutatum. Eur. J. Plant Pathol. 2019, 153, 167–181. [Google Scholar] [CrossRef]
  139. Liu, Q.; Xi, Z.; Gao, J.; Meng, Y.; Lin, S.; Zhang, Z. Effects of Exogenous 24-Epibrassinolide to Control Grey Mould and Maintain Postharvest Quality of Table Grapes. Int. J. Food Sci. Technol. 2016, 51, 1236–1243. [Google Scholar] [CrossRef]
  140. Naeemi, M.H.; Kalateh Jari, S.; Zarinnia, V.; Fatehi, F. Changes in Physio-Biochemical Status of Cut of Rose (Rosa Hybrida L. Cv. Samurai) Flowers under Methyl Jasmonate, Brassinosteroid, and Fungal Elicitor to Control Gray Mold. Sci. Hortic. 2022, 306, 111402. [Google Scholar] [CrossRef]
  141. Zhang, L.; Ahammed, G.J.; Li, X.; Wei, J.P.; Li, Y.; Yan, P.; Zhang, L.P.; Han, W.Y. Exogenous Brassinosteroid Enhances Plant Defense Against Colletotrichum Gloeosporioides by Activating Phenylpropanoid Pathway in Camellia sinensis L. J. Plant Growth Regul. 2018, 37, 1235–1243. [Google Scholar] [CrossRef]
  142. Li, X.; Ahammed, G.J.; Li, Z.; Tang, M.; Yan, P.; Han, W. Decreased Biosynthesis of Jasmonic Acid via Lipoxygenase Pathway Compromised Caffeine-Induced Resistance to Colletotrichum Gloeosporioides Under Elevated CO2 in Tea Seedlings. Phytopathology 2016, 106, 1270–1277. [Google Scholar] [CrossRef] [PubMed]
  143. Song, Y.; Hu, C.; Xue, Y.; Gu, J.; He, J.; Ren, Y. 24-Epibrassinolide Enhances Mango Resistance to Colletotrichum Gloeosporioides via Activating Multiple Defense Response. Sci. Hortic. 2022, 303, 111249. [Google Scholar] [CrossRef]
  144. Manzhalesava, N.; Litvinovskaya, R.P.; Poljanskaja, S.N.; Khripach, V.A. The Effect of Phytohormonal Steroids in Combination with Succinic Acid on the Resistance of Hordeum Vulgare L. to Helminthosporium Teres Sacc. Open Agric. J. 2022, 16, e187433152207130. [Google Scholar] [CrossRef]
  145. Nahar, K.; Kyndt, T.; Hause, B.; Höfte, M.; Gheysen, G. Brassinosteroids Suppress Rice Defense against Root-Knot Nematodes through Antagonism with the Jasmonate Pathway. Mol. Plant-Microbe Interact. 2013, 26, 106–115. [Google Scholar] [CrossRef] [PubMed]
  146. Song, L.; Xu, X.; Wang, F.; Wang, Y.; Xia, X.; Shi, K.; Zhou, Y.; Zhou, J.; Yu, J. Brassinosteroids Act as a Positive Regulator for Resistance against Root-Knot Nematode Involving Respiratory Burst Oxidase Homolog-Dependent Activation of MAPKs in Tomato. Plant Cell Environ. 2018, 41, 1113–1125. [Google Scholar] [CrossRef] [PubMed]
  147. Sahni, S.; Prasad, B.D.; Liu, Q.; Grbic, V.; Sharpe, A.; Singh, S.P.; Krishna, P. Overexpression of the Brassinosteroid Biosynthetic Gene DWF4 in Brassica Napus Simultaneously Increases Seed Yield and Stress Tolerance. Sci. Rep. 2016, 6, 28298. [Google Scholar] [CrossRef] [PubMed]
  148. Campos, M.L.; de Almeida, M.; Rossi, M.L.; Martinelli, A.P.; Litholdo Junior, C.G.; Figueira, A.; Rampelotti-Ferreira, F.T.; Vendramim, J.D.; Benedito, V.A.; Pereira Peres, L.E. Brassinosteroids Interact Negatively with Jasmonates in the Formation of Anti-Herbivory Traits in Tomato. J. Exp. Bot. 2009, 60, 4347–4361. [Google Scholar] [CrossRef] [PubMed]
  149. Pan, G.; Liu, Y.; Ji, L.; Zhang, X.; He, J.; Huang, J.; Qiu, Z.; Liu, D.; Sun, Z.; Xu, T.; et al. Brassinosteroids Mediate Susceptibility to Brown Planthopper by Integrating with the Salicylic Acid and Jasmonic Acid Pathways in Rice. J. Exp. Bot. 2018, 69, 4433–4442. [Google Scholar] [CrossRef] [PubMed]
  150. Abe, H. Advances in Brassinosteroid Research and Prospects for Its Agricultural Application. J. Pestic. Sci. 1989, 55, 10–14. [Google Scholar]
  151. Fujita, F. Prospects for Brassinolide Utilization in Agriculture. Chem. Biol. 1985, 23, 715–725. [Google Scholar]
  152. Takeuchi, T. Application of Brassinosteroids in Agriculture in Japan. In Proceedings of the 19th Annual Meeting of Plant Growth Regulator Society of America, Washington, DC, USA, 6–7 August 1992; pp. 343–352. [Google Scholar]
  153. Bhanu, A.N. Brassinosteroids: Relevance in Biological Activities of Plants and Agriculture. J. Plant Sci. Res. 2019, 35, 1–15. [Google Scholar] [CrossRef]
  154. Garciglia, R.S.; Rodríguez, M.A.C.; del Río, R.E. Uso de Brasinoesteroides y Sus Análogos En La Agricultura. Biológicas 2008, 18–27. [Google Scholar]
  155. Vargas-Vázquez, M.L.P.; Irizar-Garza, M.B.G. Efecto Del Brasinoesteroide y Densidad de Población En La Acumulación de Biomasa y Rendimiento de Ayocote (Phaseolus Coccineus L.). Rev. Chapingo Ser. Hortic. 2005, XI, 269–272. [Google Scholar] [CrossRef]
  156. Torres-Ruiz, B.L.; Espinosa-Calderón, A.; Mendoza-Rodríguez, M.; de la Rodríguez, O.J.L.; Irizar Garza, M.B.; Sahagún-Castellanos Ruiz, J. Efecto de Brasinoesteroides En Híbridos de Maíz Androestériles y Fértiles. Agron. Mesoam. 2006, 18, 155. [Google Scholar] [CrossRef]
  157. Serna, M.; Hernández, F.; Coll, F.; Amorós, A. Brassinosteroid Analogues Effect on Yield and Quality Parameters of Field-Grown Lettuce (Lactuca Sativa L.). Sci. Hortic. 2012, 143, 29–37. [Google Scholar] [CrossRef]
  158. He, Y.; Li, J.; Ban, Q.; Han, S.; Rao, J. Role of Brassinosteroids in Persimmon (Diospyros Kaki L.) Fruit Ripening. J. Agric. Food Chem. 2018, 66, 2637–2644. [Google Scholar] [CrossRef] [PubMed]
  159. Babalık, Z.; Demirci, T.; Aşcı, Ö.A.; Baydar, N.G. Brassinosteroids Modify Yield, Quality, and Antioxidant Components in Grapes (Vitis Vinifera Cv. Alphonse Lavallée). J. Plant Growth Regul. 2020, 39, 147–156. [Google Scholar] [CrossRef]
  160. Perez-Borroto, L.S.; Guzzo, M.C.; Posada, G.; Peña Malavera, A.N.; Castagnaro, A.P.; Gonzalez-Olmedo, J.L.; Coll-García, Y.; Pardo, E.M. A Brassinosteroid Functional Analogue Increases Soybean Drought Resilience. Sci. Rep. 2022, 12, 11294. [Google Scholar] [CrossRef] [PubMed]
  161. Zahedipour-Sheshglani, P.; Asghari, M. Impact of Foliar Spray with 24-Epibrassinolide on Yield, Quality, Ripening Physiology and Productivity of the Strawberry. Sci. Hortic. 2020, 268, 109376. [Google Scholar] [CrossRef]
  162. Choudhary, S.P.; Yu, J.-Q.; Yamaguchi-Shinozaki, K.; Shinozaki, K.; Tran, L.-S.P. Benefits of Brassinosteroid Crosstalk. Trends Plant Sci. 2012, 17, 594–605. [Google Scholar] [CrossRef] [PubMed]
  163. Khripach, V.A.; Zhabinskii, V.N.; Khripach, N.B. New Practical Aspects of Brassinosteroids and Results of Their Ten-Year Agricultural Use in Russia and Belarus. In Brassinosteroids; Springer: Dordrecht, The Netherlands, 2003; pp. 189–230. [Google Scholar]
  164. Zhu, T.; Tan, W.-R.; Deng, X.-G.; Zheng, T.; Zhang, D.-W.; Lin, H.-H. Effects of Brassinosteroids on Quality Attributes and Ethylene Synthesis in Postharvest Tomato Fruit. Postharvest Biol. Technol. 2015, 100, 196–204. [Google Scholar] [CrossRef]
  165. Chai, Y.M.; Zhang, Q.; Tian, L.; Li, C.L.; Xing, Y.; Qin, L.; Shen, Y.Y. Brassinosteroid Is Involved in Strawberry Fruit Ripening. Plant Growth Regul. 2013, 69, 63–69. [Google Scholar] [CrossRef]
Figure 1. Benefits induced by brassinosteroids in molecular, biochemical, physiological, anatomical, nutritional and morpho-agronomical aspects.
Figure 1. Benefits induced by brassinosteroids in molecular, biochemical, physiological, anatomical, nutritional and morpho-agronomical aspects.
Agronomy 14 00840 g001
Table 1. Effects of exogenous application of brassinosteroids on different plant species and abiotic stresses.
Table 1. Effects of exogenous application of brassinosteroids on different plant species and abiotic stresses.
Stress ConditionsBrassinosteroids and
Concentration
Plant SpeciesBRs ApplicationBenefits Induced by BRs
Al [0–10 mM]24-epiBL/28-homoBL [10−8 M]Vigna radiataLeafBiochemical [27]
CdEpiBLSolanum nigrum Biochemical [28]
Cd [0–12 mg.kg−1]24-epiBL/28-homoBL [10−8 M]Lycopersicon esculentumLeafPhysiological and biochemical [29]
Cd [1.0 mM]24-epiBL [5 µM]Phaseolus vulgarisLeafMorphological and biochemical [30]
Cd [1.0 mM]24-epiBL/28-homoBL [3.0 µM]Raphanus sativusSeedBiochemical [31]
Cd [300 µM]24-epiBL [100 nM]Brassica NapusCulture mediumPhysiological [32]
Cd [1.0 μM]BRs [0.01–10 μM]Oryza sativaNutrient solutionMolecular [33]
Cu [100 mg.kg−1]24-epiBL [0.01 μM]Cucumis sativusLeafMorphological, physiological and biochemical [34]
Cu [50–150 mg.kg−1]28-HomoBL [10−10–10−6 M]Brassica junceaSeedMorphological, physiological and biochemical [35]
Fe [250–6250 μM]24-epiBL [0–10 nM]Oryza sativaRootPhysiological [36]
Ni [0.0–0.6 mM]24-epiBL [10−11–10−7 M]Brassica junceaLeafBiochemical [37]
Ni [50–100 μM]28-homoBL [0.01 µM]Triticum aestivumLeafMorphological, physiological and biochemical [38]
Pb [0–200 μM]24-epiBL [0–100 nM]Oryza sativaLeafAnatomical and biochemical [39]
Pb [3 mM]24-epiBL [10−10–10−6 M]Brassica junceaSeedBiochemical [40]
Zn [0.0–2.0 mM]24-epiBL [10−10–10−6 M]Brassica junceaSeedMorphological and biochemical [41]
Zn [1–10 mM]28-homoBL [0.5–2 μM]Raphanus sativusSeedBiochemical [42,43]
DroughtEpiBL/HomoBL [1–5 µM]Phaseolus
vulgaris
LeafMorpho-agronomical [44]
Drought24-epiBL [0.01 mg.L−1]Capsicum
annuum
LeafPhysiological [45]
Drought24-epiBL [1 μg.L−1]Cucumis sativusRootPhysiological and biochemical [46]
Drought24-epiBL [0.01–1 µM]Lycopersicon esculentumLeafBiochemical [47]
Drought24-epiBL/28-homoBL [0.5–2 µM]Raphanus sativusSeedBiochemical [43]
Drought24-epiBL [1.0 μM]Capsicum annumLeafPhysiological and biochemical [48]
Drought24-epiBL [0.1 μM]Carthamus tinctoriusLeafBiochemical [49]
Drought24-epiBL [0–100 nM]Glycine maxSeedMorphological and biochemical [50]
Drought24-epiBL [0.4 μM]Triticum aestivumSeedBiochemical, hormonal and anatomical [51]
Drought24-epiBL [0.10 μM]Vitis viniferaLeafHormonal and biochemical [52]
Drought28-homoBL [0.01 μM]Brassica junceaLeafBiochemical [34]
DroughtBL [1 mg.L−1]Arachis hypogaeaSeedBiochemical [53]
DroughtBL [0.1 mg.L−1]Glycine maxLeafBiochemical and physiological [54]
DroughtBL [0–0.4 mg.L−1]Robinia pseudoacaciaLeafMorphological and biochemical [55]
DroughtBL [0–0.4mg.L−1]Xanthoceras sorbifoliaRootMorphological [56]
DroughtBiobras16 [16–0.1 mg. L−1]Carica papayaLeafMorphological and biochemical [57]
DroughtBR analogue [10% w/v]Passiflora edulis f. edulisSubstrateMorphological [58]
FloodingBL [1.0 mM]Zea maysLeafBiochemical and physiological [59]
High temperature24-epiBL [0.2 mg.L−1]Camellia sinensisLeafBiochemical [60]
High temperatureBL [10–6 M]Brassica NapusLeafHormonal [61]
High temperature24-epiBL [0.01–1.0 mg.L−1]Lycopersicon esculentumLeafPhysiological and biochemical [62]
Low temperatureBL [1–9 μM]Solanum lycopersicumFruitBiochemical [63]
Low temperature24-epiBL [0.1 μM]Cucumis sativusLeafBiochemical, physiological and morphological [64]
Low temperature28-homoBL [10–8–10–6 μM]Cucumis sativusLeafBiochemical, physiological and morphological [35]
Low temperature24-epiBL [10 μM]Vigna radiataLeafBiochemical and anatomical [65]
Low temperature24-epiBL [1.00 μM]Brassica NapusLeafBiochemical [66]
Low temperature24-epiBL [0.1 μM]Capsicum annuumLeafBiochemical, physiological and nutritional [37]
Low temperature24-epiBL [0.5 μM]Vitis viniferaSeedlingBiochemical, physiological and molecular [67]
NaClBRs [0–0.2 mg.L−1]Cucumis sativusRoot and leafBiochemical [68]
CaSO4 [5 t.ha−1]BL [0–0.3 mg.L−1]Trifolium alexandrinumLeafBiochemical and morpho-agronomical [69]
Imidacloprid [0.01–0.02%]24-epiBL [10−11–10−7 M]Oryza sativaSeedBiochemical [70,71]
NaCl24-epiBLCapsicum annuum Biochemical and nutritional [71]
NaCl [0–150mM]24-epiBL [0–0.2 mg.L−1]Lycopersicon esculentum Biochemical [72]
NaCl [0–150 mM]24-epiBL [0–2.5 mg.L−1]Mentha piperitaLeafBiochemical and morphological [73]
NaCl [0–250mM]24-epiBL [0–50 nM]Eucalyptus urophyllaLeafPhysiological and nutritional [74]
NaCl [0.3–0.4 M]24-epiBL [3 μM]Hordeum vulgareSeedAnatomical and genetics [75]
NaCl [0–100 mM]24-epiBL [0–3 μM]Lactuca
sativa
Seed and leafMorphological, anatomical and nutritional [76]
NaCl [0–100 mM]24-epiBL [10−4−1 µM]; 28-homoBL [10−4–1 µM]Zea maysSeedBiochemical and nutritional [77]
NaCl [0–100 mM]BL [0.25 mg.L−1]Zea maysSeed and leafBiochemical [78]
NaCl [100 mM]24-epiBL [0–100 mM]Glycine maxSeedNutritional [79]
NaCl [13.6 dS/m]BL [5 µM]Medicago sativaSeedMorphological and biochemical [80]
NaCl [150 mM]24-epiBL [0.01 μM]Cucumis sativusLeafMorphological, physiological and biochemical [35]
NaCl [150 mM]24-epiBL [0.013–0.038 mg.L−1]Triticum aestivumLeafMorphological, physiological and biochemical [81]
NaCl [2%]24-epiBL [0.4 µM]Triticum aestivumSeedHormonal [82]
NaCl [200 mM]BL [0.005–0.05 mg.L−1]Cucumis sativusRoot and leafBiochemical, physiological and morphological [83]
NaCl [250 mM]24-epiBL [0–10 µM]Cucumis sativusSeedBiochemical and molecular [84]
NaCl [25–100 mM]28-homoBL [10−11–10−7 M]Zea maysSeedMorphological and biochemical [85]
NaCl [25–150 mM]BL [0.05 mg.L−1]Vigna unguiculataLeafBiochemical [86]
NaCl [35 mM]24-epiBL [0–1 μM]Fragaria x ananassaLeafNutritional, physiological and biochemical [87]
NaCl [400 mM]24-epiBL [10−7–0.5×10−9 M]Cajanus cajanSubstrateBiochemical, physiological, anatomical and nutritional [88]
NaCl [50–150 mM]28-HomoBL [10−10–10−6 M]Brassica junceaLeafMorpho-agronomical [89]
NaCl [75–150 mM]24-epiBL [1–100 nM]Pisum sativum Morphological and biochemical [90]
NaCl [90 mM]24-epiBL [0–0.2 mg.L−1]Solanum melongenaLeafMorphological, biochemical and nutritional [91]
NaCl, CaCl2, MgSO4 [4.7–9.4 dS/m]24-epiBL [0–0.1 mg.L−1]Triticum aestivumLeafNutritional and biochemical [92]
NaCl [500 mM]24-epiBL [0–2.0 μM]Kandelia obovataRootBiochemical [93]
24-epiBL = 24-epibrassinolide; 28-HomoBL = 28-homobrassinolide; BL = brassinolide; BRs = brassinosteroids.
Table 2. Brassinosteroids and analogues applied in crops, with respective concentrations and effects.
Table 2. Brassinosteroids and analogues applied in crops, with respective concentrations and effects.
Commercial NameBrassinosteroids/
Analogue
ConcentrationEffects
1 BrassinolideBrassinolide0.5–1.0 mg L−1Higher rice yield (10%)
2 Biobras-16®Spirostanic analogue of castasterone0.4 mg L−1Increased seed yield (68%)
3 CIDEF-4(80% steroid compound)0.3 mg L−1Higher yield in fertile maize hybrids (2.5 times more)
4 DI-31 (BB16)Spirostane BR analogues30 mg ha−1Increase in lettuce production (25.93%)
4 DI-100Spirostane BR analogues30 mg ha−1Increase in lettuce production (31.08%)
5 BrassinazoleTriazole compound5.0 µMIncreased firmness and cellulose content in persimmon fruits
6 24-eBL
(Phyto Technology Laboratories, US).
24-epibrassinolide0.2 mg L−1Increased yield per vine, berry weight, cluster weight and specific gravity
Source: 1 [111]; 2 [155]; 3 [156]; 4 [157]; 5 [158]; 6 [159]. Adapted from [154].
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

Silva, A.P.S.d.; Alencar, A.A.d.S.; Sudré, C.P.; Araújo, M.d.S.B.d.; Lobato, A.K.d.S. Brassinosteroids: Relevant Evidence Related to Mitigation of Abiotic and Biotic Stresses in Plants. Agronomy 2024, 14, 840. https://doi.org/10.3390/agronomy14040840

AMA Style

Silva APSd, Alencar AAdS, Sudré CP, Araújo MdSBd, Lobato AKdS. Brassinosteroids: Relevant Evidence Related to Mitigation of Abiotic and Biotic Stresses in Plants. Agronomy. 2024; 14(4):840. https://doi.org/10.3390/agronomy14040840

Chicago/Turabian Style

Silva, Aminthia Pombo Sudré da, Antônio André da Silva Alencar, Cláudia Pombo Sudré, Maria do Socorro Bezerra de Araújo, and Allan Klynger da Silva Lobato. 2024. "Brassinosteroids: Relevant Evidence Related to Mitigation of Abiotic and Biotic Stresses in Plants" Agronomy 14, no. 4: 840. https://doi.org/10.3390/agronomy14040840

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