Next Article in Journal
Advancing Fruit Preservation: Ecofriendly Treatments for Controlling Fruit Softening
Previous Article in Journal
Morphological and Molecular Characterization of the Potato Rot Nematode, Ditylenchus destructor, Parasitizing Garlic in Korea
Previous Article in Special Issue
The Mitigating Effects of Biostimulant Amendments on the Response of Purslane Plants Grown under Drought Stress Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Drought Stress Effects and Ways for Improving Drought Tolerance in Impatiens walleriana Hook.f.—A Review

by
Marija Milovančević
*,
Milana Trifunović-Momčilov
,
Olga Radulović
,
Snežana Milošević
and
Angelina Subotić
Institute for Biological Research “Siniša Stanković”, National Institute of Republic of Serbia, Department for Plant Physiology, University of Belgrade, Bulevar despota Stefana 142, 11108 Belgrade, Serbia
*
Author to whom correspondence should be addressed.
Horticulturae 2024, 10(9), 903; https://doi.org/10.3390/horticulturae10090903
Submission received: 18 July 2024 / Revised: 19 August 2024 / Accepted: 25 August 2024 / Published: 26 August 2024
(This article belongs to the Special Issue Horticultural Production under Drought Stress)

Abstract

:
Drought is one of the main abiotic stresses affecting plant growth and development. Reduced plant yield and quality are primarily caused by the reductions in photosynthesis, mineral uptake, metabolic disorders, damages from the increased production of reactive oxygen species, and many other disruptions. Plants utilize drought resistance mechanisms as a defense strategy, and the systems’ activation is dependent upon several factors, including plant genotype, onthogenesis phase, drought intensity and duration, and the season in which the drought occurs. Impatiens walleriana is a worldwide popular flowering plant recognized for its vibrant flower colors, and is an indispensable plant in pots, gardens and other public areas. It prefers well-draining, moisturized soil, and does not perform well in overly dry or waterlogged conditions. Consequently, inadequate water supply is a common problem for this plant during production, transportation, and market placement, which has a substantial impact on plant performance overall. This review article outlines certain features of morphological, physiological, and molecular alterations induced by drought in ornamental, drought-sensitive plant species I. walleriana, as well as research carried out to date with the aim to improve the drought tolerance. Stress proteins aquaporins and dehydrins, whose molecular structure was described for the first time in this plant species, are highlighted specifically for their role in drought stress. Furthermore, the effective improvement of drought tolerance in I. walleriana by exogenous application of Plant Growth Regulators and Plant Growth-Promoting Bacteria is discussed in detail. Finally, this review can provide valuable insights for improving plant resilience and productivity in the face of water scarcity, which is critical for sustainable agriculture and horticulture.

1. General Introduction

The term drought usually refers to the physical lack of water in the substrate caused by an insufficient amount of atmospheric precipitation or an intense process of transpiration in plants [1]. However, there are other different types and definitions of drought [2]. One of them is meteorological drought, which characterizes drought as deficiency of water in soil, plants and atmosphere. A hydrological drought is characterized by a shortage of water in streams and rivers, typically measured by the number of days when the water level or discharge falls below a predetermined threshold. Agricultural drought is characterized as deficiency of water that negatively affects crop production. On the other hand, physiological drought is the state of soil and water in plants that limits plant growth and production. The relationship between physiological and other types of drought is not clear-cut; a meteorological drought does not always imply a physiological or hydrological drought. As a result, the type of plant and, in particular, its ontogenesis stage determines the physiological drought stage. Physiological drought could occur as an effect of low temperatures, which inhibit the activity of the root system. In addition, an increased concentration of salt in the substrate or flooding can negatively affect water uptake. At the global level, drought is a major problem and a limiting factor in the production of numerous crops and commercially important plant species. Drought induces tissue dehydration, inhibition of cell division, elongation and differentiation, oxidative stress manifested by increment in reactive oxygen species (ROS), and the slowing down of basic physiological and metabolic processes such as photosynthesis, transpiration, and mineral uptake, thus leading to decrease in overall plant growth and yield (Figure 1) [3,4]. The sensitivity of a plant species to drought depends on the genotype, ontogenetic stage of development, additional stress factors, frequency, intensity and drought duration [1,3,5]. Plant species respond to drought stress with different resistance mechanisms. Drought resistance mechanisms include drought escape or “ephemeroid strategy”, drought avoidance, and drought tolerance (Figure 1) [6,7]. In the genus Impatiens, there are proposed mechanisms of selection for drought stress tolerance and avoidance in Impatiens capensis, depending on the season when drought occurs [8]. Plants react to the negative effects of drought using different strategies within the resistance mechanisms, at the physiological, biochemical and molecular level. Depending on the genotype, we can talk about drought-tolerant and drought-intolerant plant species [9,10]. However, prolonged and intense drought can lead to death of even the most tolerant plant species, and considering this, various measures are taken in the production of commercially important plant species whose productivity can be limited by drought [11,12,13,14,15].
Drought escape or “ephemeroid strategy” is observed in plants at the end of the season when there is not enough rainfall in the habitat, so the plants are forced to finish their life cycle faster [16]. Drought avoidance implies the maintenance of high water potential in tissues through rapid accumulation of abscisic acid (ABA) which affects stomatal closure and transpiration reduction [6]. However, drought avoidance is very often related to the activation of drought tolerance mechanisms, especially in cases where drought periods are intense and prolonged [17]. Drought tolerance implies the plant’s ability to maintain physiological processes and functions even at a lower water potential in tissues. The accumulation of osmoprotectants, components of antioxidant protection, proteins with a protective role, transporters and various products of secondary metabolism, genes expression modulation and changes in phytohormonal signalization, underly the tolerance of plants to drought [6,7,18,19,20,21]. The accumulation of amino acid proline is often considered as drought stress indicator, since it has roles as osmoprotectant and antioxidant therewithal [22]. In addition to proline, the components with strong antioxidant activity are different phenolic compounds [23], as well as photosynthetic pigments such as carotenoids [24]. The antioxidant enzymes such as superoxide dismutase (SOD), catalase (CAT), peroxidase (POX), ascorbate-gluthatione cycle enzymes (dehydroascorbate reductase—DHAR, monodehydroascorbate reductase—MDHAR, and glutathione reductase—GR) represent the main classes of enzymatic antioxidants and some of the most investigated components of the antioxidant defense system in general [7]. The cause of all changes at the physiological and biochemical levels lies in the expression of certain genes during drought. The perception of a drought stimulus by the receptor initiates a signaling cascade that leads to the activation/deactivation of genes associated with the plants’ responses to drought stress. Gene expression can be associated with the synthesis of ion transporters, water channels—aquaporins (AQPs), stress proteins, and enzymes involved in the biosynthesis of osmoprotectants and antioxidant components, as well as many transcription factors that regulate the expression of other genes [20]. In a complex network of interactions, the altered phytohormones concentrations under drought play a major role in the regulation of gene expression and coordination of the plant physiological response [21].

2. Impatiens walleriana—Species Characteristics

Impatiens walleriana belongs to the genus Impatiens (Balsaminaceae), which includes more than 1000 species and is one of the largest genera among flowering plants [25]. Species of the genus Impatiens are mainly herbaceous, annual or perennial, with laying or upright shoots, while woody representatives of this genus are rare [26]. The Impatiens species are the most abundant in the tropical parts of Africa, Madagascar, the Himalayan region, India and Sri Lanka. They represent the typical members of tropical and subtropical forests at altitudes of 500 to 800 m, and sometimes up to 5000 m [25]. In addition, the species can be found in the temperate regions of Asia, Europe and North America [27]. Although there are some exceptions, most of Impatiens species do not tolerate prolonged drought conditions or prolonged exposure to direct sunlight. Therefore, these species are mostly restricted in their distribution to wet habitats, such as tropical rain forests, places next to rivers, streams or wetlands [25]. Due to their beautiful appearance and long flowering period, many species of the genus Impatiens are cultivated as decorative plants worldwide.
I. walleriana (Figure 2) is an annual herb, and represents one of the most popular species of the Impatiens genus. The height of the plant ranges between 30 and 70 cm. The shape of the leaves is simple, with serrated ends and a spiral arrangement, but, often at the tips of the shoots there are leaves facing each other, i.e., oppositely arranged. The flowers are zygomorphic, with three sepals and five sepals [27]. The corolla consists of one upright petal, and the remaining four are grouped into two lateral pairs. Both corolla and sepals exhibit variability in size, shape and color, as an adaptation to different pollinators [26]. Fruits are berries or explosive capsules. The color of the flower can be red, white, orange, purple and pink. The plants are very branched, with a high water content, which makes them tender and fragile. Impatiens species require a great presence of water in the substrate. The lack of water leads to a rapid decrease in the cells’ turgor and dehydration of the tissues. Inadequate water supply in the process of plant production and during transport to sales facilities can slow down growth and affect the decorative properties of I. walleriana. Because of the relationship between their morphological characteristics and their native environments, proper cultivation of plants in the context of water availability requires careful consideration. I. walleriana is one of the three plant species (Impatiens hawkeri and Impatiens balsamina) of the genus Impatiens which have been commercially produced in Serbia for many years, and, due to its decorative properties and long flowering period, it is considered one of the most popular horticultural species around the world.
The earliest investigations concerning I. walleriana started more than 15 years ago in our laboratory. In one of the first reports, a successful regeneration protocol for I. walleriana in vitro was conducted using urea-type cytokinins [28]. Subsequent studies were focused on the production of Tomato spotted wilt virus—free plants using the meristem-tip culture technique in vitro [29,30], as well genetic transformation in order to obtain the virus-resistant plants [31,32]. As will be covered in more details in the following paragraphs, the majority of the following studies mainly focused on I. walleriana responses to drought and ways to improve the drought-tolerance. The aim of this review is to explore and summarize the current advancements in understanding and improving drought tolerance in I. walleriana through various biotechnological approaches. Given the plant’s popularity and sensitivity to water stress, this review will synthesize findings related to the plant’s responses to drought, with an emphasis on identifying effective strategies for enhancing its resilience in conditions of inadequate water supply. Ultimately, this review aims to provide valuable insights for improving plant growth and development under water scarcity, which could be of importance also for other plant species in light of the improvement in their tolerance to abiotic stress.

3. Morphological Changes in Drought-Stressed I. walleriana

The first reaction of plants to drought is a biophysical response based on changes at the cellular level [33]. Plant growth is impaired primarily by inhibition of cell division, elongation and differentiation, due to loss of cell turgor and reduced water potential [1,3,33]. Cell growth is considered as one of the most sensitive physiological processes under stressful conditions. Decreasing the water potential in cells during dehydration reduces the value of turgor, which under usual physiological conditions exerts pressure on the cell wall and increases its plasticity, enabling cell growth. By inhibiting the process of growth and development of individual cells, the growth and development of organs and therefore the organism as a whole is reduced [1]. The number and size of leaves per plant are reduced, which directly affects the reduced intensity of photosynthesis and the assimilation of nutrients [4]. Therefore, growth is reduced, as well as the fresh and dry weight of plants. Depending on the intensity and duration of the dry period, which is in nature usually related with high temperatures and intensity, many plants twist the leaves (leaf rolling) or change their orientation [34,35]. In this way, the water released into the atmosphere through transpiration is reduced. On the other hand, drought can contribute to increasing the water use efficiency (WUE) by the root system. Namely, when plants suffer a lack of water in the substrate, the growth relationship between the aerial part and the root system in rice is in favor of the roots, in order to improve the absorption of water and nutrients [36]. Faster growth of the root system increases the possibility for the plant to reach water in the deeper layers of the soil where available water exists, thus mitigating the negative effect of drought on growth and development [37]. Certainly, the growth of the root system during drought depends on the plant genotype, but also on the intensity and duration of the drought [38]. Plant species that have a shallow root system, such as potatoes, are more sensitive to the effects of drought compared to plant species with a deeper root system, such as corn or clover [38].
Several studies conducted on I. walleriana in the last 20 years, have showed its sensitivity to drought stress regarding the changes in growth and development. One of the earliest studies pointed that I. wallerana Hook ‘Rosebud Pink’ variety was one of the most dehydration-intolerant herbaceous plants among 30 tested taxa (ornamental tree, shrub, and herbaceous annual and perennial species), exposed to water withholding until they reached the lethal point (fewer than eight live leaves remained) [39]. Two years later, a study was conducted in which drought stress in I. walleriana was simulated using different concentrations of polyethylene glycol (PEG8000). The results indicated an altered morphology and nutrient uptake in hydroponically grown I. walleriana ‘Dazzler pink’ [40]. Namely, plants were shorter, narrower, and with reduced root length when grown in solutions containing increasing PEG8000 concentrations. A following study included the evaluation of drought stress responses of 17 ornamental herbaceous plants, among which was I. walleriana ‘Tempo White’ variety [41]. Five irrigation regimes were tested (0, 25, 50, 75 and 100% of the reference evapotranspiration—ET0, required for optimum Kentucky bluegrass growth), and the authors concluded that I. walleriana ‘Tempo White’ performed poorest among all the tested genera. That means that I. walleriana plants grow well only under 100% ET0, with respect to the measurement of plant biomass, changes in percent cover, and visual evaluations. A year later, the authors in [42] offered a more thorough analysis of I. walleriana’s ‘Deep Rose’ reactions to drought, including changes in morphological features. The experimental design included three irrigation points—80 (control), 60 (mild drought) and 30% (severe drought) of soil water content (SWC), with three 10-day drought cycles separated by a 10-day period of standard watering. Drought stress slightly reduced plant height, increased root length under 30% SWC, and significantly decreased number of flowers at both drought intensities. Another research study [43] described the effects of competition and reduced water content within the container on I. walleriana ‘Cajun Violet’ growth and flowering. The results indicated that reduced water content significantly reduced height, shoot number, dry weight, and flower number of I. walleriana. In another interesting report, it has been concluded that water stress did not significantly affect the growth of the I. walleriana ‘DeZire’ variety, despite an obvious tendency to decrease plant height, number of flower buds, and fresh and dry weight [44]. The water stress effect on I. walleriana ‘DeZire’ growth did not show statistical significance, which the authors attributed to the experimental design that used degrees of freedom that are more conservative in the test of the complete plot component. Polyethylene glycol-imposed drought (0, 1, 2, and 3% PEG8000) in vitro reduced plant height, fresh weight, and the number of leaves and shoots per plant of I. walleriana ‘Busy Lizzie’, with the strongest effect of 3% PEG8000 [45]. Similar results were described where 3% PEG8000 decreased fresh weight and proliferation rate of in vitro-grown I. walleriana ‘Xtreme Scarlet’ [46]. Drought imposed to potted I. walleriana ‘Accent Premium Red’ plants grown in the plant growth chamber induced wilting and floral abortion [47], while drought-stressed I. walleriana ‘Xtreme Scarlet’ plants showed a significant drop in water potential in leaves, which was reflected in reduced fresh and dry weight, as well as total leaf area [48,49]. In two I. walleriana cultivars, the drought imposed in the greenhouse significantly reduced plant height, the number of leaves per plant, leaf area, leaf fresh and dry weight, stem diameter, stem fresh and dry weight, the number of axillary shoots, root volume, root fresh and dry weight, and also total plant fresh and dry weight [50]. Despite that, by monitoring flowering, the authors noticed that drought reduced the number, diameter and longevity of flowers. Within the genus Impatiens, the negative consequences of drought stress in combination with high temperature were noted also for Impatiens glandulifera morphological traits [51]. Namely, drought in combination with high temperatures induced leaf senescence and decreased leaf number, as well as dropping flower production and shortening their life span. In one similar research study using different Impatiens clones, the authors defined drought-tolerance among their characteristics [52]. More drought-tolerant clones were less affected by the stress, had less reduced fresh and dry weight, and produced more flowers than the others. One research study also indicated leaf and stem anatomical anomalies in Impatiens balsamina under drought stress [53].
Figure 3 shows morphological differences between well-watered and drought-stressed I. walleriana shoots and roots. It can be observed that drought induced leaves wilting and rolling, and significantly reduced plant growth and flowering. On the other hand, drought caused obvious differences in root growth, where drought-stressed I. walleriana had better root development as a mechanism for plant adaptation to drought. The similar results for better root growth of drought-stressed I. walleriana were described earlier [42].

4. Physiological Changes in Drought-Stressed I. walleriana

4.1. Photosyntesis, Photosyntetic Pigments and Osmotic Adjustment

Inhibited growth and division of cells due to loss of turgidity leads to disruption of numerous physiological and biochemical processes such as photosynthesis, respiration, uptake of nutrients and other metabolic processes [1,3,54,55]. The greatest impact of drought can be seen through the reduction in the photosynthesis process. The plant’s ability to adapt to changed environmental conditions is directly or indirectly related to changes at the level of the photosynthetic apparatus. Lower activity of the photosynthetic apparatus is a common consequence of water deficiency in plant tissues, due to both limited stomatal conductivity and numerous damages attributed to “non-stomatal mechanisms” [1,54]. During dehydration, plants close their stomata by the action of the ABA, in order to limit the release of water through transpiration, and at the same time limit the influx of CO2. Non-stomatal mechanisms include disruption of the biosynthesis of the main photosynthetic pigment—chlorophyll, functional and structural changes of chloroplasts, photosynthetic enzymes and disruption of the process of accumulation, transport and distribution of assimilates [1]. Both mechanisms are equally important and the dominance of one over the other depends on the plant species, developmental stage, and intensity and duration of stress, as well as the influence of other external factors. Stomatal closure ensures that plants reduce leaf water loss (LWL), and maintain a certain relative water content (RWC), which usually drops because of water deficiency. Despite that, since photosynthesis is altered, it is also common to expect changes in plant nutritional composition, since the production of assimilates is reduced.
Lack of water in the substrate reduces the concentration of photosynthetic pigments and the activity of photosynthetic enzymes, which results in disruption of the entire electron transport chain. It is known that the concentration of chlorophyll is positively correlated with the intensity of photosynthesis [1]. The decrease in chlorophyll concentration during drought is a direct consequence of photo-oxidation and degradation of pigments [3,33]. Most research on the effects of drought on the concentration of photosynthetic pigments indicates a reduction in the concentration of chlorophyll in plant leaves [56,57]. Depending on the genotype and sensitivity to drought, in different plant species the concentration of chlorophyll was reduced or even increased during drought [58,59,60]. On the other hand, carotenoids, as auxiliary photosynthetic pigments, play a role in protecting the photosynthetic apparatus from oxidative stress [61]. In addition, carotenoids are precursors in the biosynthesis of the phytohormone ABA, which plays a very important role in the mechanisms of plant resistance to drought [61]. In many plant species, increased concentration of carotenoids during drought is detected [62,63], or even decreased carotenoid concentration [64,65,66].
With decreasing water potential in cells and tissue, plants start to accumulate substances with an osmotic regulation role [67]. The accumulation of such compounds lowers the water potential inside the cell, and thus creates the difference in water potential between the cell and its surroundings. Accordingly, water is allowed to be taken in by the cells, since the external water potential is higher and provides water flow through the concentration gradient. One of the most investigated substances with an osmoprotection role is amino acid proline, which is often considered as an indicator of drought stress tolerance in plants [68,69]. Proline is an essential amino acid whose accumulation in plants increases under the various abiotic stressogenic factors, including drought. The effect of proline during tissue dehydration caused by drought can be twofold. As an osmoprotectant, proline affects the change in the osmotic and water potential in the cells and thus promotes the transport of water from a place of higher to a place of lower water potential. As an antioxidant, proline can “catch” free radicals or prevent the production of ROS [70,71].
Research conducted on I. walleriana points at drought-sensitivity regarding some physiological processes, such as nutrient composition changes. It has been described that PEG8000-treated I. walleriana ‘Dazzler pink’ had less zink, copper, nitrogen and calcium, and more phosphorus and nickel in the leaves, despite showing no visible nutritional deficiency symptoms [40]. In accordance with this, it has been concluded that changes in I. walleriana morphology were influenced by PEG8000 decreasing the water potential of the substrate and water absorption in that way. Mild and severe drought stress (60 and 30% SWC) caused a decrease in RWC and pigment content (chlorophyll a + b) in I. walleriana ‘Deep Rose’, especially during plant exposure to 30% SWC [42]. In contrast, drought-stressed I. walleriana increased free amino acid and ammonium content in leaves under the increasing water stress and after every drought period. However, since these parameters were also increased in control plants, it is concluded that the increment range was more time than drought dependent. The similar results were described for in vitro-grown I. walleriana ‘Busy Lizzie’ subjected to increasing concentration of PEG8000, which caused significant reduction in RWC, chlorophyll and total pigment content, as well as an increment in LWL [45]. On the other hand, the plants exposed to PEG8000 had considerably higher proline content in the leaves, indicating the role of proline in osmotic adjustment and osmoprotection. Slightly different results were described for I. walleriana ‘Xtreme Scarlet’ subjected to drought in the growth chamber, with a noticeable reduction in proline content, but, on the other hand, increased total chlorophyll and carotenoids, indicating that different physical conditions, stress duration and variety could affect plant response to drought [48]. Similar results regarding the increment in pigment composition in I. walleriana leaves were described three years later [72]. The same authors reported similar results for in vitro-grown I. walleriana ‘Xtreme Scarlet’ exposed to PEG8000 induced drought, and indicated significantly increased total chlorophyll and carotenoid content and proline, as well as total amino acid content [46]. Additionally, a significant increment in proline accumulation in two I. walleriana cultivars ‘Tempo’ and ‘Salmon’ subjected to drought was described, while plants showed lower pigment content, including chlorophyll a and b, total chlorophyll and carotenoids [50]. Among the genus Impatiens, it has been described that drought-tolerance of some Impatiens clones was related to the stomatal closure mechanisms which prevent excessive water release, but, at the same time, it did not negatively affect plant growth [52]. Similarly, drought-tolerance of invasive Impatiens parviflora was correlated with high WUE and adjusted water potential, which provides adequate water status in the cells [73].

4.2. Changes in Endogenous Abscisic Acid

The generally accepted role of ABA during the exposure of plants to various stressogenic factors leading to tissue dehydration characterized this hormone as a “stress hormone”. In mechanisms to avoid drought, ABA plays a key role because it affects stomatal closure and prevents water loss from the leaves through transpiration.
The endogenous concentration of ABA in plant tissue is regulated by biosynthesis, degradation, translocation and conjugation with other compounds. During tissue dehydration, ABA is synthesized de novo, while during tissue rehydration the concentration of ABA decreases. The key gene in ABA biosynthesis is NCED (9-cis-epoxycarotenoid dioxygenase), which encodes an enzyme with a role in converting cis-neoxanthin and cis-violaxanthin to xanthoxin [74]. In addition to NCED, the AAO (abscisic aldehyde oxidase) gene, which encodes the enzyme responsible for the last step in the synthesis of the biologically active form of ABA in cytoplasm, plays a very important role, that is, converting ABA aldehyde to ABA. Among genes involved in biosynthesis, the genes responsible for ABA catabolism are also important, since the phytohormone concentration is dependent on both biosynthesis and catabolism. ABA degradation is catalyzed by a cytochrome P450 monooxygenase (P450) encoded by the CYP707As gene family, in which ABA is converted to phaseic acid [74]. During tissue dehydration, it was shown that NCED gene expression increased in Cistus creticus, followed by increased endogenous ABA concentration [75]. Similar results were described for barley [76], and for three different carrot cultivars, in which the expression of NCED1 and NCED2 genes was significantly increased during tissue dehydration [77]. Also, it was shown that increased overexpression of NCED3 gene in soybean improved drought tolerance and increased the expression of AAO3, as well as the genes involved in ABA signal transduction [78]. The reduced expression of catabolic gene CYP707A2 under drought was described in Petunia × hybrida, but without changes in the expression of biosynthetic genes NCED1, NCED2, AAO31 and AAO32 [79]. Similar findings were reported in [80], demonstrating that when A. thaliana was subjected to dehydration, decreased expression of the ABA catabolic genes CYPT707A1/3 was detected.
Similarly to the available literature data, the changes in ABA metabolic genes in drought-stressed I. walleriana were investigated. Namely, the changes in I. walleriana ‘Xtreme Scarlet’ ABA metabolic genes (two genes involved in biosynthesis—IwNCED4 and IwAAO2, and one gene involved in catabolism—ABA8ox3, which belongs to the CYP707As gene family) upon exposure to different drought stress regimes in a growth chamber, were described [48]. The obtained results indicated that drought altered the expression level of these genes mainly by increasing it, which was followed by increased ABA content in I. walleriana leaves. The level of ABA in I. walleriana leaves increased by 2.92 and 4.3 times compared to control plants grown in moderate and intense drought, indicating an ABA role in mediating drought signal, depending on the stress intensity.

4.3. Changes in Stress-Related Proteins

Specifically, our group carried out two important studies on the molecular characterization of stress-related proteins—AQPs and dehydrins (DHNs), as well as the variations in their gene expression linked to I. walleriana drought-tolerance responses.

4.3.1. Aquaporins

Aquaporins (AQPs) are transmembrane proteins, which belongs to the Major Intrinsic Protein (MIP) family, and are responsible for water and small-molecule transport across membranes [81,82,83]. They are crucial for plant adaptation to environmental alterations which imply changes in water status in plant tissues, and many researches have pointed to their significant role [84,85,86,87]. With the exception of intracellular bacteria and thermophilic archaea, MIPs are found in every kingdom of living organisms [82,83,88]. All MIPs are composed of six transmembrane domains, with the protein’s N and C terminal ends found in the cytoplasm. Transmembrane regions are connected with five loops, of which two highly conserved ones (loops B and E) carry a conserved Asn-Pro-Ala (NPA) motif, found deep in the pore. Indeed, the most noticeable aquaporin trademark sequence is the NPA motif, which forms a pore for water and other solutes transport. Studies on biochemistry and functionality have also indicated that aquaporins are active in tetrameric formation, including four monomers in a holoprotein structure [83]. Plants comprise five AQP subfamilies, separated by intracellular localization and the functionality of their isoforms [82,83]. The subfamily with the highest number of members is represented by the Plasma membrane Intrinsic Proteins (PIPs), and, as implied by the subfamily name, PIPs are mainly found in the plasma membrane [83,88]. Two evolutionary groups, PIP1 and PIP2, comprise the PIP subfamily, and their water permeability properties, as well N- and C-terminal lengths, vary. Usually, PIP2 members produce a higher water permeability compared to PIP1 members. Another very important subfamily for regulation of water transport is the Tonoplast Intrinsic Proteins (TIPs), whose members are localized on the vacuole membrane—the tonoplast. Given the role of the vacuole in controlling turgor and osmotic adjustment in cells, the presence and activity of TIP isoforms in water transport across the vacuole is crucial [83,88].
Molecular characterization and expression profiling of four AQP isoforms in drought-stressed I. walleriana ‘Xtreme Scarlet’ have been performed [89]. Since members of the PIP and TIP subfamilies are recognized for their capacity for water transport, I. walleriana isoforms, which belong to these families, were selected for inclusion in the investigation. The sequenced leaf transcriptome of I. walleriana (RNASeq) was searched for sequences, and several bioinformatics methods were used to perform that. Gel electrophoresis was used to confirm that four of the identified isoforms (IwPIP1;4, IwPIP2;2, IwPIP2;7, and IwTIP4;1) exhibited the stable expression, and those isoforms were further tested in expression analysis. Before the expression analysis, detailed insight into molecular structure (sequence length, number of amino acids, molecular weight, isoelectric point, instability index, subcellular localization, transmembrane helices, stereochemical properties and 3D models of structures) of I. walleriana AQPs was performed using different available bioinformatical tools, and in Figure 4, in the first row, 3D models of IwAQPs holoproteins are presented. The authors explained in detail the IwAQPs’ structure, pointing to the crucial characteristics of MIP family members. All of the analyzed IwAQPs included the tetrameric structure of holoprotein, conserved NPA motifs, and transmembrane helices (TMh’s). Additionally, the isoelectric point corresponded to their predicted subcellular localization, while the instability index indicated the stable structures. A 3D model of the pore’s characteristics was obtained, and this model showed that the composition of amino acid residues in the pore of I. walleriana aquaporins vary, which could affect hydropathy and suggests that different solutes may be transported. Stereochemical analysis showed that most of the amino acids of the analyzed I. walleriana AQPs fall into regions which are energetically allowed, while the holoprotein structure of all IwAQPs indicated the presence of four monomers in each.
Drought stress and recovery affected the AQP gene expression in I. walleriana leaves differently [89]. The expression analysis of genes coding I. walleriana AQPs have revealed that the IwPIP2;7 gene was highly responsive to drought and recovery from drought, demonstrating its role in drought resistance mechanisms. On the other hand, it was concluded that other three AQP isoforms (IwPIP1;4, IwPIP2;2, and IwTIP4;1) could also be important in recovery after stress, maintaining the optimal water flow through the cells.

4.3.2. Dehydrins

Dehydrins are thermostable and highly hydrophilic proteins which are involved in plant responses to drought stress [90,91,92]. They could protect cell membranes and proteins from dehydration, thus protecting membrane integrity, regulating ROS homeostasis, antioxidant activity, photosynthetic pigment content and photosynthesis rates and osmotic adjustment, as well as morphological traits [93,94,95]. Dehydrins belong to group 2 of LEA (Late Embryogenesis Proteins) proteins, and are widely distributed in plants [91,96]. They consist of various conservative motifs such as K, Y and S and, according to their presence, could be divided in five structural subgroups—Kn, SKn, KnS, YnSKn, and YnKn [96]. The model of the DHN protection role of proteins under water deficit was proposed, indicating two possible mechanism [97]. In the first one, DHNs protect proteins via a molecular shield, in which the presence of DHN molecules maintains the quantity and arrangement of water molecules necessary to preserve the integrity of the enzymes without necessarily interacting with them. In the second one, DHNs must interact with proteins, as monomers, dimers or in any other oligomeric form.
The following study provided the insight into molecular structure of the dehydrin isoform characterized in the I. walleriana transcriptome [47]. In the I. walleriana transcriptome, three dehydrin isoforms—IwDhn1, IwDhn2.1 and IwDhn2.2—were detected (Figure 4, the second row), and their architecture, length, number of amino acids and protein isoelectric point were characterized [46]. In terms of IwDHNs architecture, there are two types: IwDhn2.1 (Y3SK1) and IwDhn2.2 (Y3SK2), which have a YnSKn composition, and IwDhn1, which is an SKn type. Upon exposure to drought in the plant growth chamber, I. walleriana ‘Accent Premium Red’ had slightly increased IwDhn1 gene expression, while the expression of other two isoforms—IwDhn2.1 and IwDhn2.2—was strongly induced [47]. Authors have concluded that the IwDhn1 protein is important, because of its almost constitutive expression under normal and stress conditions, but its expression profile implies that it is not crucial for drought stress protection in I. walleriana. Based on its structure (SK2 type), this isoform is not dehydration-inducible, but more likely to be responsive to other types of abiotic stresses, according to the literature data [47]. On the other hand, based on the strong induction of isoforms IwDhn2.1 and IwDhn2.2 during drought, it could be assumed that they have a role in plants for coping with stress. This could be explained by the specificities in their architecture and structure, and presence of amino acids which provide membrane binding and protection.

5. Biochemical Changes in Drought-Stressed I. walleriana

Reactive Oxygen Species and Antioxidant Defense

Regarding the drought effects on a plant’s growth and development, one of the most studied is secondary oxidative stress, i.e., the excessive production of ROS, which overwhelms the capacity of the antioxidant system to remove those chracteristics [98]. Depending on their nature and quantity, ROS can be toxic to the plants, but can also play a role as secondary messengers in the initiation of signaling pathways that activate defense mechanisms in plants [99]. The most frequently mentioned ROS are singlet oxygen (O21), hydrogen peroxide (H2O2), superoxide anion radical (O2•−), hydroxyl radical (OH•), perhydroxyl radical (HO2•), peroxyl radical (RO2•) and nitrogen monoxide (NO). Among the mentioned ROS, O21, H2O2, O2 and OH• are the most represented, with mutual differences in stability, reactivity and the possibility of transport through the cell membranes [100]. The consequences of oxidative stress are damage to the protein and nucleic acid structures through oxidative modifications, as well as membrane lipid peroxidation, which leads to the accumulation of oxidative stress indicators such as malondialdehyde (MDA) [98]. Considering proteins, the oxidative modifications of methionine, thiol groups of cysteine, site-specific amino-acid modification, fragmentation of peptide chains, and the aggregation of cross-linked reaction products, as well as carbonylation, are the most frequently mentioned effects of ROS [101,102,103]. The oxidative modifications of DNA usually involve modifications of nitrogenous bases and breaking the sugars’ bonds [104]. Lipid peroxidation of cell membranes is one of the most investigated consequences of ROS action on the structure and function of membrane lipids [19,63,98]. Precursors for the products of lipid peroxidation are polyunsaturated fatty acids, the main components of membrane lipids. Through the phases of initiation, propagation and termination, polyunsaturated fatty acids react with ROS, forming the different intermediers and final products such as MDA [105]. Malondialdehyde is a tricarbon dialdehyde whose concentration is often analyzed, and is considered as an indicator of cell membrane damage during oxidative stress [105,106,107,108,109]. Reactive oxygen species production in plants was confirmed in chloroplasts, peroxisomes, mitochondria, cell membranes and the apoplast [110].
Antioxidant enzymes are responsible for removing ROS and detoxifying plant tissues under stressful conditions. Among them, SOD, POX, CAT, DHAR, MDHAR, and GR represent the main classes of enzymatic antioxidants [111]. Superoxide dismutase catalyzes the reaction of converting O2 to H2O2 in chloroplasts, peroxisomes, cytoplasm and mitochondria [110]. Based on the presence of metals as cofactors and sites of activity, four SOD isoforms have been detected in living organisms: Fe-SOD, Mn-SOD, Cu/Zn-SOD and Ni-SOD, while the Ni-SOD isoform was not found in higher plants [110,111]. Catalase degrades H2O2 to water and oxygen primarily in peroxisomes, but also in cytoplasm and mitochondria [110,111], and, until now, three genes encoding CAT isoforms (CAT1-3) have been identified in the A. thaliana genome. Peroxidase, such as ascorbate peroxidase (APX), DHAR, MDHAR, and GR maintain H2O2 balance in the cell via the ascorbate-glutathione cycle [110,111,112]. Firstly, APX converts H2O2 to H2O by oxidizing ascorbate to monodehydroascorbate (MDHA), and ascorbate can be regenerated by the enzyme MDHAR, which uses NAD(P)H, as a cofactor. Monodehydroascorbate can spontaneously convert to dehydroascorbate (DHA), and the enzyme DHAR, which oxidizes glutathione (GSH) to glutathione disulfide (GSSG), regenerates ascorbate from dehydroascorbate. In the final step, the enzyme GR regenerates GSH from GSSG using NAD(P)H as a cofactor. Among components with strong antioxidant potential there are also some non-enzymatic compounds, such as phenolics [113,114]. Phenolics represent a group of secondary plant metabolites that contain aromatic rings with one or more hydroxyl groups in their structure [113]. The antioxidant activity of phenolic compounds is reflected by their ability to “capture” free radicals, to donate electrons or protons, or to bind metals that participate in the production of ROS. By binding metals, they can prevent, for example, the Fenton reaction, which produces the most reactive forms of oxygen [113,115].
Several studies with I. walleriana have shown that drought can cause oxidative stress and alterations in the antioxidant defense system. The changes in oxidative status were usually manifested as an increment in oxidative stress indicators—H2O2, MDA—while on the other hand, the activities of antioxidant enzymes such as SOD, POX and CAT were also altered. In one such report, the authors documented a significant increment of H2O2 and MDA in I. walleriana ‘Busy Lizzie’ leaves exposed to increasing PEG8000 concentration in vitro [45]. However, the highest increment in oxidative stress indicators was detected when plants were grown on media with the highest PEG8000 concentration (3%). Due to disturbed oxidative status in cells, increased activities of antioxidant enzymes—SOD, POX, and CAT—were determined, with the highest values observed in plants exposed to strongest drought stress, namely 3% PEG8000. In addition, 3% PEG8000 caused the induction of all eight POX isoforms, but particularly isoforms B and I. As a response to drought induced by PEG8000 (3%) in vitro, in I. walleriana ‘Xtreme Scarlet’ leaves the increased content of total polyphenol content, H2O2, and MDA was also detected [46]. Recently the authors [47] conducted an experiment with potted I. walleriana ‘Accent Premium Red’ exposed to drought in the plant growth chamber, and recorded the increased MDA content, as well as the induction of chloroplastic Cu/ZnSOD and two peroxidase isoforms. Also, upon exposure of I. walleriana ‘Xtreme Scarlet’ to drought stress in a growth chamber, an increment in total polyphenol and flavonoid content was detected in leaves, as well as the increment in total antioxidant capacity [48]. Such results indicated the significant role of phenolic compounds in plant responses to drought stress. Additionally, drought-stressed plants exhibited changes in oxidative stress indicators and antioxidant enzymes. Namely, depending on the drought intensity, an increment in H2O2 and MDA was observed, as well as the increased activity of SOD, POX, and CAT. Drought-stressed plants expressed one Mn-SOD and two Cu/Zn-SOD (A and B) isoforms, similar to those in control plants, while a total of four POX isoforms were detected and, among them, the D isoforms was specific to non-stressed I. walleriana plants. Three years later, the same authors described similar results for drought-stressed I. walleriana ‘Xtreme Scarlet’ grown in a growth chamber, and observed increased H2O2 and MDA content, as well as increased activity of antioxidant enzymes and phenolic content [72]. In similar research, as responses to drought, I. walleriana increased the activity of enzymatic antioxidants (guaiacol peroxidase—GPX, pyrogallol peroxidase—PPX, ascorbate peroxidase—APX, and CAT) and total phenolic and flavonoid contents, as well as the total antioxidant capacity of leaves [116]. Drought stress also increased the MDA, POX and APX in two I. walleriana cultivars grown under controlled conditions in a greenhouse [50].

6. Mitigation Strategies for Drought Stress

Elicitation is the process of applying certain compounds with the aim of inducing the signaling pathways associated with the stimulating of any type of plant defense [117]. Based on the elicitor’s nature, the common classification includes two groups—abiotic and biotic elicitors [118,119]. However, inside these two major groups of elicitors there are some differences in classification of individual groups of certain compounds. For example, according to one classification [118], abiotic elicitors include physical, chemical and hormonal factors, while biotic elicitors include substances of biological origin, such as plant cell-wall polysaccharides and microorganisms. In contrast to this classification, another one [119] classifies hormonal factors in a biotic group of elicitors as substances generated in living organisms, namely plants.
Plant hormones are naturally occurring, low-molecular-weight substances produced in plants. Phytohormones are responsible for coordinating all aspects of plant growth and development, including the responses to abiotic and biotic stress factors. As synthetic compounds, plant hormones are commonly used in the elicitation process to improve plant defense against different stressors. Along with some other compounds, synthetic plant hormones are usually referred to as Plant Growth Regulators (PGRs) [120,121]. The most used and investigated PGRs in plant drought tolerance are abscisic acid (ABA), salycilic acid (SA), and jasmonates—jasmonic acid (JA), methyl jasmonate (MeJA), and jasmonil–isoleucine (JA-Ile). Undoubtedly, Plant Growth Promoting Bacteria (PGPB) are the biotic elicitors that require the greatest consideration in future research. Interestingly, recently isolated Pseudomonas, Pantoea, Acinetobacter and Chryseobacterium bacteria from the rhizosphere of various Mediterranean medicinal plants (Thymus, Sarcocornia, Mentha), are hypothesized to increase the drought and salinity tolerance of these plant species through indole3-acetic-acid (IAA) production, phosphorus solubilization, and ACC (1-aminocyclopropane-1-carboxylate) deaminase activity [122]. It is estimated that as much as 80% of rhizosphere-associated bacteria are the source of IAA [122,123]. Since mechanisms of drought and salinity tolerance often overlap, the total number of PGPB that alleviate drought is probably much higher. In our laboratory, we identified four rhizosphere-associated Pseudomonas strains with the dual ability to produce and degrade IAA, and monitored their PGPB effects on duckweed (Lemna minor) [124]. The most promising is P. gessardii C31-106/3, which increases biomass production, modulation of duckweeds’ antioxidant enzymatic activity, and reduction in H2O2 content [125]. In fact, there is an abundance of studies with different PGPB strains that may also have the potential to improve drought stress tolerance in ornamental plants, including I. walleriana (see Section The PGRs and PGPB Application for Drought-Tolerance Improvement in I. walleriana). Table 1 presents the selected publications from the last decade describing the effects of ABA, SA, jasmonates, and PGPBs on plant drought-tolerance improvement, conducted in controlled ex vitro or in vitro conditions.

The PGRs and PGPB Application for Drought-Tolerance Improvement in I. walleriana

One of the earliest reports that described I. walleriana drought-tolerance improvement was conducted fourteen years ago [163]. In this experiment, drench and spray ABA (0 or 500 mg_L−1) applications were implemented for several drought-stressed bedding plants, including I. walleriana ‘Xtreme Lavender’. The application of ABA significantly delayed visible wilting symptoms and increased shelf life of I. walleriana, with more effectiveness with drench application than spraying. The next study evaluated the effect of a commercial extract of Giant Knotweed (Fallopia sachalinensis F. Schmidt) on drought tolerance of I. walleriana ‘Super Elfin XP [164]. The commercial extract, named Regalia®, was foliar-applied in different concentrations and under different substrate volumetric water contents in two independent experiments. Foliar application of Regalia® at a certain concentration improved some morphological (root dry weight) and physiological attributes (soluble protein content, leaf greenness, net photosynthesis etc.) in I. walleriana. Despite this being observed, the authors concluded that these results could not be related to elevated drought tolerance because overall growth of plants was not enhanced in either of the two applied experiments. Two years later, the efficacy of several commercially available antitranspirants in enhancing drought stress tolerance in bedding plants, including I. walleriana ‘Double Fiesta Ole Purple Stripe’, was evaluated [165]. While three physiological antitranspirants—two sugar alcohol-based compounds (SACs) and a biologically active ABA (s-ABA)—were added into the substrate, two physical antitranspirants—β-pinene polymer (βP) and vinyl–acrylic polymer (VP) were sprayed on the plants grown in greenhouse conditions when they reached a marketable stage of at least one open flower per plant. After that, half of the plants in each group were water stressed until all treated plants attained a visual wilt status rating of three or lower, defined previously as unmarketable [166]. The results of the study indicated that βP and s-ABA were effective in enhancing shelf life and delaying wilting symptoms of I. walleriana, with more effectiveness noticed for s-ABA. Based on these data, it appears that foliar application of βP or drenching s-ABA could allow I. walleriana to withstand a water shortage during transport and/or retailing. The potential role of SA as a stress-ameliorating agent in I. walleriana ‘Buzzy Lizy’, grown in vitro under PEG-induced drought, was also proposed [45]. Salycilic acid in graded concentrations (0, 1, 2, and 3 mM) was added in media supplemented with PEG8000 (0, 1, 2, and 3%), and consequently enhanced growth and improved some physiological and biochemical parameters in drought-stressed I. walleriana. For this purpose, the authors recommended the use of 2–3 mM SA for improving I. walleriana drought tolerance, with no growth-retarding, but only beneficial, effects. In the followed experiment, the authors had chosen 2 mM of SA as foliar treatment for drought-stressed I. walleriana ‘Accent Premium Red’ grown in the plant growth chamber [47]. Salycilic acid was applied on the second day of the experimental setup to well-watered plants, and plants were further subjected to water deprivation until the fourteenth day of the experiment. In the same way as in the first two experimental groups, two more plant groups received distilled-water spraying instead of SA, and were also well-watered and drought stressed. Samples were taken from five different points: the start of the experiment, 24 h after SA/distilled-water application, on the 11th and 14th day of drought duration, and 24 h after rehydration. The study’s findings showed that SA had no effects on well-watered plants, but had ameliorating effects on drought-exposed plants. That was manifested by preventing wilting, preserving RWC, increasing proline accumulation, modulating antioxidative activities, and significantly lowering lipid peroxidation, but had no effect on flower preservation. Additionally, the application of SA altered the expression of three DHNs isoforms in plants under drought stress, with one isoform being slightly enhanced after the treatment. However, under drought, the alterations in dehydrin expression were the same as in plant foliar sprayed with distilled water, so the SA effects were not considered noteworthy. The authors considered this plant response as an SA drought-ameliorating effect, so DHNs are not needed as much as in drought-stressed plants without SA treatment.
Recently, it was described how PGPB can stimulate I. walleriana ‘Super Elfin Ruby’ growth under drought stress by enhancing plant drought tolerance [167]. Namely, two Pseudomonas sp. strains—P. poae 29G9 and P. fluorescens 90F12-2—stimulated the increment of shoot biomass and flower numbers of I. walleriana grown under low-nutrient conditions and after recovery from drought stress in greenhouse conditions.
Further investigations also refer to the potential role of SA in ameliorating drought effects. Recent research has also indicated the power of exogenous SA application on morpho-physiological and molecular characteristics of two I. walleriana cultivars (‘Tempo and Salmon’) grown under water deficit stress [50]. Three SA concentration (0, 1, and 2 mM) were applied weekly as foliar treatment to I. walleriana grown in a greenhouse under three levels of soil moisture content (95, 85, and 75% of field capacity). The results indicated that SA, especially in 2 mM concentration, improved the drought-tolerance of I. walleriana through increased plant height, leaf area, number of auxiliary shoots, root volume, stem diameter, number of leaves, chlorophyll content, and antioxidant enzymes activities, and on the other hand, reduced electrolyte leakage and oxidative damage. In addition, the authors noted that the Salmon cultivar is more drought-tolerant than the Tempo cultivar, and, accordingly, Salmon is recommended for cultivation in areas with low water supply. Additionally, the potential of foliar and drench application of a secondary messenger molecule, calcium, in increasing shelf life and delaying wilting symptoms, was reported in drought-stressed I. walleriana [168].
The latest investigations of our group have revealed the role of exogenously applied MeJA as a potential regulator of drought tolerance of in vitro- and ex vitro-grown I. walleriana ‘Xtreme Scarlet’ [46,49,72]. I. walleriana plants grown on media supplemented with 5 μM MeJA for seven days had better morpho-physiological performance during a further twenty days on the media supplemented with 3% PEG [46]. In water-stressed plants, pretreatment with 5 μM MeJA for seven days increased growth parameters (fresh weight, height, number of leaves per plant and proliferation rate) compared to control plants without MeJA pretreatment. In these plants, decreased photosynthetic pigment content was detected, as well as proline and total amino acid, total polyphenol and DPPH (1,1′-diphenyl-2-picrylhydrazyl) activity, and H2O2 and MDA content. Accordingly, the activities of antioxidant enzymes (SOD, POX, and CAT) were also affected. It is interesting to note that the application of 5 μM MeJA also provided the best growth (fresh weight and height) of non-stressed plants in comparison to the plant group grown on hormone-free media during the entire experimental period. Namely, when MeJA was applied in pretreatment for seven days, and, subsequently, in treatment for twenty days, plant growth was improved and visual characteristics of tested plants were the best. In this research, the authors also investigated the effects of pretreatment with higher MeJA concentration (50 and 100 μM) on drought ameliorating effect on I. walleriana, but it seems that these concentrations had mainly inhibitory effects. Based on the mentioned in vitro research, MeJA (5 and 50 μM) was applied as a foliar treatment to drought-stressed I. walleriana grown ex vitro in a growth chamber under controlled conditions [72]. The highest concentration (100 μM MeJA) was omitted in this experiment, since it was assumed that it would suppress I. walleriana growth. Foliar treatment with 20 mL of each MeJA concentration was applied seven days before and on the day when drought stress (15 and 5% SWC) was imposed. The control plants and one drought-stressed plant group were treated with distilled water. The obtained results indicated that foliar application of 50 μM MeJA had the strongest effect on physio-biochemical responses of drought-stressed plants, increasing pigment content, and decreasing oxidative stress in plant cells. Additionally, in drought-stressed plants sprayed with 50 µM MeJA increased expression of ABA metabolic genes was detected, while of the four analyzed aquaporin gene isoforms, the expression of IwPIP1;4 and IwPIP2;7 was strongly induced. Foliar application of 50 µM MeJA also increased the chlorophyll index and Nitrogene Balance Index (NBI) of I. walleriana at 5% SWC, indicating the effect of the elicitor on plant drought tolerance [49]. The latest research aimed to evaluate the potential role of sodium nitroprusside (SNP) on alleviating water deficit caused by PEG in vitro in I. walleriana ‘Xtreme Scarlet’, monitoring the changes at a physiological and biochemical level [169]. The findings of this investigation suggested that SNP might be advantageous for biochemical parameters when applied only in combination with PEG, but not solely, to I. walleriana shoots cultivated in vitro. The summarized effects of exogenously applied elicitors on I. walleriana drought-tolerance improvement are presented in the Figure 5.

7. Conclusions and Future Perspectives

Drought stress affects plants, and ways to cope with this abiotic stress factor have been reported in numerous studies in recent years. Whether they are conducted in the laboratory, greenhouse or in the field, undoubtedly all of them demonstrated the harmful effects of this stressor on different aspects of plant growth and development. On the other hand, plants have evolved drought resistance mechanisms, and, depending on the genotype, duration and stress intensity, respond differently at a morphological, physiological, biochemical, and molecular level. One of the most important ornamental plants in horticultural production, I. walleriana, known for its sensitivity to drought, has been the object of investigation in recent years. In the last 20 years, many studies have investigated the responses of I. walleriana plants to drought, and at the same time, the diverse ways to improve drought tolerance. Drought sensitivity of I. walleriana has been reported at different levels of organization, indicating also drought-tolerance dependence on the variety of plant. It has been reported that many exogenously applied chemicals could improve I. walleriana tolerance, such as hormones ABA, SA, and MeJA, as well as antitranspirants, calcium and PGPB. Additionally, improvements in understanding of I. walleriana drought-tolerance mechanisms has been achieved by molecular characterization and understanding the role of some crucial proteins in drought stress response—AQPs and DHNs.
Further investigations will focus on the phytohormonal network involved in drought-stress responses of I. walleriana, as well as finding additional ways for drought-tolerance improvement. A promising strategy is to test various PGPBs in improving the drought tolerance of I. walleriana, especially those strains that were tested on other plants. Moreover, the effects of other types of abiotic stress factors (e.g., osmotic stress, temperature stress) on I. walleriana growth and development should be examined in future studies.

Author Contributions

Conceptualization, A.S., M.T.-M. and M.M.; writing—original draft preparation, M.M.; writing manuscript parts about PGPBs, O.R.; writing—review and editing, A.S., M.T.-M., O.R. and S.M.; visualization, O.R. and S.M.; supervision, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data are included in the manuscript.

Acknowledgments

This research was supported by the Ministry of Science, Technological Development and Innovation of the Republic of Serbia, contract number: 451-03-66/2024-03/200007.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Murtaza, G.; Rasool, F.; Habib, R.; Javed, T.; Sardar, K.; Ayub, M.M.; Ayub, M.A.; Rasool, A. A review of morphological, physiological and biochemical responses of plants under drought stress conditions. Imp. J. Interdiscip. Res. 2016, 2, 1600–1606. [Google Scholar]
  2. Novák, V. Physiological drought—How to quantify it? In Bioclimatology and Natural Hazards. Springer: Dordrecht, The Netherlands, 2009; pp. 89–95. [Google Scholar] [CrossRef]
  3. Anjum, S.A.; Xie, X.; Wang, L.C.; Saleem, M.F.; Man, C.; Lei, W. Morphological, physiological and biochemical responses of plants to drought stress. Afr. J. Agric. Res. 2011, 6, 2026–2032. [Google Scholar]
  4. Anjum, S.A.; Ashraf, U.; Zohaib, A.; Tanveer, M.; Naeem, M.; Ali, I.; Tabassum, T.; Nazir, U. Growth and developmental responses of crop plants under drought stress: A review. Zemdirb. Agric. 2017, 104, 267–276. [Google Scholar] [CrossRef]
  5. Kapoor, D.; Bhardwaj, S.; Landi, M.; Sharma, A.; Ramakrishnan, M.; Sharma, A. The impact of drought in plant metabolism: How to exploit tolerance mechanisms to increase crop production. Appl. Sci. 2020, 10, 5692. [Google Scholar] [CrossRef]
  6. Basu, S.; Ramegowda, V.; Kumar, A.; Pereira, A. Plant adaptation to drought stress. F1000Research 2016, 5, 1554. [Google Scholar] [CrossRef]
  7. Laxa, M.; Liebthal, M.; Telman, W.; Chibani, K.; Dietz, K.J. The role of the plant antioxidant system in drought tolerance. Antioxidants 2019, 8, 94. [Google Scholar] [CrossRef]
  8. Heschel, M.S.; Riginos, C. Mechanisms of selection for drought stress tolerance and avoidance in Impatiens capensis (Balsaminaceae). Am. J. Bot. 2005, 92, 37–44. [Google Scholar] [CrossRef]
  9. You, J.; Zhang, Y.; Liu, A.; Li, D.; Wang, X.; Dossa, K.; Zhou, R.; Yu, J.; Zhang, Y.; Wang, L.; et al. Transcriptomic and metabolomic profiling of drought-tolerant and susceptible sesame genotypes in response to drought stress. BMC Plant Biol. 2019, 19, 267. [Google Scholar] [CrossRef]
  10. Li, H.; Mo, Y.; Cui, Q.; Yang, X.; Guo, Y.; Wei, C.; Yang, J.; Zhang, Y.; Ma, J.; Zhang, X. Transcriptomic and physiological analyses reveal drought adaptation strategies in drought-tolerant and-susceptible watermelon genotypes. Plant Sci. 2019, 278, 32–43. [Google Scholar] [CrossRef]
  11. Ashraf, M. Inducing drought tolerance in plants: Recent advances. Biotechnol. Adv. 2010, 28, 169–183. [Google Scholar] [CrossRef]
  12. Fang, Y.; Xiong, L. General mechanisms of drought response and their application in drought resistance improvement in plants. Cell. Mol. Life Sci. 2015, 72, 673–689. [Google Scholar] [CrossRef] [PubMed]
  13. Samota, M.K.; Sasi, M.; Awana, M.; Yadav, O.P.; Mithra, S.V.A.; Tyagi, A.; Kumar, S.; Singh, A. Elicitor-induced biochemical and molecular manifestations to improve drought tolerance in rice (Oryza sativa L.) through seed-priming. Front. Plant Sci. 2017, 8, 934. [Google Scholar] [CrossRef]
  14. Nataraja, K.N.; Dhanyalakshmi, K.H.; Govind, G.; Oelmüller, R. Activation of drought tolerant traits in crops: Endophytes as elicitors. Plant Signal. Behav. 2022, 17, 2120300. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, H.; Sun, X.; Dai, M. Improving crop drought resistance with plant growth regulators and rhizobacteria: Mechanisms, applications, and perspectives. Plant Commun. 2022, 3, 100228. [Google Scholar] [CrossRef]
  16. Shavrukov, Y.; Kurishbayev, A.; Jatayev, S.; Shvidchenko, V.; Zotova, L.; Koekemoer, F.; De Groot, S.; Soole, K.; Langridge, P. Early flowering as a drought escape mechanism in plants: How can it aid wheat production? Front. Plant Sci. 2017, 8, 1950. [Google Scholar] [CrossRef]
  17. Kooyers, N.J. The evolution of drought escape and avoidance in natural herbaceous populations. Plant Sci. 2015, 234, 155–162. [Google Scholar] [CrossRef] [PubMed]
  18. Cheng, Z.; Dong, K.; Ge, P.; Bian, Y.; Dong, L.; Deng, X.; Li, X.; Yan, Y. Identification of leaf proteins differentially accumulated between wheat cultivars distinct in their levels of drought tolerance. PLoS ONE 2015, 10, e0125302. [Google Scholar] [CrossRef]
  19. Gharibi, S.; Tabatabaei, B.E.S.; Saeidi, G.; Goli, S.A.H. Effect of drought stress on total phenolic, lipid peroxidation, and antioxidant activity of Achillea species. Appl. Biochem. Biotechnol. 2016, 178, 796–809. [Google Scholar] [CrossRef]
  20. Kaur, G.; Asthir, B. Molecular responses to drought stress in plants. Biol. Plant. 2017, 61, 201–209. [Google Scholar] [CrossRef]
  21. Wahab, A.; Abdi, G.; Saleem, M.H.; Ali, B.; Ullah, S.; Shah, W.; Mumtaz, S.; Yasin, G.; Muresan, C.C.; Marc, R.A. Plants’ physio-biochemical and phyto-hormonal responses to alleviate the adverse effects of drought stress: A comprehensive review. Plants 2022, 11, 1620. [Google Scholar] [CrossRef]
  22. Signorelli, S. The fermentation analogy: A point of view for understanding the intriguing role of proline accumulation in stressed plants. Front. Plant Sci. 2016, 7, 1339. [Google Scholar] [CrossRef] [PubMed]
  23. Varela, M.C.; Arslan, I.; Reginato, M.A.; Cenzano, A.M.; Luna, M.V. Phenolic compounds as indicators of drought resistance in shrubs from Patagonian shrublands (Argentina). Plant Physiol. Biochem. 2016, 104, 81–91. [Google Scholar] [CrossRef]
  24. Dawood, M.G.; El-Awadi, M.E.S.; Sadak, M.S.; El-Lethy, S.R. Research article comparison between the physiological role of carrot root extract and β-carotene in inducing Helianthus annuus L. Drought Tolerance. Asian J. Biol. Sci. 2019, 12, 231–241. [Google Scholar] [CrossRef]
  25. Janssens, S.B.; Knox, E.B.; Huysmans, S.; Smets, E.F.; Merckx, V.S. Rapid radiation of Impatiens (Balsaminaceae) during Pliocene and Pleistocene: Result of a global climate change. Mol. Phylogenet. Evol. 2009, 52, 806–824. [Google Scholar] [CrossRef] [PubMed]
  26. Fischer, E. Balsaminaceae. In Flowering Plants·Dicotyledons: Celastrales, Oxalidales, Rosales, Cornales, Ericales; Springer: Berlin/Heidelberg, Germany, 2004; pp. 20–25. [Google Scholar] [CrossRef]
  27. Grey-Wilson, C. Hydrocera triflora, its floral morphology and relationship with Impatiens: Studies in Balsaminaceae: V. Kew Bulletin 1980, 35, 213–219. [Google Scholar] [CrossRef]
  28. Subotić, A.; Jevremović, S.; Cingel, A.; Milošević, S. Effect of urea—Type citokinins on axillary shoots regeneration of Impatiens walleriana L. Biotechnol. Biotechnol. Equip. 2008, 22, 817–819. [Google Scholar] [CrossRef]
  29. Milošević, S.; Simonović, A.; Cingel, A.; Jevremović, S.; Todorović, S.; Filipović, B.; Subotić, A. Response of antioxidative enzymes to long-term Tomato spotted wilt virus infection and virus elimination by meristem-tip culture in two Impatiens species. Physiol. Mol. Plant Pathol. 2012, 79, 79–88. [Google Scholar] [CrossRef]
  30. Milošević, S.; Subotić, A.; Cingel, A.; Jevremović, S.; Stanković, I.; Bulajić, A.; Krstić, B. Virus elimination from ornamental plants with the use of in vitro culture techniques. Pestic. Phytomed. 2012, 27, 203–211. [Google Scholar] [CrossRef]
  31. Milošević, S.; Simonović, A.; Cingel, A.; Nikolić, D.; Ninković, S.; Subotić, A. Introduction of dsRNA-specific ribonuclease pac1 into Impatiens walleriana provides resistance to Tomato spotted wilt virus. Sci. Hortic. 2013, 164, 499–506. [Google Scholar] [CrossRef]
  32. Milošević, S.; Cingel, A.; Subotić, A. Agrobacterium-mediated transformation of ornamental species: A review. Genetika 2015, 47, 1149–1164. [Google Scholar] [CrossRef]
  33. Fathi, A.; Tari, D.B. Effect of drought stress and its mechanism in plants. Int. J. Life Sci. 2016, 10, 1–6. [Google Scholar] [CrossRef]
  34. Saglam, A.; Kadioglu, A.; Demiralay, M.; Terzi, R. Leaf rolling reduces photosynthetic loss in maize under severe drought. Acta Bot. Croat. 2014, 73, 315–323. [Google Scholar] [CrossRef]
  35. Cal, A.J.; Sanciangco, M.; Rebolledo, M.C.; Luquet, D.; Torres, R.O.; McNally, K.L.; Henry, A. Leaf morphology, rather than plant water status, underlies genetic variation of rice leaf rolling under drought. Plant Cell Environ. 2019, 42, 1532–1544. [Google Scholar] [CrossRef]
  36. Xu, W.; Cui, K.; Xu, A.; Nie, L.; Huang, J.; Peng, S. Drought stress condition increases root to shoot ratio via alteration of carbohydrate partitioning and enzymatic activity in rice seedlings. Acta Physiol. Plant. 2015, 37, 9. [Google Scholar] [CrossRef]
  37. Yang, X.; Lu, M.; Wang, Y.; Wang, Y.; Liu, Z.; Chen, S. Response mechanism of plants to drought stress. Horticulturae 2021, 7, 50. [Google Scholar] [CrossRef]
  38. Lipiec, J.; Doussan, C.; Nosalewicz, A.; Kondracka, K. Effect of drought and heat stresses on plant growth and yield: A review. Int. Agrophys. 2013, 27, 463–477. [Google Scholar] [CrossRef]
  39. Augé, R.M.; Stodola, A.J.; Moore, J.L.; Klingeman, W.E.; Duan, X. Comparative dehydration tolerance of foliage of several ornamental crops. Sci. Hortic. 2003, 98, 511–516. [Google Scholar] [CrossRef]
  40. Burnett, S.; van Iersel, M.; Thomas, P. PEG-8000 alters morphology and nutrient concentration of hydroponic impatiens. HortScience 2005, 40, 1768–1772. [Google Scholar] [CrossRef]
  41. Henson, D.Y.; Newman, S.E.; Hartley, D.E. Performance of selected herbaceous annual ornamentals grown at decreasing levels of irrigation. HortScience 2006, 41, 1481–1486. [Google Scholar] [CrossRef]
  42. Chyliński, W.K.; Łukaszewska, A.J.; Kutnik, K. Drought response of two bedding plants. Acta Physiol. Plant. 2007, 29, 399–406. [Google Scholar] [CrossRef]
  43. Blanusa, T.; Vysini, E.; Cameron, R.W. Growth and flowering of Petunia and Impatiens: Effects of competition and reduced water content within a container. HortScience 2009, 44, 1302–1307. [Google Scholar] [CrossRef]
  44. Andersson, N.E. The influence of water stress and air velocity on growth of Impatiens walleriana and Petunia × hybrid. Sci. Hortic. 2011, 128, 146–151. [Google Scholar] [CrossRef]
  45. Antonić, D.; Milošević, S.; Cingel, A.; Lojić, M.; Trifunović-Momčilov, M.; Petrić, M.; Subotić, A.; Simonović, A. Effects of exogenous salicylic acid on Impatiens walleriana L. grown in vitro under polyethylene glycol-imposed drought. S. Afr. J. Bot. 2016, 105, 226–233. [Google Scholar] [CrossRef]
  46. Đurić, M.; Subotić, A.; Trifunović-Momčilov, M.; Milošević, S. Improvement of water deficit stress tolerance of Impatiens walleriana shoots grown in vitro by methyl jasmonate. Plant Cell Tissue Organ Cult. 2023, 154, 351–365. [Google Scholar] [CrossRef]
  47. Antonić, D.D.; Subotić, A.R.; Dragićević, M.B.; Pantelić, D.; Milošević, S.M.; Simonović, A.D.; Momčilović, I. Effects of exogenous salicylic acid on drought response and characterization of dehydrins in Impatiens walleriana. Plants 2020, 9, 1589. [Google Scholar] [CrossRef] [PubMed]
  48. Đurić, M.; Subotić, A.; Prokić, L.; Trifunović-Momčilov, M.; Cingel, A.; Vujičić, M.; Milošević, S. Morpho-physiological and molecular evaluation of drought and recovery in Impatiens walleriana grown ex vitro. Plants 2020, 9, 1559. [Google Scholar] [CrossRef]
  49. Đurić, M.; Subotić, A.; Prokić, L.; Trifunović-Momčilov, M.; Milošević, S. Foliar application of methyl jasmonate affects impatiens walleriana growth and leaf physiology under drought stress. Plant Signal. Behav. 2023, 18, 2219936. [Google Scholar] [CrossRef]
  50. Safari, M.; Mousavi-Fard, S.; Rezaei Nejad, A.; Sorkheh, K.; Sofo, A. Exogenous salicylic acid positively affects morpho-physiological and molecular responses of Impatiens walleriana plants grown under drought stress. Int. J. Environ. Sci. Technol. 2021, 19, 969–984. [Google Scholar] [CrossRef]
  51. Descamps, C.; Boubnan, N.; Jacquemart, A.L.; Quinet, M. Growing and flowering in a changing climate: Effects of higher temperatures and drought stress on the bee-pollinated species Impatiens glandulifera royle. Plants 2021, 10, 988. [Google Scholar] [CrossRef]
  52. Shintiavira, H.; Purba, A.E.; Kartikaningrum, S.; Koseki, A. Identifying Drought-Tolerant Impatiens Genotypes by Using Water Stress Treatment. Caraka Tani J. Sustain. Agric. 2023, 38, 40–52. [Google Scholar] [CrossRef]
  53. Todd, G.W.; Richardson, P.E.; Sengupta, S.P. Leaf and stem anatomical anomalies in a drought-susceptible species, Impatiens balsamina, under conditions of drought stress. Bot. Gaz. 1974, 135, 121–126. [Google Scholar] [CrossRef]
  54. Farooq, M.; Wahid, A.; Kobayashi, N.S.M.A.; Fujita, D.B.S.M.A.; Basra, S.M. Plant drought stress: Effects, mechanisms and management. Sustain. Agric. 2009, 23, 407–418. [Google Scholar]
  55. Ansari, W.A.; Atri, N.; Pandey, M.; Singh, A.K.; Singh, B.; Pandey, S. Influence of drought stress on morphological, physiological and biochemical attributes of plants: A review. Biosci. Biotechnol. Res. Asia 2019, 16, 697–709. [Google Scholar] [CrossRef]
  56. Mathobo, R.; Marais, D.; Steyn, J.M. The effect of drought stress on yield, leaf gaseous exchange and chlorophyll fluorescence of dry beans (Phaseolus vulgaris L.). Agric. Water Manag. 2017, 180, 118–125. [Google Scholar] [CrossRef]
  57. Shivakrishna, P.; Reddy, K.A.; Rao, D.M. Effect of PEG-6000 imposed drought stress on RNA content, relative water content (RWC), and chlorophyll content in peanut leaves and roots. Saudi J. Biol. Sci. 2018, 25, 285–289. [Google Scholar]
  58. Zaefyzadeh, M.; Quliyev, R.A.; BABAYEVA, S.; Abbasov, M.A. The effect of the interaction between genotypes and drought stress on the superoxide dismutase and chlorophyll content in durum wheat landraces. Turk. J. Biol. 2009, 33, 1–7. [Google Scholar] [CrossRef]
  59. Pirzad, A.; Shakiba, M.R.; Zehtab-Salmasi, S.; Mohammadi, S.A.; Darvishzadeh, R.; Samadi, A. Effect of water stress on leaf relative water content, chlorophyll, proline and soluble carbohydrates in Matricaria chamomilla L. J. Med. Plants Res. 2011, 5, 2483–2488. [Google Scholar]
  60. Dastborhan, S.; Ghassemi-Golezani, K. Influence of seed priming and water stress on selected physiological traits of borage. Folia Hortic. 2015, 27, 151–159. [Google Scholar] [CrossRef]
  61. Sun, T.; Rao, S.; Zhou, X.; Li, L. Plant carotenoids: Recent advances and future perspectives. Mol. Hortic. 2022, 2, 3. [Google Scholar] [CrossRef]
  62. Irani, S.; Majidi, M.M.; Mirlohi, A.; Zargar, M.; Karami, M. Assessment of drought tolerance in sainfoin: Physiological and drought tolerance indices. Agron. J. 2015, 107, 1771–1781. [Google Scholar] [CrossRef]
  63. Sarker, U.; Oba, S. Drought stress effects on growth, ROS markers, compatible solutes, phenolics, flavonoids, and antioxidant activity in Amaranthus tricolor. Appl. Biochem. Biotechnol. 2018, 186, 999–1016. [Google Scholar] [CrossRef] [PubMed]
  64. Goodarzian Ghahfarokhi, M.; Mansurifar, S.; Taghizadeh-Mehrjardi, R.; Saeidi, M.; Jamshidi, A.M.; Ghasemi, E. Effects of drought stress and rewatering on antioxidant systems and relative water content in different growth stages of maize (Zea mays L.) hybrids. Arch. Agron. Soil. Sci. 2015, 61, 493–506. [Google Scholar] [CrossRef]
  65. Khoyerdi, F.F.; Shamshiri, M.H.; Estaji, A. Changes in some physiological and osmotic parameters of several pistachio genotypes under drought stress. Sci. Hortic. 2016, 198, 44–51. [Google Scholar] [CrossRef]
  66. Okunlola, G.O.; Olatunji, O.A.; Akinwale, R.O.; Tariq, A.; Adelusi, A.A. Physiological response of the three most cultivated pepper species (Capsicum spp.) in Africa to drought stress imposed at three stages of growth and development. Sci. Hortic. 2017, 224, 198–205. [Google Scholar] [CrossRef]
  67. Ozturk, M.; Turkyilmaz Unal, B.; García-Caparrós, P.; Khursheed, A.; Gul, A.; Hasanuzzaman, M. Osmoregulation and its actions during the drought stress in plants. Physiol. Plant. 2021, 172, 1321–1335. [Google Scholar] [CrossRef]
  68. Mwadzingeni, L.; Shimelis, H.; Tesfay, S.; Tsilo, T.J. Screening of bread wheat genotypes for drought tolerance using phenotypic and proline analyses. Front. Plant Sci. 2016, 7, 182193. [Google Scholar] [CrossRef] [PubMed]
  69. de Oliveira, L.A.R.; Cardoso, M.N.; de Oliveira, A.C.A.; Machado, C.D.A.; Cardoso, B.T.; Muniz, A.D.S.; Ledo, A.D.S. Effects of In Vitro Drought Stress on Growth, Proline Accumulation and Antioxidant Defense in Sugarcane. 2018. Available online: https://www.ccsenet.org/journal/index.php/jas/article/view/73121 (accessed on 25 May 2024).
  70. Signorelli, S.; Coitiño, E.L.; Borsani, O.; Monza, J. Molecular mechanisms for the reaction between •OH radicals and proline: Insights on the role as reactive oxygen species scavenger in plant stress. J. Phys. Chem. B 2014, 118, 37–47. [Google Scholar] [CrossRef] [PubMed]
  71. Soares, C.; Carvalho, M.E.; Azevedo, R.A.; Fidalgo, F. Plants facing oxidative challenges—A little help from the antioxidant networks. Environ. Exp. Bot. 2019, 161, 4–25. [Google Scholar] [CrossRef]
  72. Đurić, M.; Subotić, A.; Prokić, L.; Trifunović-Momčilov, M.; Milošević, S. Alterations in Physiological, Biochemical, and Molecular Responses of Impatiens walleriana to Drought by Methyl Jasmonate Foliar Application. Genes 2023, 14, 1072. [Google Scholar] [CrossRef]
  73. Quinet, M.; Descamps, C.; Coster, Q.; Lutts, S.; Jacquemart, A.L. Tolerance to water stress and shade in the invasive Impatiens parviflora. Int. J. Plant Sci. 2015, 176, 848–858. [Google Scholar] [CrossRef]
  74. Chen, K.; Li, G.J.; Bressan, R.A.; Song, C.P.; Zhu, J.K.; Zhao, Y. Abscisic acid dynamics, signaling, and functions in plants. J. Integr. Plant Biol. 2020, 62, 25–54. [Google Scholar] [CrossRef] [PubMed]
  75. Munné-Bosch, S.; Falara, V.; Pateraki, I.; López-Carbonell, M.; Cela, J.; Kanellis, A.K. Physiological and molecular responses of the isoprenoid biosynthetic pathway in a drought-resistant Mediterranean shrub, Cistus creticus exposed to water deficit. J. Plant Physiol. 2009, 166, 136–145. [Google Scholar] [CrossRef]
  76. Zeng, X.; Bai, L.; Wei, Z.; Yuan, H.; Wang, Y.; Xu, Q.; Tang, Y.; Nyima, T. Transcriptome analysis revealed the drought-responsive genes in Tibetan hulless barley. BMC Genom. 2016, 17, 1–12. [Google Scholar] [CrossRef] [PubMed]
  77. Zhang, R.-R.; Wang, Y.-H.; Li, T.; Tan, G.-F.; Tao, J.-P.; Su, X.-J.; Xu, Z.-S.; Tian, Y.-S.; Xiong, A.-S. Effects of simulated drought stress on carotenoid contents and expression of related genes in carrot taproots. Protoplasma 2021, 258, 379–390. [Google Scholar] [CrossRef]
  78. Molinari, M.D.C.; Fuganti-Pagliarini, R.; Marin, S.R.R.; Ferreira, L.C.; Barbosa, D.D.A.; Marcolino-Gomes, J.; Oliveira, M.C.N.D.; Mertz-Henning, L.M.; Kanamori, N.; Takasaki, H.; et al. Overexpression of AtNCED3 gene improved drought tolerance in soybean in greenhouse and field conditions. Genet. Mol. Biol. 2020, 43, e20190292. [Google Scholar] [CrossRef]
  79. Kim, J.; Malladi, A.; Van Iersel, M.W. Physiological and molecular responses to drought in Petunia: The importance of stress severity. J. Exp. Bot. 2012, 63, 6335–6345. [Google Scholar] [CrossRef] [PubMed]
  80. Wang, Z.; Wang, F.; Hong, Y.; Yao, J.; Ren, Z.; Shi, H.; Zhu, J.K. The flowering repressor SVP confers drought resistance in Arabidopsis by regulating abscisic acid catabolism. Mol. Plant 2018, 11, 1184–1197. [Google Scholar] [CrossRef]
  81. Martínez-Ballesta, M.D.C.; Carvajal, M. Mutual interactions between aquaporins and membrane components. Front. Plant Sci. 2016, 7, 1322. [Google Scholar] [CrossRef]
  82. Maurel, C.; Boursiac, Y.; Luu, D.T.; Santoni, V.; Shahzad, Z.; Verdoucq, L. Aquaporins in plants. Physiol. Rev. 2015, 95, 1321–1358. [Google Scholar] [CrossRef]
  83. Kapilan, R.; Vaziri, M.; Zwiazek, J.J. Regulation of aquaporins in plants under stress. Biol. Res. 2018, 51, 1–11. [Google Scholar] [CrossRef]
  84. Grondin, A.; Mauleon, R.; Vadez, V.; Henry, A. Root aquaporins contribute to whole plant water fluxes under drought stress in rice (Oryza sativa L.). Plant Cell Environ. 2016, 39, 347–365. [Google Scholar] [CrossRef] [PubMed]
  85. Patel, J.; Mishra, A. Plant aquaporins alleviate drought tolerance in plants by modulating cellular biochemistry, root-architecture, and photosynthesis. Physiol. Plant. 2021, 172, 1030–1044. [Google Scholar] [CrossRef] [PubMed]
  86. Nicolas-Espinosa, J.; Carvajal, M. Genome-wide identification and biological relevance of broccoli aquaporins. Plant Genome 2022, 15, e20262. [Google Scholar] [CrossRef]
  87. Lopez-Zaplana, A.; Martinez-Garcia, N.; Carvajal, M.; Bárzana, G. Relationships between aquaporins gene expression and nutrient concentrations in melon plants (Cucumis melo L.) during typical abiotic stresses. Environ. Exp. Bot. 2022, 195, 104759. [Google Scholar] [CrossRef]
  88. Kaldenhoff, R.; Fischer, M.J.A.P. Aquaporins in plants. Acta Physiol. 2006, 187, 169–176. [Google Scholar] [CrossRef] [PubMed]
  89. Đurić, M.J.; Subotić, A.R.; Prokić, L.T.; Trifunović-Momčilov, M.M.; Cingel, A.D.; Dragićević, M.B.; Simonović, A.D.; Milošević, S.M. Molecular characterization and expression of four aquaporin genes in Impatiens walleriana during drought stress and recovery. Plants 2021, 10, 154. [Google Scholar] [CrossRef]
  90. Hassan, N.M.; El-Bastawisy, Z.M.; El-Sayed, A.K.; Ebeed, H.T.; Alla, M.M.N. Roles of dehydrin genes in wheat tolerance to drought stress. J. Adv. Res. 2015, 6, 179–188. [Google Scholar] [CrossRef]
  91. Sun, Z.; Li, S.; Chen, W.; Zhang, J.; Zhang, L.; Sun, W.; Wang, Z. Plant dehydrins: Expression, regulatory networks, and protective roles in plants challenged by abiotic stress. Int. J. Mol. Sci. 2021, 22, 12619. [Google Scholar] [CrossRef]
  92. Riyazuddin, R.; Nisha, N.; Singh, K.; Verma, R.; Gupta, R. Involvement of dehydrin proteins in mitigating the negative effects of drought stress in plants. Plant Cell Rep. 2022, 41, 519–533. [Google Scholar] [CrossRef]
  93. Yang, Z.; Sheng, J.; Lv, K.; Ren, L.; Zhang, D. Y2SK2 and SK3 type dehydrins from Agapanthus praecox can improve plant stress tolerance and act as multifunctional protectants. Plant Sci. 2019, 284, 143–160. [Google Scholar] [CrossRef]
  94. Guo, X.; Zhang, L.; Wang, X.; Zhang, M.; Xi, Y.; Wang, A.; Zhu, J. Overexpression of Saussurea involucrata dehydrin gene SiDHN promotes cold and drought tolerance in transgenic tomato plants. PLoS ONE 2019, 14, e0225090. [Google Scholar] [CrossRef] [PubMed]
  95. Meng, Y.-C.; Zhang, H.-F.; Pan, X.-X.; Chen, N.; Hu, H.-F.; Haq, S.U.; Khan, A.; Chen, R.-G. CaDHN3, a pepper (Capsicum annuum L.) dehydrin gene enhances the tolerance against salt and drought stresses by reducing ROS accumulation. Int. J. Mol. Sci. 2021, 22, 3205. [Google Scholar] [CrossRef] [PubMed]
  96. Kosová, K.; Prášil, I.T.; Vítámvás, P. Role of dehydrins in plant stress response. In Handbook of Plant and Crop Stress, 4th ed.; CRC Press: Boca Raton, FL, USA, 2019; pp. 175–196. [Google Scholar]
  97. Cuevas-Velazquez, C.L.; Rendón-Luna, D.F.; Covarrubias, A.A. Dissecting the cryoprotection mechanisms for dehydrins. Front. Plant Sci. 2014, 5, 108360. [Google Scholar] [CrossRef] [PubMed]
  98. Hussain, S.; Rao, M.J.; Anjum, M.A.; Ejaz, S.; Zakir, I.; Ali, M.A.; Ahmad, N.; Ahmad, S. Oxidative stress and antioxidant defense in plants under drought conditions. In Plant Abiotic Stress Tolerance; Hasanuzzaman, M., Hakeem, K., Nahar, K., Alharby, H., Eds.; Springer: Cham, Switzerland, 2019. [Google Scholar] [CrossRef]
  99. Furlan, A.L.; Bianucci, E.; Castro, S. Signaling role of ROS in modulating drought stress tolerance. In Drought Stress. Tolerance in Plants, Vol. 1: Physiology and Biochemistry; Springer: Cham, Switzerland, 2016; pp. 309–330. [Google Scholar] [CrossRef]
  100. Waszczak, C.; Carmody, M.; Kangasjärvi, J. Reactive oxygen species in plant signaling. Annu. Rev. Plant Biol. 2018, 69, 209–236. [Google Scholar] [CrossRef] [PubMed]
  101. Pyngrope, S.; Bhoomika, K.; Dubey, R.S. Oxidative stress, protein carbonylation, proteolysis and antioxidative defense system as a model for depicting water deficit tolerance in Indica rice seedlings. Plant Growth Regul. 2013, 69, 149–165. [Google Scholar] [CrossRef]
  102. Waszczak, C.; Akter, S.; Jacques, S.; Huang, J.; Messens, J.; Van Breusegem, F. Oxidative post-translational modifications of cysteine residues in plant signal transduction. J. Exp. Bot. 2015, 66, 2923–2934. [Google Scholar] [CrossRef]
  103. AnAnjum, N.A.; Sofo, A.; Scopa, A.; Roychoudhury, A.; Gill, S.S.; Iqbal, M.; Lukatkin, A.S.; Pereira, E.; Duarte, A.C.; Ahmad, I. Lipids and proteins—Major targets of oxidative modifications in abiotic stressed plants. Environ. Sci. Pollut. Res. 2015, 22, 4099–4121. [Google Scholar] [CrossRef]
  104. Maurya, A.K. Oxidative stress in crop plants. In Agronomic Crops: Volume 3: Stress. Responses and Tolerance; Springer: Singapore, 2020; pp. 349–380. [Google Scholar] [CrossRef]
  105. Asthir, B. Protective mechanisms of heat tolerance in crop plants. J. Plant Interact. 2015, 10, 202–210. [Google Scholar] [CrossRef]
  106. Chugh, V.; Kaur, N.; Gupta, A.K. Evaluation of oxidative stress tolerance in maize (Zea mays L.) seedlings in response to drought. Indian J. Biochem. Biophys. 2011, 48, 47–53. [Google Scholar]
  107. Moharramnejad, S.A.J.J.A.D.; Sofalian, O.; Valizadeh, M.; Asghari, A.; Shiri, M.; Ashraf, M.U.H.A.M.M.A.D. Response of maize to field drought stress: Oxidative defense system, osmolytes’ accumulation and photosynthetic pigments. Pak. J. Bot. 2019, 51, 799–807. [Google Scholar] [CrossRef]
  108. Zhang, C.; Shi, S.; Liu, Z.; Yang, F.; Yin, G. Drought tolerance in alfalfa (Medicago sativa L.) varieties is associated with enhanced antioxidative protection and declined lipid peroxidation. J. Plant Physiol. 2019, 232, 226–240. [Google Scholar] [CrossRef]
  109. Zhang, Y.; Luan, Q.; Jiang, J.; Li, Y. Prediction and utilization of malondialdehyde in exotic pine under drought stress using near-infrared spectroscopy. Front. Plant Sci. 2021, 12, 735275. [Google Scholar] [CrossRef] [PubMed]
  110. Noctor, G.; Reichheld, J.P.; Foyer, C.H. ROS-related redox regulation and signaling in plants. In Seminars in Dell & Developmental Biology; Academic Press: Cambridge, MA, USA, 2018; Volume 80, pp. 3–12. [Google Scholar] [CrossRef]
  111. Dvořák, P.; Krasylenko, Y.; Zeiner, A.; Šamaj, J.; Takáč, T. Signaling toward reactive oxygen species-scavenging enzymes in plants. Front. Plant Sci. 2021, 11, 618835. [Google Scholar] [CrossRef]
  112. Kapoor, D.; Singh, S.; Kumar, V.; Romero, R.; Prasad, R.; Singh, J. Antioxidant enzymes regulation in plants in reference to reactive oxygen species (ROS) and reactive nitrogen species (RNS). Plant Gene 2019, 19, 100182. [Google Scholar] [CrossRef]
  113. Naikoo, M.I.; Dar, M.I.; Raghib, F.; Jaleel, H.; Ahmad, B.; Raina, A.; Khan, F.A.; Naushin, F. Role and regulation of plants phenolics in abiotic stress tolerance: An overview. In Plant Signaling Molecules; Elsevier: Amsterdam, The Netherlands, 2019; pp. 157–168. [Google Scholar]
  114. Parvin, K.; Nahar, K.; Mohsin, S.M.; Al Mahmud, J.; Fujita, M.; Hasanuzzaman, M. Plant phenolic compounds for abiotic stress tolerance. In Managing Plant Production under Changing Environment; Springer: Singapore, 2022; pp. 193–237. [Google Scholar]
  115. Kumar, M.; Tak, Y.; Potkule, J.; Choyal, P.; Tomar, M.; Meena, N.L.; Kaur, C. Phenolics as plant protective companion against abiotic stress. In Plant Phenolics in Sustainable Agriculture; Springer: Singapore, 2020; Volume 1, pp. 277–308. [Google Scholar]
  116. Matijević, A.; Šakonjić, A.; Murtić, S. Antioxidant response of Impatiens walleriana L. to drought. Acta Agric. Slov. 2022, 118, 1–7. [Google Scholar] [CrossRef]
  117. Angelova, Z.; Georgiev, S.; Roos, W. Elicitation of plants. Biotechnol. Biotechnol. Equip. 2006, 20, 72–83. [Google Scholar] [CrossRef]
  118. Naik, P.M.; Al-Khayri, J.M. Impact of abiotic elicitors on in vitro production of plant secondary metabolites: A review. J. Adv. Res. Biotechnol. 2016, 1, 1–7. [Google Scholar] [CrossRef]
  119. Thakur, M.; Bhattacharya, S.; Khosla, P.K.; Puri, S. Improving production of plant secondary metabolites through biotic and abiotic elicitation. J. Appl. Res. Med. Aromat. Plants 2019, 12, 1–12. [Google Scholar] [CrossRef]
  120. Rademacher, W. Plant growth regulators: Backgrounds and uses in plant production. J. Plant Growth Regul. 2015, 34, 845–872. [Google Scholar] [CrossRef]
  121. Agudelo-Morales, C.E.; Lerma, T.A.; Martínez, J.M.; Palencia, M.; Combatt, E.M. Phytohormones and Plant Growth Regulators—A Review. J. Sci. Technol. Appl. 2021, 10, 27–65. [Google Scholar] [CrossRef]
  122. LeLeontidou, K.; Genitsaris, S.; Papadopoulou, A.; Kamou, N.; Bosmali, I.; Matsi, T.; Madesis, P.; Vokou, D.; Karamanoli, K.; Mellidou, I. Plant growth promoting rhizobacteria isolated from halophytes and drought-tolerant plants: Genomic characterisation and exploration of phyto-beneficial traits. Sci. Rep. 2020, 10, 14857. [Google Scholar] [CrossRef]
  123. Nordstedt, N.P.; Jones, M.L. Isolation of rhizosphere bacteria that improve quality and water stress tolerance in greenhouse ornamentals. Front. Plant Sci. 2020, 11, 549301. [Google Scholar] [CrossRef] [PubMed]
  124. Popržen, T.; Nikolić, I.; Krstić-Milošević, D.; Uzelac, B.; Trifunović-Momčilov, M.; Marković, M.; Radulović, O. Characterization of the IAA-Producing and-Degrading Pseudomonas Strains Regulating Growth of the Common Duckweed (Lemna minor L.). Int. J. Mol. Sci. 2023, 24, 17207. [Google Scholar] [CrossRef] [PubMed]
  125. Popržen, T.; Jevremović, S.; Milošević, S.; Đurić, M.; Uzelac, B.; Stanković, S.; Radulović, O. Antioxidative Response of Duckweed (Lemna minor L.) to Rhizosphere-Associated Pseudomonas Strains and Exogenous Indole-3-Acetic Acid. Horticulturae 2024, 10, 562. [Google Scholar] [CrossRef]
  126. Du, Y.L.; Wang, Z.Y.; Fan, J.W.; Turner, N.C.; He, J.; Wang, T.; Li, F.M. Exogenous abscisic acid reduces water loss and improves antioxidant defence, desiccation tolerance and transpiration efficiency in two spring wheat cultivars subjected to a soil water deficit. Funct. Plant Biol. 2013, 40, 494–506. [Google Scholar] [CrossRef] [PubMed]
  127. Dias, M.C.; Oliveira, H.; Costa, A.; Santos, C. Improving elms performance under drought stress: The pretreatment with abscisic acid. Environ. Exp. Bot. 2014, 100, 64–73. [Google Scholar] [CrossRef]
  128. Zhou, L.; Xu, H.; Mischke, S.; Meinhardt, L.W.; Zhang, D.; Zhu, X.; Li, X.; Fang, W. Exogenous abscisic acid significantly affects proteome in tea plant (Camellia sinensis) exposed to drought stress. Hortic. Res. 2014, 1, 14029. [Google Scholar] [CrossRef]
  129. Abdelaal, K.A. Effect of salicylic acid and abscisic acid on morpho-physiological and anatomical characters of faba bean plants (Vicia faba L.) under drought stress. J. Plant Prod. 2015, 6, 1771–1788. [Google Scholar] [CrossRef]
  130. Wei, L.; Wang, L.; Yang, Y.; Wang, P.; Guo, T.; Kang, G. Abscisic acid enhances tolerance of wheat seedlings to drought and regulates transcript levels of genes encoding ascorbate-glutathione biosynthesis. Front. Plant Sci. 2015, 6, 458. [Google Scholar] [CrossRef]
  131. Li, Z.; Yu, J.; Peng, Y.; Huang, B. Metabolic pathways regulated by abscisic acid, salicylic acid and γ-aminobutyric acid in association with improved drought tolerance in creeping bentgrass (Agrostis stolonifera). Physiol. Plant. 2017, 159, 42–58. [Google Scholar] [CrossRef]
  132. Mohammadi, M.H.S.; Etemadi, N.; Arab, M.M.; Aalifar, M.; Arab, M.; Pessarakli, M. Molecular and physiological responses of Iranian Perennial ryegrass as affected by Trinexapac ethyl, Paclobutrazol and Abscisic acid under drought stress. Plant Physiol. Biochem. 2017, 111, 129–143. [Google Scholar] [CrossRef]
  133. Nawaz, M.; Wang, Z. Abscisic acid and glycine betaine mediated tolerance mechanisms under drought stress and recovery in Axonopus compressus: A new insight. Sci. Rep. 2020, 10, 6942. [Google Scholar] [CrossRef]
  134. Gai, Z.; Wang, Y.; Ding, Y.; Qian, W.; Qiu, C.; Xie, H.; Sun, L.; Jiang, Z.; Ma, Q.; Wang, L.; et al. Exogenous abscisic acid induces the lipid and flavonoid metabolism of tea plants under drought stress. Sci. Rep. 2020, 10, 12275. [Google Scholar] [CrossRef] [PubMed]
  135. AAwan, S.A.; Khan, I.; Rizwan, M.; Zhang, X.; Brestic, M.; Khan, A.; El-Sheikh, M.A.; Alyemeni, M.N.; Ali, S.; Huang, L. Exogenous abscisic acid and jasmonic acid restrain polyethylene glycol-induced drought by improving the growth and antioxidative enzyme activities in pearl millet. Physiol. Plant. 2021, 172, 809–819. [Google Scholar] [CrossRef]
  136. Alam, M.M.; Hasanuzzaman, M.; Nahar, K.; Fujita, M. Exogenous salicylic acid ameliorates short-term drought stress in mustard (Brassica juncea L.) seedlings by up-regulating the antioxidant defense and glyoxalase system. Aust. J. Crop Sci. 2013, 7, 1053. [Google Scholar]
  137. Abbaspour, J.; Ehsanpour, A. The impact of salicylic acid on some physiological responses of Artemisia aucheri Boiss. under in vitro drought stress. Acta Agric. Slov. 2016, 107, 287–298. [Google Scholar] [CrossRef]
  138. Chavoushi, M.; Najafi, F.; Salimi, A.; Angaji, S.A. Improvement in drought stress tolerance of safflower during vegetative growth by exogenous application of salicylic acid and sodium nitroprusside. Ind. Crops Prod. 2019, 134, 168–176. [Google Scholar] [CrossRef]
  139. Ghassemi-Golezani, K.; Bilasvar, H.M.; Nassab, A.D.M. Improving rapeseed (Brassica napus L.) plant performance by exogenous salicylic acid and putrescine under gradual water deficit. Acta Physiol. Plant. 2019, 41, 1–8. [Google Scholar] [CrossRef]
  140. Abdelaal, K.A.A.; Attia, K.A.; Alamery, S.F.; El-Afry, M.M.; Ghazy, A.I.; Tantawy, D.S.; Al-Doss, A.A.; El-Shawy, E.-S.E.; Abu-Elsaoud, A.M.; Hafez, Y.M. Exogenous application of proline and salicylic acid can mitigate the injurious impacts of drought stress on barley plants associated with physiological and histological characters. Sustainability 2020, 12, 1736. [Google Scholar] [CrossRef]
  141. Alam, M.; Hayat, K.; Ullah, I.; Sajid, M.; Ahmad, M.; Basit, A.; Ahmad, I.; Muhammad, A.; Akbar, S.; Hussain, Z. Improving okra (Abelmoschus esculentus L.) growth and yield by mitigating drought through exogenous application of salicylic acid. FRES Environ. Bulle 2020, 29, 529–535. [Google Scholar]
  142. Hosain, M.T.; Kamrunnahar, M.; Rahman, M.; Hossain, M.; Munshi, M.; Rahman, S. Drought stress response of rice yield (Oryza sativa L.) and role of exogenous salicylic acid. Int. J. Biosci. 2020, 16, 222–230. [Google Scholar]
  143. Shemi, R.; Wang, R.; Gheith, E.-S.M.S.; Hussain, H.A.; Cholidah, L.; Zhang, K.; Zhang, S.; Wang, L. Role of exogenous-applied salicylic acid, zinc and glycine betaine to improve drought-tolerance in wheat during reproductive growth stages. BMC Plant Biol. 2021, 21, 1–15. [Google Scholar] [CrossRef] [PubMed]
  144. Anjum, S.A.; Tanveer, M.; Hussain, S.; Tung, S.A.; Samad, R.A.; Wang, L.; Khan, I.; Rehman, N.U.; Shah, A.N.; Shahzad, B. Exogenously applied methyl jasmonate improves the drought tolerance in wheat imposed at early and late developmental stages. Acta Physiol. Plant. 2016, 38, 1–11. [Google Scholar] [CrossRef]
  145. Mohamed, H.I.; Latif, H.H. Improvement of drought tolerance of soybean plants by using methyl jasmonate. Physiol. Mol. Biol. Plants 2017, 23, 545–556. [Google Scholar] [CrossRef]
  146. Fugate, K.K.; Lafta, A.M.; Eide, J.D.; Li, G.; Lulai, E.C.; Olson, L.L.; Deckard, E.L.; Khan, M.F.R.; Finger, F.L. Methyl jasmonate alleviates drought stress in young sugar beet (Beta vulgaris L.) plants. J. Agron. Crop. Sci. 2018, 204, 566–576. [Google Scholar] [CrossRef]
  147. Karamian, R.; Ghasemlou, F.; Amiri, H. Physiological evaluation of drought stress tolerance and recovery in Verbascum sinuatum plants treated with methyl jasmonate, salicylic acid and titanium dioxide nanoparticles. Plant Biosyst.-Int. J. Deal. All Asp. Plant Biol. 2020, 154, 277–287. [Google Scholar] [CrossRef]
  148. Yosefi, A.; Mozafari, A.A.; Javadi, T. Jasmonic acid improved in vitro strawberry (Fragaria × ananassa Duch.) resistance to PEG-induced water stress. Plant Cell Tissue Organ Cult. 2020, 142, 549–558. [Google Scholar] [CrossRef]
  149. Shirani Bidabadi, S.; Sharifi, P. Strigolactone and methyl jasmonate-induced antioxidant defense and the composition alterations of different active compounds in Dracocephalum kotschyi Boiss under drought stress. J. Plant Growth Regul. 2021, 40, 878–889. [Google Scholar] [CrossRef]
  150. Abeed, A.H.; Eissa, M.A.; Abdel-Wahab, D.A. Effect of exogenously applied jasmonic acid and kinetin on drought tolerance of wheat cultivars based on morpho-physiological evaluation. J. Soil. Sci. Plant Nutr. 2021, 21, 131–144. [Google Scholar] [CrossRef]
  151. Ahmad Lone, W.; Majeed, N.; Yaqoob, U.; John, R. Exogenous brassinosteroid and jasmonic acid improve drought tolerance in Brassica rapa L. genotypes by modulating osmolytes, antioxidants and photosynthetic system. Plant Cell Rep. 2022, 41, 603–617. [Google Scholar] [CrossRef]
  152. Meng, Y.; Liao, P.; Chen, Y.; Weng, W.; Chen, L.; Xu, F.; Hu, Q.; Xing, Z.; Wei, H.; Gao, H.; et al. Exogenous Application of Methyl Jasmonate Promotes Yield and Grain Quality of Rice under Terminal Drought Stress. Agronomy 2023, 13, 1903. [Google Scholar] [CrossRef]
  153. Naseem, H.; Bano, A. Role of plant growth-promoting rhizobacteria and their exopolysaccharide in drought tolerance of maize. J. Plant Interact. 2014, 9, 689–701. [Google Scholar] [CrossRef]
  154. Gusain, Y.S.; Singh, U.S.; Sharma, A.K. Bacterial mediated amelioration of drought stress in drought tolerant and susceptible cultivars of rice (Oryza sativa L.). Afr. J. Biotechnol. 2015, 14, 764–773. [Google Scholar] [CrossRef]
  155. Gontia-Mishra, I.; Sapre, S.; Sharma, A.; Tiwari, S. Amelioration of drought tolerance in wheat by the interaction of plant growth-promoting rhizobacteria. Plant Biol. 2016, 18, 992–1000. [Google Scholar] [CrossRef]
  156. Niu, X.; Song, L.; Xiao, Y.; Ge, W. Drought-tolerant plant growth-promoting rhizobacteria associated with foxtail millet in a semi-arid agroecosystem and their potential in alleviating drought stress. Front. Microbiol. 2018, 8, 2580. [Google Scholar] [CrossRef]
  157. Tiepo, A.N.; Hertel, M.F.; Rocha, S.S.; Calzavara, A.K.; De Oliveira, A.L.M.; Pimenta, J.A.; Oliveira, H.C.; Bianchini, E.; Stolf-Moreira, R. Enhanced drought tolerance in seedlings of Neotropical tree species inoculated with plant growth-promoting bacteria. Plant Physiol. Biochem. 2018, 130, 277–288. [Google Scholar] [CrossRef]
  158. Zhang, M.; Yang, L.; Hao, R.; Bai, X.; Wang, Y.; Yu, X. Drought-tolerant plant growth-promoting rhizobacteria isolated from jujube (Ziziphus jujuba) and their potential to enhance drought tolerance. Plant Soil. 2020, 452, 423–440. [Google Scholar] [CrossRef]
  159. He, A.; Niu, S.; Yang, D.; Ren, W.; Zhao, L.; Sun, Y.; Meng, L.; Zhao, Q.; Paré, P.W.; Zhang, J. Two PGPR strains from the rhizosphere of Haloxylon ammodendron promoted growth and enhanced drought tolerance of ryegrass. Plant Physiol. Biochem. 2021, 161, 74–85. [Google Scholar] [CrossRef]
  160. Rashid, U.; Yasmin, H.; Hassan, M.N.; Naz, R.; Nosheen, A.; Sajjad, M.; Ilyas, N.; Keyani, R.; Jabeen, Z.; Mumtaz, S.; et al. Drought-tolerant Bacillus megaterium isolated from semi-arid conditions induces systemic tolerance of wheat under drought conditions. Plant Cell Rep. 2021, 41, 549–569. [Google Scholar] [CrossRef]
  161. Ferioun, M.; Srhiouar, N.; Tirry, N.; Belahcen, D.; Siang, T.C.; Louahlia, S.; El Ghachtouli, N. Optimized drought tolerance in barley (Hordeum vulgare L.) using plant growth-promoting rhizobacteria (PGPR). Biocatal. Agric. Biotechnol. 2023, 50, 102691. [Google Scholar] [CrossRef]
  162. Rigi, F.; Saberi, M.; Ebrahimi, M. Improved drought tolerance in Festuca ovina L. using plant growth promoting bacteria. J. Arid. Land. 2023, 15, 740–755. [Google Scholar] [CrossRef]
  163. Waterland, N.L.; Campbell, C.A.; Finer, J.J.; Jones, M.L. Abscisic acid application enhances drought stress tolerance in bedding plants. HortScience 2010, 45, 409–413. [Google Scholar] [CrossRef]
  164. Cochran, D.R.; Harkess, R.L.; Knight, P.R.; Tomaso-Peterson, M.; Blythe, E.K.; Gilliam, C.H. Evaluation of a Commercial Extract of Giant Knotweed on Drought Tolerance of Impatiens. HortScience 2014, 49, 1034–1040. [Google Scholar] [CrossRef]
  165. Park, S.; Mills, S.A.; Moon, Y.; Waterland, N.L. Evaluation of antitranspirants for enhancing temporary water stress tolerance in bedding plants. HortTechnology 2016, 26, 444–452. [Google Scholar] [CrossRef]
  166. Waterland, N.L.; Finer, J.J.; Jones, M.L. Benzyladenine and gibberellic acid application prevents abscisic acid-induced leaf chlorosis in pansy and viola. HortScience 2010, 45, 925–933. [Google Scholar] [CrossRef]
  167. Nordstedt, N.P.; Chapin, L.J.; Taylor, C.G.; Jones, M.L. Identification of Pseudomonas spp. that increase ornamental crop quality during abiotic stress. Front. Plant Sci. 2020, 10, 1754. [Google Scholar] [CrossRef] [PubMed]
  168. Park, S.; Waterland, N.L. Evaluation of calcium application methods on delaying plant wilting under water deficit in bedding plants. Agronomy 2021, 11, 1383. [Google Scholar] [CrossRef]
  169. Đurić, M.; Trifunović-Momčilov, M.; Milošević, S.; Marković, M.; Radulović, O.; Subotić, A.; Uzelac, B. Does Sodium Nitroprusside Alleviate Water Deficit Stress in Impatiens walleriana Shoots Grown In Vitro? Agriculture 2023, 13, 1903. [Google Scholar] [CrossRef]
Figure 1. Drought effects on plant growth and development (left side), and plant resistance mechanisms to drought (right side).
Figure 1. Drought effects on plant growth and development (left side), and plant resistance mechanisms to drought (right side).
Horticulturae 10 00903 g001
Figure 2. I. walleriana with different color of flowers.
Figure 2. I. walleriana with different color of flowers.
Horticulturae 10 00903 g002
Figure 3. Morphological differences between well-watered and drought-stressed I. walleriana. (a,b) well-watered shoots and roots; (c,d) drought-stressed shoots and roots.
Figure 3. Morphological differences between well-watered and drought-stressed I. walleriana. (a,b) well-watered shoots and roots; (c,d) drought-stressed shoots and roots.
Horticulturae 10 00903 g003
Figure 4. 3D structures of I. walleriana aquaporins (IwPIP1;4, IwPIP2;2, IwPIP2;7 and IwTIP4;1), and dehydrins (IwDhn1, IwDhn2.1 and IwDhn2.2), obtained by using the software SWISS-MODEL (https://swissmodel.expasy.org/) and PHYRE2 (http://www.sbg.bio.ic.ac.uk/phyre2/html/page.cgi?id=index).
Figure 4. 3D structures of I. walleriana aquaporins (IwPIP1;4, IwPIP2;2, IwPIP2;7 and IwTIP4;1), and dehydrins (IwDhn1, IwDhn2.1 and IwDhn2.2), obtained by using the software SWISS-MODEL (https://swissmodel.expasy.org/) and PHYRE2 (http://www.sbg.bio.ic.ac.uk/phyre2/html/page.cgi?id=index).
Horticulturae 10 00903 g004
Figure 5. Summarized effects of exogenously applied elicitors on I. walleriana drought-tolerance improvement.
Figure 5. Summarized effects of exogenously applied elicitors on I. walleriana drought-tolerance improvement.
Horticulturae 10 00903 g005
Table 1. The effects of PGRs and PGPBs on drought-tolerance improvement in different plant species. ABA—abscisic acid, ASA—ascorbate, Chl—chlorophyll, DW—dry weight, FW—fresh weight, GSH—glutathione, H2O2—hydrogen peroxide, JA—jasmonic acid, MDA—malondialdehyde, MeJA—methyl jasmonate, O2•−—superoxide anion radical, PGPB—Plant Growth-Promoting Bacteria, PGRs—Plant Growth Regulators, RWC—Relative Water Content, WUE—Water Use Efficiency.
Table 1. The effects of PGRs and PGPBs on drought-tolerance improvement in different plant species. ABA—abscisic acid, ASA—ascorbate, Chl—chlorophyll, DW—dry weight, FW—fresh weight, GSH—glutathione, H2O2—hydrogen peroxide, JA—jasmonic acid, MDA—malondialdehyde, MeJA—methyl jasmonate, O2•−—superoxide anion radical, PGPB—Plant Growth-Promoting Bacteria, PGRs—Plant Growth Regulators, RWC—Relative Water Content, WUE—Water Use Efficiency.
ElicitorPlant SpeciesConcentration/
Bacterial Strains
Way of ApplicationEffectsReferences
ABATriticum aestivum10 µMSoil drenchingIncreased ABA concentration, reduced stomatal conductance, decreased oxidative stress damages, increased antioxidant enzyme activities[126]
Ulmus minor50 and 100 μMFoliar applicationIncreased DW, reduced water loss, enhancement of antioxidant capacity[127]
Camellia sinensis50 mg L−1Foliar applicationLeaf proteome changes[128]
Vicia faba0.1 mMFoliar applicationImproved morphological and anatomical characters, Chl concentrations, and yield[129]
Triticum aestivum10 μMSupplemented in Hoagland solutionIncreased shoot length, and shoot and root dry weights; decreased H2O2 and MDA, increased content of GSH and ASA[130]
Agrostis stolonifera5 μMFoliar applicationLower electrolyte leakage, greater RWC, accumulation of organic acids[131]
Lolium perenne0.054 kg ai ha−1Foliar applicationEnhanced RWC, decreased electrolyte leakage and H2O2, increased soluble sugar content and antioxidant enzyme activity[132]
Axonopus compressus100 μMFoliar applicationReduced oxidative stress, increased pigment content, osmolyte accumulation[133]
Camellia sinensis50 mg L−1Foliar applicationMetabolic changes, increased phenolic content[134]
Pennisetum glaucum100 μMSupplemented in Hoagland solutionImproved Chl and RWC, increased activities of antioxidative enzymes[135]
SABrassica juncea50 μMFoliar applicationIncreased RWC, Chl content, antioxidant enzyme activity, decreased H2O2 and lipid peroxidation level[136]
Vicia faba1 mMFoliar applicationImproved morphological and anatomical characters, Chl concentrations, and yield[129]
Artemisia aucheri0.01 and 0.1 mMIn vitroImproved DW and FW, Chl and carotenoid contents, increased soluble carbohydrates, increased biosynthesis of phenolic compounds[137]
Agrostis stolonifera10 μMFoliar applicationLower electrolyte leakage, greater RWC, accumulation of amino acids and carbohydrates[131]
Carthamus tinctorius250 μMFoliar applicationActivation of non-enzymatic antioxidant defense system, increased proline content, decreased oxidative stress[138]
Brassica napus1 mMFoliar applicationIncreased antioxidant enzyme activities, water content, membrane integrity and Chl, improvement in grain and oil yields[139]
Hordeum vulgare0.5 mMFoliar applicationIncreased stem length, DW, Chl, RWC, activity of antioxidant enzymes, and grain yield; decreased lipid peroxidation, electrolyte leakage, O2·− and H2O2[140]
Abelmoschus esculentus80, 160 and 240 mgL−lFoliar applicationSA at 240 mgL−1 improved the best all of the growth and yield attributes; had minimum days to flowering and picking, and maximum single-pod weight, average pod length, plant height, number of leaves per plant, number of pods per plant, stem diameter and yield[141]
Oryza sativa250, 500, 750 and 1000 µM m−2Foliar applicationImproved grain yield and harvest index when 750 µM m−2 SA applied[142]
Triticum aestivum140 mg L− 1Foliar applicationIncreased grain yield, decreased MDA, H2O2 and O2•−, increased proline, soluble sugars and antioxidant enzyme activity[143]
JasmonatesTriticum aestivum0.5 mM MeJAFoliar applicationImproved dry biomass, number of grains per spike, and grain weight and yield[144]
Glycine max20 µM MeJAFoliar applicationIncreased growth parameters, RWC, photosynthetic pigments, cell wall components, unsaturated fatty acids, and phenolic compounds[145]
Beta vulgaris0.01, 0.1, 1 or 10 μM MeJAFoliar application1 and 10 μM MeJA reduced moderate- and severe-drought effects on RWC, photosynthesis rate, substomatal CO2 concentration and WUE, and altered drought-induced changes in proline accumulation[146]
Verbascum sinuatum200 µM MeJAIn vitroMeJA negatively affected growth parameters and increased the content of MDA, H2O2, total saponin and activity of peroxidase and polyphenol oxidase[147]
Fragaria ×
ananassa
0.01 and
0.05 mM JA
In vitroImproved growth, RWC, and pigment content[148]
Dracocephalum kotschyi0.5 mM MeJAFoliar applicationHigher FW and DW, lower electrolyte leakage, MDA, H2O2, total phenol content, total antioxidant activity and antioxidant power assay[149]
Triticum sativum0.1 mM JAFoliar applicationImproved growth, restoration of shoot/root ratio, accumulation of osmolytes, regulated activity of antioxidant enzymes[150]
Pennisetum glaucum100 μMSupplemented in Hoagland solutionImproved Chl and RWC, increased activities of antioxidative enzymes[135]
Brassica rapa10 µM JAFoliar applicationImproved photosynthetic rate, photosynthetic pigments, stomatal conductance, transpiration rate and antioxidant defence, increased osmolyte accumulation, and decreased membrane damage[151]
Oryza sativa100 μmol L−1 MeJAFoliar applicationIncreased grain yield and quality[152]
PGPBsZea maysProteus penneri,
Pseudomonas aeruginosa, and
Alcaligenes faecalis
Seed primingImproved plant biomass, root and shoot length, leaf area, RWC, protein and sugar content[153]
Oryza sativaPseudomonas sp., Bacillus cereus, Arthrobacter nitroguajacolicusSoil inoculationEnhanced growth, higher proline content and antioxidant enzyme activities, lower MDA and H2O2[154]
Triticum aestivumKlebsiella sp., Enterobacter ludwigii, Flavobacterium sp.Seed priming, soil inoculationTranscriptomic changes, improved root length and number, shoot DW, root FW and DW, and physiological and biochemical parameters[155]
Setaria italicaPseudomonas fluorescens,
P. migulae, Enterobacter hormaechei
Seed priming, soil inoculationStimulated seed germination and seedling growth[156]
Trema micrantha, Cariniana estrellensisAzospirillum brasilense, Bacillus sp., Azomonas sp., Azorhizophillus sp.Seed priming, soil inoculationIncreased drought tolerance, growth parameters, and physiological and biochemical attributes[157]
Ziziphus jujubaPseudomonas,
Bacillus, Serratia
Soil inoculationPseudomonas lini and Serratia plymuthica increased plant height, shoot and root dry matter, RWC, and antioxidant enzyme activities; decreased MDA and ABA[158]
Thymus, Sarcocornia, MenthaPseudomonas, Pantoea, AcinetobacterSoil inoculationin vitro PGP-associated traits, including phosphate solubilization, indole-3-acetic acid production, and 1-aminocyclopropane-1-carboxylate deaminase activity; increased tolerance to salinity and drought[122]
Petunia, PelargoniumPseudomonas, Arthrobacter, Herbaspirillium,Soil inoculationIncreased biomass; increased number of flowers; mitigation of the reduction in photosynthetic parameters of water-stressed P. hybrida and Pelargonium × hortorum[123]
Haloxylon ammodendronBacillus sp.,
Pseudomonas sp.
Seed priming, soil inoculationImproved root system and growth, RWC, photosynthetic capacity, antioxidant enzyme activities and regulated ABA content[159]
Triticum aestivumBacillus megaterium,
B. licheniformis
Seed priming, soil inoculationIncreased germination index, promptness index, seedling vigor index, FW and DW, RWC, photosynthetic pigments, osmolytes, and activities of defense-related antioxidant enzymes[160]
Hordeum vulgarefour bacterial isolates (MFC1, MFE3, MFF2, and MFF5)Soil inoculationIncreased shoot dry weight, RWC, Chl content, photosynthesis efficiency, and proline content; decreased MDA and H2O2[161]
Festuca ovinaAzotobacter vinelandii,
Pantoea agglomeran, Pseudomonas putida
Seed primingImproved seed germination, plant growth, and nutrient uptake[162]
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

Milovančević, M.; Trifunović-Momčilov, M.; Radulović, O.; Milošević, S.; Subotić, A. Drought Stress Effects and Ways for Improving Drought Tolerance in Impatiens walleriana Hook.f.—A Review. Horticulturae 2024, 10, 903. https://doi.org/10.3390/horticulturae10090903

AMA Style

Milovančević M, Trifunović-Momčilov M, Radulović O, Milošević S, Subotić A. Drought Stress Effects and Ways for Improving Drought Tolerance in Impatiens walleriana Hook.f.—A Review. Horticulturae. 2024; 10(9):903. https://doi.org/10.3390/horticulturae10090903

Chicago/Turabian Style

Milovančević, Marija, Milana Trifunović-Momčilov, Olga Radulović, Snežana Milošević, and Angelina Subotić. 2024. "Drought Stress Effects and Ways for Improving Drought Tolerance in Impatiens walleriana Hook.f.—A Review" Horticulturae 10, no. 9: 903. https://doi.org/10.3390/horticulturae10090903

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