Next Article in Journal
Role of Aiolos and Ikaros in the Antitumor and Immunomodulatory Activity of IMiDs in Multiple Myeloma: Better to Lose Than to Find Them
Next Article in Special Issue
Transcriptomics Reveals Fast Changes in Salicylate and Jasmonate Signaling Pathways in Shoots of Carbonate-Tolerant Arabidopsis thaliana under Bicarbonate Exposure
Previous Article in Journal
Nanotechnology Therapy for Alzheimer′s Disease Memory Impairment Attenuation
Previous Article in Special Issue
Whole-Transcriptome RNA Sequencing Reveals the Global Molecular Responses and CeRNA Regulatory Network of mRNAs, lncRNAs, miRNAs and circRNAs in Response to Salt Stress in Sugar Beet (Beta vulgaris)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plant Morphological, Physiological and Anatomical Adaption to Flooding Stress and the Underlying Molecular Mechanisms

Chongqing Institute of Green and Intelligent Technology, Chinese Academy of Sciences, Chongqing 400714, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(3), 1088; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22031088
Submission received: 18 December 2020 / Revised: 17 January 2021 / Accepted: 19 January 2021 / Published: 22 January 2021
(This article belongs to the Special Issue Environmental Stress and Plants)

Abstract

:
Globally, flooding is a major threat causing substantial yield decline of cereal crops, and is expected to be even more serious in many parts of the world due to climatic anomaly in the future. Understanding the mechanisms of plants coping with unanticipated flooding will be crucial for developing new flooding-tolerance crop varieties. Here we describe survival strategies of plants adaptation to flooding stress at the morphological, physiological and anatomical scale systemically, such as the formation of adventitious roots (ARs), aerenchyma and radial O2 loss (ROL) barriers. Then molecular mechanisms underlying the adaptive strategies are summarized, and more than thirty identified functional genes or proteins associated with flooding-tolerance are searched out and expounded. Moreover, we elaborated the regulatory roles of phytohormones in plant against flooding stress, especially ethylene and its relevant transcription factors from the group VII Ethylene Response Factor (ERF-VII) family. ERF-VIIs of main crops and several reported ERF-VIIs involving plant tolerance to flooding stress were collected and analyzed according to sequence similarity, which can provide references for screening flooding-tolerant genes more precisely. Finally, the potential research directions in the future were summarized and discussed. Through this review, we aim to provide references for the studies of plant acclimation to flooding stress and breeding new flooding-resistant crops in the future.

1. Introduction

Along with climate changes, flooding shows a heavier tendency triggering severe crop reduction in both yield and quality around the world [1]. In the United States, for instance, losses of crop yield induced by flooding followed by drought from 2000 to 2011 [2]. In 2016, a total of 217 million dollars was paid for damages or controls associated with flooding, which cost 3.4 times higher than that for droughts (www.rma.usda.gov/data/cause). In some developing countries, because of their poor drainage systems, the impact of frequent flooding on crop systems is more serious, which usually aggravates poverty and food insecurity [1]. To reduce the losses, it is essential to unravel the mechanisms of plants against flooding stress for developing new flooding-tolerance crops.
Generally, flooding can be classified into two forms depending on water depth: waterlogging and submergence [3,4]. Waterlogging is the condition that water exists on the soil surface and only plant roots are surrounded by water, while submergence is the state that the whole plant partially or completely immerses in water [3,5,6]. One of the immediate impacts caused by flooding is the deficiency of oxygen [7,8]. Previous studies showed that oxygen diffused in water approximately 10,000 times slower than that in air, and oxygen permeating into water-immersed soil is about 320,000 times less than that into soil full of gas [3,9,10]. The deficiency of oxygen in soil restricts plant growth resulting in the reduction of crop yield [3,11]. Under waterlogging stress, the growth of the plant was impeded due to enhanced anaerobic respiration [3,11]. Meanwhile, adventitious roots (ARs), aerenchyma and radial O2 loss (ROL) barriers emerged in roots for the exchange of gas [3,11]. When exposed to submergence stress, plants have evolved two main strategies to resist the adversity: low oxygen quiescence syndrome (LOQS) for complete submergence and low oxygen escape syndrome (LOES) for partial submergence [8,12]. Plants with LOQS strategy demonstrate restricted growth through keeping the minimum energy and carbon consumptions for prolonging underwater survival [13] (Figure 1A). After the flooding recedes, these plants can be recovered rapidly [11]. Escaping from water (LOES) is another strategy that plants maintain the upper leaves in the aerial environment for getting adequate oxygen, light and carbon dioxide [8,11]. Accordingly, several morphology and anatomical traits of plants change with the water depth, such as deepwater rice and Rumex palustris with internode or petiole elongation as well as aerenchyma formation [14,15,16] (Figure 1A).
Recently, much progress has been made in the mechanisms of plant acclimation to flooding stress [17,18,19]. For example, Nagai et al. reported two genes controlled internode elongation antagonistically in rice, the ‘accelerator’ ACCELERATOR OF INTERNODE ELONGATION 1 (ACE1) and the ‘decelerator’ DECELERATOR OF INTERNODE ELONGATION 1 (DEC1) [18]. Tang et al. showed that two transcription factors of the WRKY family (WRKY33 and WRKY12) interacted with each other to up-regulate RAP2.2 for Arabidopsis adaptation to submergence stress [19]. Bui et al. reported that transcription factor ANAC017 mediated differential submergence tolerance of Arabidopsis between juvenile and adult stages [20]. These studies contribute to our understanding of the mechanisms of plant adaptation to flooding stress, however, the detailed update including these latest research progresses has not been conducted subsequently. In this review, plant survival strategies under flooding stress in morphology, physiology and anatomy levels were investigated and exhibited initially, then the potential crucial genes or proteins and the regulatory role of phytohormones were summarized by highlighting the most recent findings of plant adapting to flooding stress. Moreover, several key open questions worthy of further study were proposed and discussed.

2. Plant Responses to Flooding Stress

Flooding is a natural occurrence with adverse effects on plant growth and development, multiple mechanisms of plants adaption to flooding stress have been discovered at morphology, physiological and anatomical scales.

2.1. The Morphology of Plants under Flooding Stress

Generally, flooding stress triggers adaptive changes in plant roots and shoots at a morphological scale, such as the formation of ARs, the growth of the shoots inhibited or accelerated (Figure 1A). Previously, some important crops including wheat, maize, rice and soybean have been studied in their responses to flooding stress. For instance, wheat, as a dryland crop, was sensitive to flooding stress [21]. When exposed to waterlogging stress, dry mass of wheat shoot and root, and the ratio of root/shoot significantly declined compared to controls, indicating that root growth was inhibited more seriously than shoot [22,23]. Maize was also intolerant against waterlogging stress and the trefoil stage was the most sensitive period for maize [24,25]. Waterlogging inhibited maize growth resulting in decline in plant height, ear height, dry weight, leaf area index and grain characteristics (such as grain number per ear and 1000-grain weight) [26]. Meanwhile, contents of chlorophyll, soluble sugar and starch in leaves, stems and roots decreased under waterlogging stress [27]. For rice, many lowland cultivars are susceptible to submergence stress and thus hardly survive as the deluge lasts for a longer period of time [3,28,29]. However, several rice cultivars can experience flash-flood (complete submergence) for around two weeks by restricting shoot elongation as well as carbohydrate consumption, and will be recovered once the flash-flood recedes [5,11]. On the other hand, deepwater rice, which adapts to the submergence where it grows, is able to maintain the top leaves in the aerial environment for catching sufficient oxygen by elongating its internodes rapidly [5,30,31]. The growth and grain yield of soybean is also affected by flooding [32]. At the seedling stage, root growth of soybean was suppressed severely after submergence prolonging for 10 days [33]. Different soybean genotypes might utilize distinct mechanisms to resist waterlogging stress and proteins associated with energy metabolism were thought to function in soybeans tolerance to flooding stress [34,35]. Generally, waterlogging stress enhances anaerobic respiration of roots of dryland plants along with energy consumption and restricts their growth eventually. Likewise, complete submergence stress inhibits the growth of plants and decreases the rate of survival with the extension of the submergence time in most cases. By contrast, partial submergence promotes the elongation of internode, petiole or other organs in some plant species for getting enough oxygen, which contributes to their rapid growth in a short time.
Adventitious roots, as a typical trait of plants acclimating to waterlogging stress, can facilitate the uptake of water and nutrient as well as the transport of gas more effectively [36] (Table 1). ARs can be developed from some non-root tissues of the plants during normal development, or under the stimulation of external stresses, including flooding, nutrient deficiency, heavy metals and wounding [37]. On the basis of physical characteristics and generation conditions, ARs can be divided into various types, such as hypocotyl roots, crown roots, brace roots, nodal roots, stem roots, junction roots and prop roots, among which the front five types can be triggered by flooding [37]. When exposed to waterlogging stress, maize formed more crown roots than controls and the waterlogging-tolerant line possessed more crown roots than the waterlogging-sensitive line [38]. In deep-water rice, the elongation of ARs was initiated after 8 to 10 h under partial submergence stress, and ethylene was found to facilitate the formation of ARs through triggering the death of epidermal cells covering the root tip [39,40]. In tomato, ARs gradually formed at the hypocotyl after three days of partial submergence, and further extended to the upper surface of the water along caulicles [41]. Furthermore, inhibiting ethylene or auxin production impaired ARs formation of tomato, which indicated that ethylene and auxin might interact with each other during the development of ARs under flooding stress [42]. Likewise, a recent study reported that waterlogging stress-induced cucumber generating a large amount of ARs, and both auxin and ethylene can stimulate the formation of ARs, however, auxin-induced ARs relied on ethylene while ethylene-induced ARs was unaffected by auxin [43]. Contrastively, the opposite result was found in Rumex palustris that auxin-induced ARs were not affected by inhibiting ethylene biosynthesis while the inhibition of auxin transport suppressed ethylene-induced ARs formation [44]. In summary, the formation of plant ARs is mainly regulated by ethylene and auxin under flooding stress, however, their roles may be diverse in different plants.

2.2. The Physiological Phenotype of Plants Response to Flooding Stress

Flooding stress leads to accumulation of the harmful substances in plants, such as reactive oxygen species (ROS) and malondialdehyde (MDA) [67]. In turn, plants activate the internal protection mechanisms to scavenge these components, such as the production of proline, glutathione and the activation of antioxidant enzymes [57,68,69] (Figure 1B, Table 2). ROS are toxic byproducts of oxidative metabolisms, such as O2.−, H2O2, OH˙ and 1O2 [70,71]. Whereas, ROS also functions as signal molecules that regulate plant development under environmental stresses [56,71]. Submergence exhibited a huge impact on ROS homeostasis. Ye et al. reported that submergence led to a decrease of H2O2 content in bermudagrass (Cynodon dactylon) leaves, though O2.− content had insignificant change [68]. Contrastively, another study showed that O2.− and H2O2 concentration in the elongation and mature zone of barley roots increased under waterlogging stress [72]. Additionally, MDA content in bermudagrass leaves increased after submergence for three weeks and catalase (CAT), glutathione reductase (GR) and peroxidase (POD) activities presented an upward trend with the extension of the time [68]. Wang et al. showed the activities of antioxidative enzymes POD and superoxide dismutase (SOD) in roots of Triarrhena sacchariflora increased first and then decreased under flooding stress as well as contents of proline and MDA compared with control plants [69].
On the other hand, flooding is accompanied by the deprivation of oxygen, which leads to aerobic respiration in a low-level and accelerates more carbohydrates consumed by glycolysis [85]. Moreover, under complete submergence stress, the insufficiency of carbon dioxide and light decreased plant photosynthesis rates [13,86] (Figure 1B). For instance, the accumulation of soluble sugar and sucrose in bermudagrass leaves declined obviously under submergence stress [68]. Loreti et al. reported that starch content in rosette leaves of Arabidopsis degraded with the extension of submerged time during the night, and glucose content was observed decline at the end of the night as well as sucrose [87]. For spring maize, the negative impact induced by waterlogging stress was different because of the distinction of duration and growth stage, and the seedling stage was the sensitive period followed by the jointing and tasseling stages [88]. Additionally, the photosynthetic rate (Pn) decreased with the waterlogging time prolonging and resulted in the reduction of total dry weight and grain yield ultimately [88].

2.3. Anatomical Adaptations to Flooding Stress

Flooding often leads to oxygen deprivation in soils and plant roots, which significantly restricts metabolic activities for plant survival and development. Therefore, amounts of aerenchyma and ROL barriers in plant roots are formed to transport adequate oxygen from shoots to roots (Figure 1C).

2.3.1. The Formation of Aerenchyma

Aerenchyma not only exists in many plants constitutively, but also is formed inductively by environmental stress, such as flooding stress [89] (Table 1). It forms in roots and shoots and consists of thin-walled cells filled with gas spaces to facilitate the diffusion of oxygen [8]. For instance, Aerenchyma formed in the mid-cortex cells of wheat roots after waterlogging for 24 h and its formation was mediated by ROS [55,56]. When applied with an ethylene precursor beforehand, wheat seedlings formed lysigenous aerenchyma in roots and these aerenchyma were further developed under hypoxic stress [90]. Zea nicaraguensis is a wild relative of maize and grew in frequently flooded areas [49]. Soil waterlogging increased aerenchyma proportion in the cross-sectional area of maize root from less than 1% in drained soil to 15% as well as from 22% to 29% in Z. nicaraguensis roots [49]. Another study reported that ARs of waterlogging-tolerant barley (Hordeum vulgare L.) genotypes exhibited remarkably higher porosity and formed aerenchyma more quickly compared to sensitive ones [57]. Aerenchyma increases plant tolerance to flooding stress by improving the oxygen-transport efficiency from shoot to root tip, and its formation mainly relies on ethylene and ROS signaling.

2.3.2. A Barrier to Radial Oxygen Loss

Except for aerenchyma, the oxygen-transport efficiency from shoot to root tip depends on another crucial factor, the formation of ROL barriers. ROL barriers usually form in the roots of waterlogging-tolerant plants, and can reduce oxygen leakage during its transport from shoot-to-root tip and impede soil phytotoxin entry simultaneously [49,51]. The ROL barriers can be formed constitutively or induced by stagnant conditions, such as flooding stress [14,91] (Figure 1C, Table 1). Many wetland species, such as Oryza sativa, Juncus effusus, Phragmites australis and Glyceria maxima, have been reported forming ROL barriers in roots [14,15,92,93]. For example, growing in stagnant nutrient solution with oxygen removal enhanced the development of ROL barriers in ARs of rice [15,45]. Long ARs of rice initiated the formation of ROL barriers quickly and developed well within 24 h under hypoxic conditions, however, short roots generated the barrier for more than 48 h [93]. The dryland species have the ability to form ROL barriers under flooding stress. For instance, seminal roots of wheat induced ROL barriers when exposed to waterlogging to enhance the internal supply of oxygen [56]. Notably, some Echinochloa species even developed a ROL barrier constitutively under aerated conditions [66].
The formation of ROL barriers can also be triggered by the organic acids [94,95]. In waterlogged soils, some monocarboxylic acids are produced by anaerobic microorganisms, such as acetic, butyric, hexanoic and propionic acids, which inhibit plant growth as accumulated to some extent and can invoke enhanced suberization or lignification in cell walls of roots [96,97]. For example, the ROL barrier in Hordeum marinum roots was induced by butyric and hexanoic acids individually or a mixture with acetic and propionic acids [94], and these four organic acids also triggered the ROL barrier in roots of rice [95].

3. Molecular Mechanism of Plant Response to Flooding Stress

On the basis of the morphology, physiology and anatomy characteristics of plant response to flooding stress, many associated-genes functioning in plant adaption to flooding stress have been reported and studied (Table 3).

3.1. The Formation of Adventitious Roots

The emergence of ARs is one of the adaptive strategies for plants coping with flooding stress, and several genes have been identified to mediate ARs formation (Figure 2). GNOM encodes a guanine-nucleotide exchange factor for ADP-ribosylation factor (ARF) [120] and can influence polar auxin transport in plants [121,122]. Studies have reported that OsGNOM1 in rice affected the development of ARs through regulating polar auxin transport, and loss function of this gene resulted in the defect of ARs formation [120]. CsARN6.1 in cucumber encoded an AAA ATPase domain-containing protein and affected waterlogging tolerance through the regulation of ARs formation, besides, the transformation of CsARN6.1 from waterlogging-tolerant cultivar into waterlogging-sensitive variety increased numbers of ARs under waterlogging conditions [117]. Ethylene can induce ARs formation and this process requires ROS as a signal triggering programmed cell death in cortex cells [123,124]. For instance, waterlogging led to the accumulation of ethylene and ROS in cucumber, ethylene-induced AR formation was inhibited by the reduction of H2O2 [43]. The above results suggest that auxin, ethylene and ROS are required for the formation of ARs under flooding stress, but the mechanisms of interaction among them need to be revealed.

3.2. The Homeostasis of Reactive Oxygen Species

Flooding stress leads to oxidative stress for plants, which triggers plant antioxidant systems rapidly [68,77]. Respiratory burst oxidase homolog (RBOH) encodes a plasma membrane-associated NADPH oxidase in plants and can generate hydrogen peroxide (H2O2) by converting O2 to O2·− [127,128,129]. Qi et al. reported that cucumber accumulated higher levels of ROS under waterlogged stress, further investigation showed that two members CsRBOHB and CsRBOHF3 were strongly upregulated and might result in ROS accumulation [43]. Hypoxia Responsive Universal Stress Protein 1 (HRU1) can be induced by submergence stress in Arabidopsis, and was regulated by an VII Ethylene Response Factor (ERF-VII) protein RAP2.12 (Related to Apetala 2.12) simultaneously; loss of HRU1 function affected H2O2 production and became sensitive to submergence and anoxia, further research exhibited that HRU1 mediated ROS production in Arabidopsis through an interaction with the GTPase ROP2 and the NADPH oxidase RbohD [130]. The transcriptome of two Brachypodium distachyon ecotypes with contrasting submergence tolerance exhibited that lots of genes associated with redox reaction significantly changed tolerant ecotype, such as genes coding for ascorbate peroxidases (ASP), ascorbate oxidases (ASO) and peroxidases (PER), which might be involved in ROS management under submergence stress [131]. ROS accumulation is harmful for plants, on the other hand, it mediates ARs and aerenchyma formation as signaling molecules. Under flooding stress, how plants keep the balance between oxidative stress and signal transduction needs to be further explored.

3.3. The Mechanism of Aerenchyma Formation

Generally, aerenchyma can be classified into two types: schizogenous aerenchyma and lysigenous aerenchyma [132]. Schizogenous aerenchyma that forms gas spaces owing to cell separation and expansion without cell death, by contrast, lysigenous aerenchyma is formed due to the death and subsequent lysis of some cells [53]. Under flooding stress, plants usually form lysigenous aerenchyma in roots to improve oxygen diffusion (e.g., Arabidopsis, maize and wheat) and several genes involved in aerenchyma formation have been investigated [53,62,90] (Figure 2). In Arabidopsis, lysigenous aerenchyma emerged in hypocotyls under hypoxic stress and this process relied on three genes LSD1, EDS1 and PAD4, which regulated the upstream ethylene signal and the generation of ROS [62]. With laser microdissection technology, Rajhi et al. investigated the gene expression profile of cortical cells in maize root when the lysigenous aerenchyma formed, and showed that the expression of RBOH gene increased sharply under waterlogging stress while MT (metallothionein) gene for scavenging ROS was down-regulated, which may lead to the generation of excess ROS and triggered programmed cell death in cortical cells for aerenchyma formation [53]. Additionally, ethylene-induced the formation of aerenchyma under waterlogged conditions [53]. In wheat roots, an ethylene precursor pre-treatment enhanced the expression of three RBOH genes for ROS generation and activated the formation of lysigenous aerenchyma under stagnant conditions, which might be one of the strategies for wheat seedlings acclimation to hypoxia stress [90]. According to the above researches, it can be concluded that both ethylene and ROS mediate the formation of aerenchyma. Ethylene may induce the generation of ROS and results in the programmed cell death of cortical cells for lysigenous aerenchyma formation eventually.

3.4. The Forming Basis of ROL Barriers

Apart from aerenchyma, plants develop ROL barriers in roots to impede oxygen leakage and several genes have been reported to mediate its formation [36] (Figure 2). Previous studies exhibited that suberin might be the major component of ROL barriers compared to lignin. For example, De Simone et al. showed that suberization but not lignification of exodermal cell walls suppressed the radial oxygen loss effectively in roots of four Amazon tree species, indicating it was suberin forming the ROL barriers under flooding stress [133]. Shiono et al. reported that the expression of most genes associated with suberin biosynthesis increased sharply in the outer part of rice roots during ROL barrier formation, including GLYCEROL-3-PHOSPHATE ACYLTRANSFERASE (GPAT) and PEROXIDASE (POD), which have also been identified to be upregulated during the formation of ROL barriers with Si supply [48,134], by contrast, few genes related to lignin biosynthesis were found in this process [88]. Additionally, some transcription factors containing WRKY, AP2, MYB, and NAC domains might participate in the regulation of suberin biosynthesis during ROL barrier formation [48,135,136,137] (Figure 2). Although suberin is considered to be more crucial in forming the ROL barrier of plant roots, the initiation and regulation mechanisms of suberin deposition remain unclear.

4. Regulatory Mechanisms of Phytohormones in Plants Response to Flooding Stress

Phytohormones play critical roles in regulating plant growth and can integrate multiple signal transduction pathways under environmental stresses [138,139]. Some phytohormones have been identified to be involved in plants adapting to flooding stress, such as ethylene, gibberellin and auxin [13] (Table 3, Figure 2 and Figure 3).

4.1. Ethylene

Ethylene is a phytohormone in the form of gas, and hardly escape from plants under flooding conditions, which led to its rapid accumulation inside the plant [13]. The increased ethylene promotes the formation of AR and aerenchyma in plants when exposed to flooding stress [43,90] (Figure 2). ERF-VIIs are ethylene-responsive transcription factors with a conserved APETALA2 (AP2) domain and have been identified to function in plants tolerance to waterlogging and submergence stresses [17,25,31,104,140] (Table 3). Two ERF-VII members SK1 and SK2 from deepwater rice were transcriptionally induced by submergence stress and contributed to the elongation of internodes so as to keep the top leaves from submerging [11,31]. Another ERF-type transcription factor Sub1A, was inducible expression under submergence stress, and overexpression of Sub1A with a tolerance-specific allele in submergence-intolerant variety enhanced its submergence tolerance [104]. ERF-VII member TaERFVII.1 from waterlogging-tolerant wheat (Triticum aestivum) was triggered expression obviously by waterlogging but not in susceptible cultivar, overexpression of TaERFVII.1 increased the waterlogging-tolerance ability of wheat with no adverse effect on its growth and grain production [17], which can be considered as a candidate gene for breeding waterlogging-resistant crops. Yu et al. reported that an ERF-VII member from maize, ZmEREB180, was concerned with waterlogging tolerance and its expression level was affected by the changes of 5ʹ-untranslated region (5′-UTR); transgenic maize of ZmEREB180 increased the survival rate through promoting ARs formation and improving antioxidant ability, which was further supported by transcriptomic analysis [25]. As transcription factors, ERF-VIIs usually function by affecting the expression of other genes, however, they are regulated by other transcription factors simultaneously. Recently, Tang et al. reported two transcription factors of WRKY33 and WRKY12 were involved in Arabidopsis adaptation to submergence stress [19]. WRKY33 interacted with WRKY12 and both of them participated in hypoxia response partially by activating the expression of ERF-VII member RAP2.2 [19]. In addition, overexpression of WRKY33 or WRKY12 in Arabidopsis enhanced resistance to hypoxia [19].
Considering the importance of ERF-VIIs for plant adaption to flooding stress, their members in the model plant Arabidopsis, main crops (rice, maize, wheat and tomato) and other reported ERF-VIIs were collected and constructed a phylogenetic tree. These members can be divided into three groups according to sequence similarity. ERF-VIIs that had been reported to be involved in plant tolerance to flooding stress belongs to Group Ⅰ and Ⅱ, no genes associated with flooding-tolerance have yet been found in Group Ⅲ (Figure 3), which can provide references for screening flooding-tolerant genes more precisely. In brief, ethylene as a major regulator mediated the formation of ARs, aerenchyma, shoot hyponasty and elongation [141]. However, the downstream genes of ethylene-signaling pathway associated with flooding-tolerance are largely unknown. For example, ERF-VIIs as transcription factors, usually function by regulating the expression of their target genes, but so far, lots of these upstream genes have not been found and studied.

4.2. Auxin

Auxin, which acts as a general coordinator of plant growth and development, has been reported to participate in several plant species adapting to waterlogging stress (Figure 2), such as box elder (Acer negundo) [143], tomato [42], sunflower [144] and cucumber [43]. Auxin functions in the initiation of root apical meristems and increased auxin content may result in the formation of AR primordia [43,145,146]. OsPIN1 is a polar auxin transporter from rice and plays an important role in auxin-dependent ARs formation; the loss function of OsPIN1 suppressed the development of ARs apparently and application of auxin partially rescued the missed phenotypes [125]. Another polar auxin transporter PIN2 of Solanum dulcamara could be induced by flooding stress and ethylene application, the silence of PIN2 or impediment of auxin transport with chemical methods inhibited AR primordium initiation [126]. Gutierrez et al. reported that three genes of AUXIN RESPONSE FACTOR (ARF) family regulated ARs initiation cooperatively in Arabidopsis [147]. ARF6 and ARF8 were positive regulators while ARF17 regulated ARs formation negatively. Meanwhile, three auxin-induced Gretchen Hagen3 (GH3) genes (GH3.3, GH3.5, and GH3.6) for the biosynthesis of acyl-acid-amido were necessary for ARs initiation in Arabidopsis hypocotyl [148]. All of these studies indicated that auxin signal participates in plant acclimation to flooding stress mainly through regulating the formation of ARs.

4.3. Gibberellin

Gibberellin was also identified to function in plant adaptation to flooding stress (Figure 4). As suffered from flash floods, submergence-tolerant plants usually restrict their growth by activation of GA signal [149]. Slender Rice-1 (SLR1) and SLR1 Like-1 (SLRL1) are GA signaling repressors in rice [150,151,152]. Previous research exhibited that Sub1A promoted the accumulation of SLR1 and SLRL1 to suppress GA signal transduction in submergence-tolerant rice under submergence stress, and affected the expression of PDC and ADH genes positively, which led to the inhibited growth of submergence-tolerant rice for energy conservation and regrowth after flooding [104,149,153]. GID1 (GIBBERELLIN INSENSITIVE DWARF1) is a soluble GA receptor and has been reported previously [154]. Under submergence conditions, loss function of GID1 in rice inhibited the degradation of chlorophyll and promoted the metabolism of carbohydrates, further analyses suggested that GID1 regulated GA signals to influence submergence tolerance through controlling carbohydrate consumption [155].
GA also plays an important role in the stem elongation of plants [156]. In lowland rice, an increase of active GA (GA1) resulted in the elongation of leaf sheath under submergence stress [157]. For deepwater rice, submergence induced the expression of gibberellin biosynthesis gene SD1 (SEMIDWARF1) that was regulated by an ethylene-responsive transcription factor OsEIL1a, and internode elongation was promoted by increasing gibberellins [110]. Recently, two genes, an ‘accelerator’ ACE1 and a ‘decelerator’ DEC1, were reported to regulate internode elongation antagonistically together with GA; ACE1, as an unknown function protein, contributed internode elongation in concert with GA while DEC1, as a zinc-finger transcription factor, had the opposite function that restricted internode elongation; these two genes were involved in deepwater rice acclimation to submergence stress collectively [18]. According to the above results, GA mainly functions in plant acclimation to flooding stress through the modulation of carbohydrate metabolism and the regulation of internode (or other organs) elongation. Actually, GA mediates flooding-tolerance of plants together with ethylene in most cases.

4.4. Other Phytohormones

Moreover, cytokinin and abscisic acid (ABA) have been identified to function in plants against flooding stress [114,158,159]. Waterlogging-induced ARs formation in wheat is relevant to four phytohormones (IAA, GA, cytokinin and ABA) and several genes associated with their biosynthesis or metabolism were inducible expression under waterlogging stress [160]. Salicylic acid (SA) also participates in the waterlogging-tolerance of plants. SA content in waterlogging-tolerant soybean lines increased significantly after waterlogging for 5 or 10 days compared to non-waterlogging conditions while SA content in sensitive lines exhibited no significant change, implying that SA mediates waterlogging-tolerance of soybean through regulating the formation of aerenchyma or ARs [161]. Another phytohormone brassinosteroid (BR) has been reported that it may affect GA signaling by interacting with Sub1A, and regulated the elongation of shoot for surviving under submergence condition [162].
In fact, different phytohormones function cooperatively in response to flooding stress. For instance, the ethylene response factor Sub1A in rice repressed the GA signal to prevent shoot elongation and carbohydrate consumption under complete submergence [104,149]. However, when exposed to partial submergence, ethylene accumulation contributed to the expression of SK1 and SK2, which facilitated internode elongation through regulating GA biosynthesis or GA signal transduction [31]. Notably, the expression of SK1 and SK2 may be regulated by one transcription factor EIN3 (ethylene-insensitive-3-like protein) through binding to their promoter regions [31]. To summarize, plant response to flooding stress is a complicated process depending on the co-regulation of multiple phytohormones and the cooperation mechanisms among different hormones need to be further revealed. Meanwhile, the studies about SA and BR regulating flooding-tolerance of plants should be strengthened in the future.

5. Conclusions and Further Perspectives

Recently, the mechanisms of plant adaption to flooding stress have been progressively studied. However, there are still several key open questions to be further clarified.
Firstly, since former studies of plants adaption to flooding stress excessively focus on model or wetland plants including Arabidopsis, rice and Rumex, the terrestrial plants should be paid more attention in the future, such as wheat, maize, barely, tomato and soybean. More importantly, it is essential to collect germplasm resources of these terrestrial plants widely owing to phenotypic variation among natural populations, and then the key loci associated with flooding-tolerance can be screened through map-based cloning, genome-wide association study (GWAS) and precision genome editing with the aid of CRISPR-Cas9 system for breeding flooding-resistant varieties.
Secondly, the uptake of multiple mineral nutrients usually becomes disorganized under flooding stress. For example, the concentrations of N, P, K and Zn in leaves of Lepidium latifolium L. exhibited a decreasing trend as the flooding-time prolongs, and were lower than those in control plants, however, Fe and Mn contents were higher in roots of flooded plants than that in unflooded ones [163]. Mineral elements are crucial for plant growth and the deficiency of mineral nutrients under flooding conditions hastens adverse effects on plants. How plants maintain mineral nutrients homeostasis under waterlogging or submergence stress is largely unknown. Additionally, the formation of ROL barriers is one of the typical strategies for plant acclimation to flooding stress, however, its prevention of oxygen loss is nonspecific, and the uptake of water or nutrients is influenced simultaneously [164,165]. Thus, the mechanism of plants balancing the impediment of oxygen loss and the uptake of water or nutrients also needs to be further investigated.
Thirdly, flooding stress usually leads to changes in metabolic activities in plants, especially outbreaks of some harmful substances, such as ROS, MDA and acetaldehyde. For instance, flooding-induced oxygen deprivation facilitates the activation of anaerobic respiration and results in the generation of ethanol and acetaldehyde. Ethanol is usually considered to be innocuous for its rapid diffusion from cells while acetaldehyde can poison plant cells [7]. Under the prolonged flooding stress, how does the plant avoid the toxic of acetaldehyde? Whether the flooding-tolerance plants have enhanced capacity or other pathways to scavenge acetaldehyde? These questions remain to be further answered.
Finally, the mechanisms about plants with the ‘quiescent strategy’ adaption to submergence stress remain to be further investigated, especially the tolerant species that can endure complete submergence stress for a long time. For instance, more than 90% of bermudagrass can survive after complete submergence for 5 months with water depth up to 25 m [166]. Under deep submergence, bermudagrass experiences multiple stresses simultaneously, including oxygen deprivation, low temperature, high water pressure and low light or even no light. How does bermudagrass adapt to the stresses over the long term? Exploring the above scientific questions will facilitate understanding the mechanisms of plant acclimation to flooding stress and provide new insights into the breeding of flooding-resistant crops in the future.

Supplementary Materials

Author Contributions

S.W. and W.J. conceived this review, W.J. drafted the manuscript, M.M. and J.C. helped revise and check the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The project was financially supported by the Strategic Priority Research Program (A) of the Chinese Academy of Sciences (Grant No. XDA23040303), the Key Program of the Natural Science Foundation of Chongqing (Grant No. cstc2020jcyj-zdxmX0018), the Application of ecological restoration and function enhancement in the water-level-fluctuation zone of Guangyang Island (Grant No. E039620201) and the National Natural Science Foundation of China (Grant No. 42061015).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

Authors declare that they have no conflict of interest.

References

  1. Tanoue, M.; Hirabayashi, Y.; Ikeuchi, H. Global-scale river flood vulnerability in the last 50 years. Sci. Rep. 2016, 6, 36021. [Google Scholar] [CrossRef] [PubMed]
  2. Bailey-Serres, J.; Lee, S.C.; Brinton, E. Waterproofing crops: Effective flooding survival strategies. Plant Physiol. 2012, 160, 1698–1709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Nishiuchi, S.; Yamauchi, T.; Takahashi, H.; Kotula, L.; Nakazono, M. Mechanisms for coping with submergence and waterlogging in rice. Rice 2012, 5, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Fukao, T.; Barrera-Figueroa, B.E.; Juntawong, P.; Pena-Castro, J.M. Submergence and waterlogging stress in plants: A review highlighting research opportunities and understudied aspects. Front. Plant Sci. 2019, 10, 340. [Google Scholar] [CrossRef] [PubMed]
  5. Catling, D. Rice in Deep Water; Macmillan Press: London, UK, 1992. [Google Scholar]
  6. Setter, T.L.; Waters, I. Review of prospects for germplasm improvement for waterlogging tolerance in wheat, barley and oats. Plant Soil 2013, 253, 1–34. [Google Scholar] [CrossRef]
  7. Bailey-Serres, J.; Voesenek, L.A.C.J. Flooding stress: Acclimations and genetic diversity. Annu. Rev. Plant Biol. 2008, 59, 313–339. [Google Scholar] [CrossRef] [Green Version]
  8. Voesenek, L.A.C.J.; Bailey-Serres, J. Flood adaptive traits and processes: An overview. New Phytol. 2015, 206, 57–73. [Google Scholar] [CrossRef]
  9. Armstrong, W.; Drew, M.C. Root growth and metabolism under oxygen deficiency. In Plant Roots: The Hidden Half; Waisel, E.A., Kafkafi, Y., Eds.; Marcel Dekker: New York, NY, USA, 2002; pp. 729–761. [Google Scholar]
  10. Colmer, T.D.; Flowers, T.J. Flooding tolerance in halophytes. New Phytol. 2008, 179, 964–974. [Google Scholar] [CrossRef]
  11. Hattori, Y.; Nagai, K.; Ashikari, M. Rice growth adapting to deepwater. Curr. Opin. Plant Biol. 2011, 14, 100–105. [Google Scholar] [CrossRef]
  12. Colmer, T.D.; Voesenek, L.A.C.J. Flooding tolerance: Suites of plant traits in variable environments. Funct. Plant Biol. 2009, 36, 665–681. [Google Scholar] [CrossRef]
  13. Loreti, E.; van Veen, H.; Perata, P. Plant responses to flooding stress. Curr. Opin. Plant Biol. 2016, 33, 64–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Visser, E.J.W.; Colmer, T.D.; Blom, C.W.P.M.; Voesenek, L.A.C.J. Changes in growth, porosity, and radial oxygen loss from adventitious roots of selected mono- and dicotyledonous wetland species with contrasting types of aerenchyma. Plant Cell Environ. 2000, 23, 1237–1245. [Google Scholar] [CrossRef] [Green Version]
  15. Colmer, T.D. Aerenchyma and an inducible barrier to radial oxygen loss facilitate root aeration in upland, paddy and deep-water rice (Oryza sativa L.). Ann. Bot. 2003, 91, 301–309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Jackson, M.B. Ethylene-promoted elongation: An adaptation to submergence stress. Ann. Bot. 2008, 101, 229–248. [Google Scholar] [CrossRef]
  17. Wei, X.N.; Xu, H.J.; Rong, W.; Ye, X.G.; Zhang, Z.Y. Constitutive expression of a stabilized transcription factor group VII ethylene response factor enhances waterlogging tolerance in wheat without penalizing grain yield. Plant Cell Environ. 2019, 42, 1471–1485. [Google Scholar] [CrossRef]
  18. Nagai, K.; Mori, Y.; Ishikawa, S.; Furuta, T.; Gamuyao, R.L.; Niimi, Y.; Hobo, T.; Fukuda, M.; Kojima, M.; Takebayashi, Y.; et al. Antagonistic regulation of the gibberellic acid response during stem growth in rice. Nature 2020, 584, 109–114. [Google Scholar] [CrossRef]
  19. Tang, H.; Bi, H.; Liu, B.; Lou, S.L.; Song, Y.; Tong, S.F.; Chen, N.N.; Jiang, Y.Z.; Liu, J.Q.; Liu, H.H. WRKY33 interacts with WRKY12 protein to up-regulate RAP2.2 during submergence induced hypoxia response in Arabidopsis thaliana. New Phytol. 2021, 229, 106–125. [Google Scholar] [CrossRef]
  20. Bui, L.T.; Shukla, V.; Giorgi, F.M.; Trivellini, A.; Perata, P.; Licausi, F.; Giuntoli, B. Differential submergence tolerance between juvenile and adult Arabidopsis plants involves the ANAC017 transcription factor. Plant J. 2020, 104, 979–994. [Google Scholar] [CrossRef]
  21. Olgun, M.; Metin Kumlay, A.; Cemal Adiguzel, M.; Caglar, A. The effect of waterlogging in wheat (T. aestivum L.). Acta Agric. Scand. Sect. B Soil Plant Sci. 2008, 58, 193–198. [Google Scholar]
  22. Haque, M.E.; Kawaguchi, K.; Komatsu, S. Analysis of proteins in aerenchymatous seminal roots of wheat grown in hypoxic soils under waterlogged conditions. Protein Pept. Lett. 2011, 18, 912–924. [Google Scholar] [CrossRef]
  23. Malik, A.I.; Islam, A.K.M.R.; Colmer, T.D. Transfer of the barrier to radial oxygen loss in roots of Hordeum marinum to wheat (Triticum aestivum): Evaluation of four H. marinum-wheat amphiploids. New Phytol. 2011, 190, 499–508. [Google Scholar] [CrossRef] [PubMed]
  24. Panozzo, A.; Dal Cortivo, C.; Ferrari, M.; Vicelli, B.; Varotto, S.; Vamerali, T. Morphological changes and expressions of AOX1A, CYP81D8, and putative PFP genes in a large set of commercial maize hybrids under extreme waterlogging. Front. Plant Sci. 2019, 10, 62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Yu, F.; Liang, K.; Fang, T.; Zhao, H.L.; Han, X.S.; Cai, M.J.; Qiu, F.Z. A group VII ethylene response factor gene, ZmEREB180, coordinates waterlogging tolerance in maize seedlings. Plant Biotechnol. J. 2019, 17, 2286–2298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Ren, B.Z.; Zhang, J.W.; Li, X.; Fan, X.; Dong, S.T.; Liu, P.; Zhao, B. Effects of waterlogging on the yield and growth of summer maize under field conditions. Can. J. Plant Sci. 2014, 94, 23–31. [Google Scholar] [CrossRef]
  27. Baranwal, S.; Singh, B.B. Effect of waterlogging on growth, chlorophylls and saccharides content in maize genotypes. Indian J. Plant Physiol. 2002, 7, 246–251. [Google Scholar]
  28. Jackson, M.B.; Ram, P.C. Physiological and molecular basis of susceptibility and tolerance of rice plants to complete submergence. Ann. Bot. 2003, 91, 227–241. [Google Scholar] [CrossRef] [Green Version]
  29. Bailey-Serres, J.; Fukao, T.; Ronald, P.; Ismail, A.; Heuer, S.; Mackill, D. Submergence tolerant rice: SUB1’s journey from landrace to modern cultivar. Rice 2010, 3, 138–147. [Google Scholar] [CrossRef] [Green Version]
  30. Kende, H.; van der Knaap, E.; Cho, H.T. Deepwater rice: A model plant to study stem elongation. Plant Physiol. 1998, 118, 1105–1110. [Google Scholar] [CrossRef] [Green Version]
  31. Hattori, Y.; Nagai, K.; Furukawa, S.; Song, X.J.; Kawano, R.; Sakakibara, H.; Wu, J.Z.; Matsumoto, T.; Yoshimura, A.; Kitano, H.; et al. The ethylene response factors SNORKEL1 and SNORKEL2 allow rice to adapt to deep water. Nature 2009, 460, 1026–1030. [Google Scholar] [CrossRef]
  32. Githiri, S.M.; Watanabe, S.; Harada, K.; Takahashi, R. QTL analysis of flooding tolerance in soybean at an early vegetative growth stage. Plant Breed. 2006, 125, 613–618. [Google Scholar] [CrossRef]
  33. Tamang, B.G.; Magliozzi, J.O.; Maroof, M.A.; Fukao, T. Physiological and transcriptomic characterization of submergence and reoxygenation responses in soybean seedlings. Plant Cell Environ. 2014, 37, 2350–2365. [Google Scholar] [CrossRef] [Green Version]
  34. Wang, X.; Sakata, K.; Komatsu, S. An integrated approach of proteomics and computational genetic modification effectiveness analysis to uncover the mechanism of flood tolerance in soybeans. Int. J. Mol. Sci. 2018, 19, 1301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Garcia, N.; da-Silva, C.J.; Cocco, K.L.T.; Pomagualli, D.; de Oliveira, F.K.; da Silva, J.V.L.; de Oliveira, A.C.B.; do Amarante, L. Waterlogging tolerance of five soybean genotypes through different physiological and biochemical mechanisms. Environ. Exp. Bot. 2020, 172, 103975. [Google Scholar] [CrossRef]
  36. Striker, G.G. Flooding stress on plants: Anatomical, morphological and physiological responses. In Botany; Mworia, J.K., Ed.; InTech: Rijeka, Croatia, 2012. [Google Scholar]
  37. Steffens, B.; Rasmussen, A. The physiology of adventitious roots. Plant Physiol. 2016, 170, 603–617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Zhai, L.H.; Liu, Z.J.; Zou, X.L.; Jiang, Y.Y.; Qiu, F.Z.; Zheng, Y.L.; Zhang, Z.X. Genome-wide identification and analysis of microRNA responding to long-term waterlogging in crown roots of maize seedlings. Physiol. Plant 2013, 147, 181–193. [Google Scholar] [CrossRef]
  39. Lorbiecke, R.; Sauter, M. Adventitious root growth and cell-cycle induction in deepwater rice. Plant Physiol. 1999, 119, 21–30. [Google Scholar] [CrossRef] [Green Version]
  40. Mergemann, H.; Sauter, M. Ethylene induces epidermal cell death at the site of adventitious root emergence in rice. Plant Physiol. 2000, 124, 609–614. [Google Scholar] [CrossRef] [Green Version]
  41. Else, M.A.; Janowiak, F.; Atkinson, C.J.; Jackson, M.B. Root signals and stomatal closure in relation to photosynthesis, chlorophyll a fluorescence and adventitious rooting of flooded tomato plants. Ann. Bot. 2009, 103, 313–323. [Google Scholar] [CrossRef] [Green Version]
  42. Vidoz, M.L.; Loreti, E.; Mensuali, A.; Alpi, A.; Perata, P. Hormonal interplay during adventitious root formation in flooded tomato plants. Plant J. 2010, 63, 551–562. [Google Scholar] [CrossRef]
  43. Qi, X.H.; Li, Q.Q.; Ma, X.T.; Qian, C.L.; Wang, H.H.; Ren, N.N.; Shen, C.X.; Huang, S.M.; Xu, X.W.; Xu, Q.; et al. Waterlogging-induced adventitious root formation in cucumber is regulated by ethylene and auxin through reactive oxygen species signalling. Plant Cell Environ. 2019, 42, 1458–1470. [Google Scholar] [CrossRef]
  44. Visser, E.J.W.; Cohen, J.D.; Barendse, G.W.M.; Blom, C.W.P.M.; Voesenek, L.A.C.J. An ethylene-mediated increase in sensitivity to auxin induces adventitious root formation in flooded Rumex palustris Sm. Plant Physiol. 1996, 112, 1687–1692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Colmer, T.D.; Gibberd, M.R.; Wiengweera, A.; Tinh, T.K. The barrier to radial oxygen loss from roots of rice (Oryza sativa L.) is induced by growth in stagnant solution. J. Exp. Bot. 1998, 49, 1431–1436. [Google Scholar] [CrossRef]
  46. Kotula, L.; Ranathunge, K.; Schreiber, L.; Steudle, E. Functional and chemical comparison of apoplastic barriers to radial oxygen loss in roots of rice (Oryza sativa L.) grown in aerated or deoxygenated solution. J. Exp. Bot. 2009, 60, 2155–2167. [Google Scholar] [CrossRef] [PubMed]
  47. Steffens, B.; Geske, T.; Sauter, M. Aerenchyma formation in the rice stem and its promotion by H2O2. New Phytol. 2011, 190, 369–378. [Google Scholar] [CrossRef] [PubMed]
  48. Shiono, K.; Yamauchi, T.; Yamazaki, S.; Mohanty, B.; Malik, A.I.; Nagamura, Y.; Nishizawa, N.K.; Tsutsumi, N.; Colmer, T.D.; Nakazono, M. Microarray analysis of laser-microdissected tissues indicates the biosynthesis of suberin in the outer part of roots during formation of a barrier to radial oxygen loss in rice (Oryza sativa). J. Exp. Bot. 2014, 65, 4795–4806. [Google Scholar] [CrossRef] [Green Version]
  49. Abiko, T.; Kotula, L.; Shiono, K.; Malik, A.I.; Colmer, T.D.; Nakazono, M. Enhanced formation of aerenchyma and induction of a barrier to radial oxygen loss in adventitious roots of Zea nicaraguensis contribute to its waterlogging tolerance as compared with maize (Zea mays ssp mays). Plant Cell Environ. 2012, 35, 1618–1630. [Google Scholar] [CrossRef]
  50. Mano, Y.; Omori, F. Relationship between constitutive root aerenchyma formation and flooding tolerance in Zea nicaraguensis. Plant Soil 2013, 370, 447–460. [Google Scholar] [CrossRef]
  51. Watanabe, K.; Takahashi, H.; Sato, S.; Nishiuchi, S.; Omori, F.; Malik, A.I.; Colmer, T.D.; Mano, Y.; Nakazono, M. A major locus involved in the formation of the radial oxygen loss barrier in adventitious roots of teosinte Zea nicaraguensis is located on the short-arm of chromosome 3. Plant Cell Environ. 2017, 40, 304–316. [Google Scholar] [CrossRef] [Green Version]
  52. Mano, Y.; Omori, F.; Takamizo, T.; Kindiger, B.; Bird, R.M.; Loaisiga, C.H. Variation for root aerenchyma formation in flooded and non-flooded maize and teosinte seedlings. Plant Soil 2006, 281, 269–279. [Google Scholar] [CrossRef]
  53. Rajhi, I.; Yamauchi, T.; Takahashi, H.; Nishiuchi, S.; Shiono, K.; Watanabe, R.; Mliki, A.; Nagamura, Y.; Tsutsumi, N.; Nishizawa, N.K.; et al. Identification of genes expressed in maize root cortical cells during lysigenous aerenchyma formation using laser microdissection and microarray analyses. New Phytol. 2011, 190, 351–368. [Google Scholar] [CrossRef]
  54. Huang, B.R.; Johnson, J.W.; Nesmith, S.; Bridges, D.C. Growth, physiological and anatomical responses of 2 wheat genotypes to waterlogging and nutrient supply. J. Exp. Bot. 1994, 45, 193–202. [Google Scholar] [CrossRef]
  55. Jiang, Z.; Song, X.F.; Zhou, Z.Q.; Wang, L.K.; Li, J.W.; Deng, X.Y.; Fan, H.Y. Aerenchyma formation: Programmed cell death in adventitious roots of winter wheat (Triticum aestivum) under waterlogging. Funct. Plant Biol. 2010, 37, 748–755. [Google Scholar] [CrossRef]
  56. Xu, Q.T.; Yang, L.; Zhou, Z.Q.; Mei, F.Z.; Qu, L.H.; Zhou, G.S. Process of aerenchyma formation and reactive oxygen species induced by waterlogging in wheat seminal roots. Planta 2013, 238, 969–982. [Google Scholar] [CrossRef] [PubMed]
  57. Zhang, X.C.; Shabala, S.; Koutoulis, A.; Shabala, L.; Johnson, P.; Hayes, D.; Nichols, D.S.; Zhou, M.X. Waterlogging tolerance in barley is associated with faster aerenchyma formation in adventitious roots. Plant Soil 2015, 394, 355–372. [Google Scholar] [CrossRef]
  58. Laan, P.; Berrevoets, M.J.; Lythe, S.; Armstrong, W.; Blom, C.W.P.M. Root morphology and aerenchyma formation as indicators of the flood-tolerance of Rumex species. J. Ecol. 1989, 77, 693–703. [Google Scholar] [CrossRef]
  59. Visser, E.J.W.; Heijink, C.J.; Vanhout, K.J.G.M.; Voesenek, L.A.C.J.; Barendse, G.W.M.; Blom, C.W.P.M. Regulatory role of auxin in adventitious root-formation in two Species of Rumex, differing in their sensitivity to waterlogging. Physiol. Plant. 1995, 93, 116–122. [Google Scholar] [CrossRef] [Green Version]
  60. Wample, R.L.; Reid, D.M. Control of adventitious root production and hypocotyl hypertrophy of sunflower (Helianthus Annuus) in response to flooding. Physiol. Plant. 1978, 44, 351–358. [Google Scholar] [CrossRef]
  61. Jackson, M.B.; Attwood, P.A. Roots of willow (Salix viminalis L) show marked tolerance to oxygen shortage in flooded soils and in solution culture. Plant Soil 1996, 187, 37–45. [Google Scholar] [CrossRef]
  62. Mühlenbock, P.; Plaszczyca, M.; Plaszczyca, M.; Mellerowicz, E.; Karpinski, S. Lysigenous aerenchyma formation in Arabidopsis is controlled by LESION SIMULATING DISEASE1. Plant Cell 2007, 19, 3819–3830. [Google Scholar] [CrossRef] [Green Version]
  63. Dawood, T.; Rieu, I.; Wolters-Arts, M.; Derksen, E.B.; Mariani, C.; Visser, E.J.W. Rapid flooding-induced adventitious root development from preformed primordia in Solanum dulcamara. AoB Plants 2014, 6. [Google Scholar] [CrossRef] [Green Version]
  64. Hua, J.F.; Han, L.W.; Wang, Z.Q.; Gu, C.S.; Yin, Y.L. Morpho-anatomical and photosynthetic responses of Taxodium hybrid ‘Zhongshanshan’ 406 to prolonged flooding. Flora 2017, 231, 29–37. [Google Scholar] [CrossRef]
  65. Xiao, B.; Jespersen, D. Morphological and physiological responses of seashore paspalum and bermudagrass to waterlogging stress. J. Am. Soc. Hortic. Sci. 2019, 144, 305–313. [Google Scholar] [CrossRef]
  66. Ejiri, M.; Shiono, K. Prevention of radial oxygen loss is associated with exodermal suberin along adventitious roots of annual wild species of Echinochloa. Front. Plant Sci. 2019, 10, 254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Jia, L.T.; Qin, X.; Lyu, D.G.; Qin, S.J.; Zhang, P. ROS production and scavenging in three cherry rootstocks under short-term waterlogging conditions. Sci. Hortic. 2019, 257, 108647. [Google Scholar] [CrossRef]
  68. Ye, T.T.; Shi, H.T.; Wang, Y.P.; Chan, Z.L. Contrasting changes caused by drought and submergence stresses in bermudagrass (Cynodon dactylon). Front. Plant Sci. 2015, 6, 951. [Google Scholar] [CrossRef] [Green Version]
  69. Wang, J.; Sun, H.; Sheng, J.J.; Jin, S.R.; Zhou, F.S.; Hu, Z.L.; Diao, Y. Transcriptome, physiological and biochemical analysis of Triarrhena sacchariflora in response to flooding stress. BMC Genet. 2019, 20, 88. [Google Scholar] [CrossRef]
  70. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 2002, 7, 405–410. [Google Scholar] [CrossRef]
  71. Choudhury, F.K.; Rivero, R.M.; Blumwald, E.; Mittler, R. Reactive oxygen species, abiotic stress and stress combination. Plant J. 2017, 90, 856–867. [Google Scholar] [CrossRef]
  72. Gill, M.B.; Zeng, F.R.; Shabala, L.; Zhang, G.P.; Yu, M.; Demidchik, V.; Shabala, S.; Zhou, M.X. Identification of QTL related to ROS formation under hypoxia and their association with waterlogging and salt tolerance in barley. Int. J. Mol. Sci. 2019, 20, 699. [Google Scholar] [CrossRef] [Green Version]
  73. Yordanova, R.Y.; Christov, K.N.; Popova, L.P. Antioxidative enzymes in barley plants subjected to soil flooding. Environ. Exp. Bot. 2004, 51, 93–101. [Google Scholar] [CrossRef]
  74. Damanik, R.I.; Maziah, M.; Ismail, M.R.; Ahmad, S.; Zain, A. Responses of the antioxidative enzymes in Malaysian rice (Oryza sativa L.) cultivars under submergence condition. Acta Physiol. Plant. 2010, 32, 739–747. [Google Scholar] [CrossRef]
  75. Tan, W.; Liu, J.; Dai, T.; Jing, Q.; Cao, W.; Jiang, D. Alterations in photosynthesis and antioxidant enzyme activity in winter wheat subjected to post-anthesis waterlogging. Photosynthetica 2008, 46, 21–27. [Google Scholar] [CrossRef]
  76. Tang, B.; Xu, S.Z.; Zou, X.L.; Zheng, Y.L.; Qiu, F.Z. Changes of antioxidative enzymes and lipid peroxidation in leaves and roots of waterlogging-tolerant and waterlogging-sensitive maize genotypes at seedling stage. Agr. Sci. China 2010, 9, 651–661. [Google Scholar] [CrossRef]
  77. Arbona, V.; Hossain, Z.; Lopez-Climent, M.F.; Perez-Clemente, R.M.; Gomez-Cadenas, A. Antioxidant enzymatic activity is linked to waterlogging stress tolerance in citrus. Physiol. Plantarum 2008, 132, 452–466. [Google Scholar] [CrossRef] [PubMed]
  78. Hossain, Z.; Lopez-Climent, M.F.; Arbona, V.; Perez-Clemente, R.M.; Gomez-Cadenas, A. Modulation of the antioxidant system in citrus under waterlogging and subsequent drainage. J. Plant Physiol. 2009, 166, 1391–1404. [Google Scholar] [CrossRef]
  79. Yu, Q.; Rengel, Z. Waterlogging influences plant growth and activities of superoxide dismutases in narrow-leafed lupin and transgenic tobacco plants. J. Plant Physiol. 1999, 155, 431–438. [Google Scholar] [CrossRef]
  80. Ushimaru, T.; Kanematsu, S.; Katayama, M.; Tsuji, H. Antioxidative enzymes in seedlings of Nelumbo nucifera germinated under water. Physiol. Plant. 2001, 112, 39–46. [Google Scholar] [CrossRef]
  81. Wang, K.H.; Jiang, Y.W. Antioxidant responses of creeping bentgrass roots to waterlogging. Crop Sci. 2007, 47, 232–238. [Google Scholar] [CrossRef]
  82. Kumutha, D.; Ezhilmathi, K.; Sairam, R.K.; Srivastava, G.C.; Deshmukh, P.S.; Meena, R.C. Waterlogging induced oxidative stress and antioxidant activity in pigeonpea genotypes. Biol. Plant. 2009, 53, 75–84. [Google Scholar] [CrossRef]
  83. Sairam, R.K.; Kumutha, D.; Ezhilmathi, K.; Chinnusamy, V.; Meena, R.C. Waterlogging induced oxidative stress and antioxidant enzyme activities in pigeon pea. Biol. Plant. 2009, 53, 493–504. [Google Scholar] [CrossRef]
  84. Liu, M.X.; Jiang, Y.W. Genotypic variation in growth and metabolic responses of perennial ryegrass exposed to short-term waterlogging and submergence stress. Plant Physiol. Biochem. 2015, 95, 57–64. [Google Scholar] [CrossRef] [PubMed]
  85. Voesenek, L.A.C.J.; Bailey-Serres, J. Flooding tolerance: O2 sensing and survival strategies. Curr. Opin. Plant Biol. 2013, 16, 647–653. [Google Scholar] [CrossRef] [PubMed]
  86. Vervuren, P.J.A.; Blom, C.W.P.M.; de Kroon, H. Extreme flooding events on the Rhine and the survival and distribution of riparian plant species. J. Ecol. 2003, 91, 135–146. [Google Scholar] [CrossRef]
  87. Loreti, E.; Valeri, M.C.; Novi, G.; Perata, P. Gene regulation and survival under hypoxia requires starch availability and metabolism. Plant Physiol. 2018, 176, 1286–1298. [Google Scholar] [CrossRef]
  88. Tian, L.X.; Li, J.; Bi, W.S.; Zuo, S.Y.; Li, L.J.; Li, W.L.; Sun, L. Effects of waterlogging stress at different growth stages on the photosynthetic characteristics and grain yield of spring maize (Zea mays L.) under field conditions. Agric. Water Manag. 2019, 218, 250–258. [Google Scholar] [CrossRef]
  89. Evans, D.E. Aerenchyma formation. New Phytol. 2003, 161, 35–49. [Google Scholar] [CrossRef]
  90. Yamauchi, T.; Watanabe, K.; Fukazawa, A.; Mori, H.; Abe, F.; Kawaguchi, K.; Oyanagi, A.; Nakazono, M. Ethylene and reactive oxygen species are involved in root aerenchyma formation and adaptation of wheat seedlings to oxygen-deficient conditions. J. Exp. Bot. 2014, 65, 261–273. [Google Scholar] [CrossRef] [Green Version]
  91. McDonald, M.P.; Galwey, N.W.; Colmer, T.D. Similarity and diversity in adventitious root anatomy as related to root aeration among a range of wetland and dryland grass species. Plant Cell Environ. 2002, 25, 441–451. [Google Scholar] [CrossRef]
  92. Soukup, A.; Armstrong, W.; Schreiber, L.; Franke, R.; Votrubova, O. Apoplastic barriers to radial oxygen loss and solute penetration: A chemical and functional comparison of the exodermis of two wetland species, Phragmites australis and Glyceria maxima. New Phytol. 2007, 173, 264–278. [Google Scholar] [CrossRef]
  93. Shiono, K.; Ogawa, S.; Yamazaki, S.; Isoda, H.; Fujimura, T.; Nakazono, M.; Colmer, T.D. Contrasting dynamics of radial O2-loss barrier induction and aerenchyma formation in rice roots of two lengths. Ann. Bot. 2011, 107, 89–99. [Google Scholar] [CrossRef]
  94. Kotula, L.; Colmer, T.D.; Nakazono, M. Effects of organic acids on the formation of the barrier to radial oxygen loss in roots of Hordeum marinum. Funct. Plant Biol. 2014, 41, 187–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Colmer, T.D.; Kotula, L.; Malik, A.; Takahashi, H.; Konnerup, D.; Nakazono, M. Rice acclimation to soil flooding: Low concentrations of organic acids can trigger a barrier to radial oxygen loss in roots. Plant Cell Environ. 2019, 42, 2183–2197. [Google Scholar] [CrossRef] [PubMed]
  96. Rao, D.N.; Mikkelsen, D.S. Effects of acetic, propionic, and butyric acids on rice seedling growth and nutrition. Plant Soil 1977, 47, 323–334. [Google Scholar] [CrossRef]
  97. Armstrong, J.; Armstrong, W. Rice and Phragmites: Effects of organic acids on growth, root permeability, and radial oxygen loss to the rhizosphere. Am. J. Bot. 2001, 88, 1359–1370. [Google Scholar] [CrossRef]
  98. Dolferus, R.; Wolansky, M.; Carroll, R.; Miyashita, Y.; Ismond, K.; Good, A. Functional analysis of lactate dehydrogenase during hypoxic stress in Arabidopsis. Funct. Plant Biol. 2008, 35, 131–140. [Google Scholar] [CrossRef] [PubMed]
  99. Hinz, M.; Wilson, I.W.; Yang, J.; Buerstenbinder, K.; Llewellyn, D.J.; Dennis, E.S.; Sauter, M.; Dolferus, R. Arabidopsis RAP2.2: An ethylene response transcription factor that is important for hypoxia survival. Plant Physiol. 2010, 153, 757–772. [Google Scholar] [CrossRef] [Green Version]
  100. Licausi, F.; Van Dongen, J.T.; Giuntoli, B.; Novi, G.; Santaniello, A.; Geigenberger, P.; Perata, P. HRE1 and HRE2, two hypoxia-inducible ethylene response factors, affect anaerobic responses in Arabidopsis thaliana. Plant J. 2010, 62, 302–315. [Google Scholar] [CrossRef]
  101. Licausi, F.; Kosmacz, M.; Weits, D.A.; Giuntoli, B.; Giorgi, F.M.; Voesenek, L.A.C.J.; Perata, P.; Van Dongen, J.T. Oxygen sensing in plants is mediated by an N-end rule pathway for protein destabilization. Nature 2011, 479, 419–422. [Google Scholar] [CrossRef]
  102. Weits, D.A.; Giuntoli, B.; Kosmacz, M.; Parlanti, S.; Hubberten, H.M.; Riegler, H. Plant cysteine oxidases control the oxygen-dependent branch of the N-end-rule pathway. Nat. Commun. 2014, 5, 3425. [Google Scholar] [CrossRef] [Green Version]
  103. Lin, I.S.; Wu, Y.S.; Chen, C.T.; Chen, G.H.; Hwang, S.G.; Jauh, G.Y.; Tzen, J.T.; Yang, C.Y. AtRBOH I confers submergence tolerance and is involved in auxin-mediated signaling pathways under hypoxic stress. Plant Growth Regul. 2017, 83, 277–285. [Google Scholar] [CrossRef]
  104. Xu, K.N.; Xu, X.; Fukao, T.; Canlas, P.; Maghirang-Rodriguez, R.; Heuer, S.; Ismail, A.M.; Bailey-Serres, J.; Ronald, P.C.; Mackill, D.J. Sub1A is an ethylene-response-factor-like gene that confers submergence tolerance to rice. Nature 2006, 442, 705–708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Lee, K.W.; Chen, P.W.; Lu, C.A.; Chen, S.; Ho, T.H.D.; Yu, S.M. Coordinated responses to oxygen and sugar deficiency allow rice seedlings to tolerate flooding. Sci. Signal. 2009, 2, ra61. [Google Scholar] [CrossRef] [PubMed]
  106. Minji, P.; Huikyeong, Y.; Hyeokgon, P.; Jun, L.; Soohwan, K.; Yongsic, H. Interference with oxidative phosphorylation enhances anoxic expression of rice α-amylase genes through abolishing sugar regulation. J. Exp. Bot. 2010, 61, 3235–3244. [Google Scholar]
  107. Kretzschmar, T.; Pelayo, M.A.; Trijatmiko, K.R.; Gabunada, L.F.; Alam, R.; Jimenez, R.; Mendioro, M.S.; Slamet-Loedin, I.H.; Sreenivasulu, N.; Bailey-Serres, J.; et al. A trehalose-6-phosphate phosphatase enhances anaerobic germination tolerance in rice. Nat. Plants 2015, 1, 15124. [Google Scholar] [CrossRef] [PubMed]
  108. Jisha, V.; Dampanaboina, L.; Vadassery, J.; Mithöfer, A.; Kappara, S.; Ramanan, R. Overexpression of an AP2/ERF type transcription factor OsEREBP1 confers biotic and abiotic stress tolerance in rice. PLoS ONE 2015, 10, e0127831. [Google Scholar] [CrossRef]
  109. Kurokawa, Y.; Nagai, K.; Huan, P.D.; Shimazaki, K.; Qu, H.Q.; Mori, Y.; Toda, Y.; Kuroha, T.; Hayashi, N.; Aiga, S.; et al. Rice leaf hydrophobicity and gas films are conferred by a wax synthesis gene (LGF1) and contribute to flood tolerance. New Phytol. 2018, 218, 1558–1569. [Google Scholar] [CrossRef] [Green Version]
  110. Kuroha, T.; Nagai, K.; Gamuyao, R.; Wang, D.R.; Furuta, T.; Nakamori, M.; Kitaoka, T.; Adachi, K.; Minami, A.; Mori, Y.; et al. Ethylene-gibberellin signaling underlies adaptation of rice to periodic flooding. Science 2018, 361, 181–185. [Google Scholar] [CrossRef] [Green Version]
  111. Liang, S.; Xiong, W.; Yin, C.; Xie, X.; Jin, Y.J.; Zhang, S.; Yang, B.; Ye, G.Y.; Chen, S.Y.; Luan, W.J. Overexpression of OsARD1 improves submergence, drought, and salt tolerances of seedling through the enhancement of ethylene synthesis in rice. Front. Plant Sci. 2019, 10, 1088. [Google Scholar] [CrossRef]
  112. Campbell, M.T.; Proctor, C.A.; Dou, Y.C.; Schmitz, A.J.; Phansak, P.; Kruger, G.R.; Zhang, C.; Walia, H. Genetic and molecular characterization of submergence response identifies Subtol6 as a major submergence tolerance locus in maize. PLoS ONE 2015, 10, e0120385. [Google Scholar] [CrossRef]
  113. Luan, H.Y.; Guo, B.J.; Shen, H.Q.; Pan, Y.H.; Hong, Y.; Lv, C.; Xu, R.G. Overexpression of barley transcription factor HvERF2.11 in Arabidopsis enhances plant waterlogging tolerance. Int. J. Mol. Sci. 2020, 21, 1982. [Google Scholar] [CrossRef] [Green Version]
  114. Zhang, J.Y.; Huang, S.N.; Wang, G.; Xuan, J.P.; Guo, Z.R. Overexpression of Actinidia deliciosa pyruvate decarboxylase 1 gene enhances waterlogging stress in transgenic Arabidopsis thaliana. Plant Physiol. Biochem. 2016, 106, 244–252. [Google Scholar] [CrossRef] [PubMed]
  115. Pan, D.L.; Wang, G.; Wang, T.; Jia, Z.H.; Guo, Z.R.; Zhang, J.Y. AdRAP2.3, a novel ethylene response factor VII from Actinidia deliciosa, enhances waterlogging resistance in transgenic tobacco through improving expression levels of PDC and ADH genes. Int. J. Mol. Sci. 2019, 20, 1189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Wang, L.J.; Gao, J.J.; Zhang, Z.X.; Liu, W.M.; Cheng, P.L.; Mu, W.T.; Su, T.; Chen, S.M.; Chen, F.D.; Jiang, J.F. Overexpression of CmSOS1 confers waterlogging tolerance in Chrysanthemum. J. Integr. Plant Biol. 2020, 62, 1059–1064. [Google Scholar] [CrossRef] [PubMed]
  117. Xu, X.W.; Ji, J.; Xu, Q.; Qi, X.H.; Weng, Y.Q.; Chen, X.H. The major-effect quantitative trait locus CsARN6.1 encodes an AAA ATPase domain-containing protein that is associated with waterlogging stress tolerance by promoting adventitious root formation. Plant J. 2018, 93, 917–930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Phukan, U.J.; Jeena, G.S.; Tripathi, V.; Shukla, R.K. MaRAP2-4, a waterlogging-responsive ERF from Mentha, regulates bidirectional sugar transporter AtSWEET10 to modulate stress response in Arabidopsis. Plant Biotechnol. J. 2018, 16, 221–233. [Google Scholar] [CrossRef] [Green Version]
  119. Yin, D.M.; Sun, D.Y.; Han, Z.Q.; Ni, D.; Norris, A.; Jiang, C.Z. PhERF2, an ethylene-responsive element binding factor, plays an essential role in waterlogging tolerance of petunia. Hortic. Res. 2019, 6, 83. [Google Scholar] [CrossRef] [Green Version]
  120. Liu, S.; Wang, J.; Wang, L.; Wang, X.; Xue, Y.; Wu, P.; Shou, H.X. Adventitious root formation in rice requires OsGNOM1 and is mediated by the OsPINs family. Cell Res. 2009, 19, 1110–1119. [Google Scholar] [CrossRef] [Green Version]
  121. Steinmann, T.; Geldner, N.; Grebe, M.; Mangold, S.; Jackson, C.L.; Paris, S.; Gälweiler, L.; Palme, K.; Jürgens, G. Coordinated polar localization of auxin efflux carrier PIN1 by GNOM ARF GEF. Science 1999, 286, 316–318. [Google Scholar] [CrossRef]
  122. Geldner, N.; Anders, N.; Wolters, H.; Keicher, J.; Kornberger, W.; Muller, P.; Delbarre, A.; Ueda, T.; Nakano, A.; Jürgens, G. The Arabidopsis GNOM ARF-GEF mediates endosomal recycling, auxin transport, and auxin-dependent plant growth. Cell 2003, 112, 219–230. [Google Scholar] [CrossRef]
  123. Steffens, B.; Sauter, M. Epidermal cell death in rice is confined to cells with a distinct molecular identity and is mediated by ethylene and H2O2 through an autoamplified signal pathway. Plant Cell 2009, 21, 184–196. [Google Scholar] [CrossRef] [Green Version]
  124. Steffens, B.; Kovalev, A.; Gorb, S.N.; Sauter, M. Emerging roots alter epidermal cell fate through mechanical and reactive oxygen species signaling. Plant Cell 2012, 24, 3296–3306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Xu, M.; Zhu, L.; Shou, H.; Wu, P. A PIN1 family gene, OsPIN1, involved in auxin-dependent adventitious root emergence and tillering in rice. Plant Cell Physiol. 2005, 46, 1674–1681. [Google Scholar] [CrossRef] [PubMed]
  126. Dawood, T.; Yang, X.; Visser, E.J.; Te Beek, T.A.; Kensche, P.R.; Cristescu, S.M.; Lee, S.; Floková, K.; Nguyen, D.; Mariani, C.; et al. A co-opted hormonal cascade activates dormant adventitious root primordia upon flooding in Solanum dulcamara. Plant Physiol. 2016, 170, 2351–2364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Overmyer, K.; Brosche, M.; Kangasjarvi, J. Reactive oxygen species and hormonal control of cell death. Trends Plant Sci. 2003, 8, 335–342. [Google Scholar] [CrossRef]
  128. Torres, M.A.; Dangl, J.L. Functions of the respiratory burst oxidase in biotic interactions, abiotic stress and development. Curr. Opin. Plant Biol. 2005, 8, 397–403. [Google Scholar] [CrossRef] [PubMed]
  129. Yamauchi, T.; Rajhi, I.; Nakazono, M. Lysigenous aerenchyma formation in maize root is confined to cortical cells by regulation of genes related to generation and scavenging of reactive oxygen species. Plant Signal. Behav. 2011, 6, 759–761. [Google Scholar] [CrossRef] [Green Version]
  130. Gonzali, S.; Loreti, E.; Cardarelli, F.; Novi, G.; Parlanti, S.; Pucciariello, C.; Bassolino, L.; Banti, V.; Licausi, F.; Perata, P. Universal stress protein HRU1 mediates ROS homeostasis under anoxia. Nat. Plants 2015, 1, 15151. [Google Scholar] [CrossRef]
  131. Rivera-Contreras, I.K.; Zamora-Hernandez, T.; Huerta-Heredia, A.A.; Capataz-Tafur, J.; Barrera-Figueroa, B.E.; Juntawong, P.; Peña-Castro, J.M. Transcriptomic analysis of submergence-tolerant and sensitive Brachypodium distachyon ecotypes reveals oxidative stress as a major tolerance factor. Sci. Rep. 2016, 6, 27686. [Google Scholar] [CrossRef] [Green Version]
  132. Jackson, M.B.; Armstrong, W. Formation of aerenchyma and the processes of plant ventilation in relation to soil flooding and submergence. Plant Biol. 1999, 1, 274–287. [Google Scholar] [CrossRef]
  133. De Simone, O.; Haase, K.; Muller, E.; Junk, W.J.; Hartmann, K.; Schreiber, L.; Schmidt, W. Apoplasmic barriers and oxygen transport properties of hypodermal cell walls in roots from four Amazonian tree species. Plant Physiol. 2003, 132, 206–217. [Google Scholar] [CrossRef] [Green Version]
  134. Fleck, A.T.; Nye, T.; Repenning, C.; Stahl, F.; Zahn, M.; Schenk, M.K. Silicon enhances suberization and lignification in roots of rice (Oryza sativa). J. Exp. Bot. 2011, 62, 2001–2011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Soler, M.; Serra, O.; Molinas, M.; Huguet, G.; Fluch, S.; Figueras, M. A genomic approach to suberin biosynthesis and cork differentiation. Plant Physiol. 2007, 144, 419–431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Ranathunge, K.; Schreiber, L.; Franke, R. Suberin research in the genomics era-new interest for an old polymer. Plant Sci. 2011, 180, 399–413. [Google Scholar] [CrossRef] [PubMed]
  137. Almeida, T.; Menéndez, E.; Capote, T.; Ribeiro, T.; Santos, C.; Gonçalves, S. Molecular characterization of Quercus suber MYB1, a transcription factor up-regulated in cork tissues. J. Plant Physiol. 2013, 170, 172–178. [Google Scholar] [CrossRef]
  138. Gray, W.M. Hormonal regulation of plant growth and development. PLoS Biol. 2004, 2, 306–313. [Google Scholar] [CrossRef] [Green Version]
  139. Wolters, H.; Jürgens, G. Survival of the flexible: Hormonal growth control and adaptation in plant development. Nat. Rev. Genet. 2009, 10, 305–317. [Google Scholar] [CrossRef]
  140. van Veen, H.; Akman, M.; Jamar, D.C.; Vreugdenhil, D.; Kooiker, M.; van Tienderen, P.; Voesenek, L.A.C.J.; Schranz, M.E.; Sasidharan, R. Group VII ethylene response factor diversification and regulation in four species from flood-prone environments. Plant Cell Environ. 2014, 37, 2421–2432. [Google Scholar] [CrossRef]
  141. Sasidharan, R.; Voesenek, L.A.C.J. Ethylene-mediated acclimations to flooding stress. Plant Physiol. 2015, 169, 3–12. [Google Scholar] [CrossRef] [Green Version]
  142. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [Green Version]
  143. Yamamoto, F.; Kozlowski, T.T. Regulation by auxin and ethylene of responses of Acer negundo seedlings to flooding of soil. Environ. Exp. Bot. 1987, 27, 329–335, 337–340. [Google Scholar] [CrossRef]
  144. Leyser, O. Auxin signaling. Plant Physiol. 2018, 176, 465–479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Verstraeten, I.; Beeckman, T.; Geelen, D. Adventitious root induction in Arabidopsis thaliana as a model for in vitro root organogenesis. Methods Mol. Biol. 2013, 959, 159–175. [Google Scholar] [PubMed]
  146. Agulló-Antón, M.Á.; Ferrández-Ayela, A.; Fernández-García, N.; Nicolás, C.; Albacete, A.; Pérez-Alfocea, F.; Sánchez-Bravo, J.; Pérez-Pérez, J.M.; Acosta, M. Early steps of adventitious rooting: Morphology, hormonal profiling and carbohydrate turnover in carnation stem cuttings. Physiol. Plant. 2014, 150, 446–462. [Google Scholar] [CrossRef] [PubMed]
  147. Gutierrez, L.; Bussell, J.D.; Pacurar, D.I.; Schwambach, J.; Pacurar, M.; Bellini, C. Phenotypic plasticity of adventitious rooting in Arabidopsis is controlled by complex regulation of AUXIN RESPONSE FACTOR transcripts and microRNA abundance. Plant Cell 2009, 21, 3119–3132. [Google Scholar] [CrossRef] [Green Version]
  148. Gutierrez, L.; Mongelard, G.; Flokova, K.; Pacurar, D.I.; Novak, O.; Staswick, P.; Kowalczyk, M.; Păcurar, M.; Demailly, H.; Geiss, G.; et al. Auxin controls Arabidopsis adventitious root initiation by regulating jasmonic acid homeostasis. Plant Cell 2012, 24, 2515–2527. [Google Scholar] [CrossRef] [Green Version]
  149. Fukao, T.; Bailey-Serres, J. Submergence tolerance conferred by Sub1A is mediated by SLR1 and SLRL1 restriction of gibberellin responses in rice. Proc. Natl. Acad. Sci. USA 2008, 105, 16814–16819. [Google Scholar] [CrossRef] [Green Version]
  150. Ikeda, A.; Ueguchi-Tanaka, M.; Sonoda, Y.; Kitano, H.; Koshioka, M.; Futsuhara, Y.; Matsuoka, M.; Yamaguchi, J. slender rice, a constitutive gibberellin response mutant, is caused by a null mutation of the SLR1 gene, an ortholog of the height-regulating gene GAI/RGA/RHT/D8. Plant Cell 2001, 13, 999–1010. [Google Scholar]
  151. Itoh, H.; Ueguchi-Tanaka, M.; Sato, Y.; Ashikari, M.; Matsuoka, M. The gibberellin signaling pathway is regulated by the appearance and disappearance of SLENDER RICE1 in nuclei. Plant Cell 2002, 14, 57–70. [Google Scholar] [CrossRef] [Green Version]
  152. Itoh, H.; Shimada, A.; Ueguchi-Tanaka, M.; Kamiya, N.; Hasegawa, Y.; Ashikari, M.; Matsuoka, M. Overexpression of a GRAS protein lacking the DELLA domain confers altered gibberellin responses in rice. Plant J. 2005, 44, 669–679. [Google Scholar] [CrossRef]
  153. Fukao, T.; Xu, K.N.; Ronald, P.C.; Bailey-Serres, J. A variable cluster of ethylene response factor-like genes regulates metabolic and developmental acclimation responses to submergence in rice. Plant Cell 2006, 18, 2021–2034. [Google Scholar] [CrossRef] [Green Version]
  154. Ueguchi-Tanaka, M.; Ashikari, M.; Nakajima, M.; Itoh, H.; Katoh, E.; Kobayashi, M. GIBBERELLIN INSENSITIVE DWARF1 encodes a soluble receptor for gibberellin. Nature 2005, 437, 693–698. [Google Scholar] [CrossRef] [PubMed]
  155. Du, H.; Chang, Y.; Huang, F.; Xiong, L.Z. GID1 modulates stomatal response and submergence tolerance involving abscisic acid and gibberellic acid signaling in rice. J. Integr. Plant Biol. 2015, 57, 954–968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Zhou, X.; Zhang, Z.L.; Park, J.; Tyler, L.; Yusuke, J.; Qiu, K. The ERF11 transcription factor promotes internode elongation by activating gibberellin biosynthesis and signaling. Plant Physiol. 2016, 171, 2760–2770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Dubois, V.; Moritz, T.; Garcia-Martinez, J.L. Comparison of the role of gibberellins and ethylene in response to submergence of two lowland rice cultivars, Senia and Bomba. J. Plant Physiol. 2011, 168, 233–241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Jackson, M.B.; Hall, K.C. Early stomatal closure in waterlogged pea-plants is mediated by abscisic-acid in the absence of foliar water deficits. Plant Cell Environ. 1987, 10, 121–130. [Google Scholar]
  159. Zhang, J.; Van Toai, T.; Huynh, L.; Preiszner, J. Development of flooding-tolerant Arabidopsis thaliana by autoregulated cytokinin production. Mol. Breed. 2000, 6, 135–144. [Google Scholar] [CrossRef]
  160. Nguyen, T.N.; Tuan, P.A.; Mukherjee, S.; Son, S.; Ayele, B.T. Hormonal regulation in adventitious roots and during their emergence under waterlogged conditions in wheat. J. Exp. Bot. 2018, 69, 4065–4082. [Google Scholar] [CrossRef]
  161. Kim, Y.H.; Hwang, S.J.; Waqas, M.; Khan, A.L.; Lee, J.H.; Lee, J.D.; Nguyen, H.T.; Lee, I. Comparative analysis of endogenous hormones level in two soybean (Glycine max L.) lines differing in waterlogging tolerance. Front. Plant Sci. 2015, 6, 714. [Google Scholar] [CrossRef] [Green Version]
  162. Schmitz1, A.J.; Folsom, J.J.; Jikamaru, Y.; Ronald, P.; Walia, H. SUB1A-mediated submergence tolerance response in rice involves differential regulation of the brassinosteroid pathway. New Phytol. 2013, 198, 1060–1070. [Google Scholar] [CrossRef] [Green Version]
  163. Chen, H.J.; Qualls, R.G.; Blank, R.R. Effect of soil flooding on photosynthesis, carbohydrate partitioning and nutrient uptake in the invasive exotic Lepidium latifolium. Aquat. Bot. 2005, 82, 250–268. [Google Scholar] [CrossRef]
  164. Colmer, T.D.; Bloom, A.J. A comparison of NH4+ and NO3- net fluxes along roots of rice and maize. Plant Cell Environ. 1998, 21, 240–246. [Google Scholar] [CrossRef]
  165. Armstrong, J.; Armstrong, W. Rice: Sulfide-induced barriers to root radial oxygen loss, Fe2+ and water uptake, and lateral root emergence. Ann. Bot. 2005, 96, 625–638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Tan, S.D.; Zhu, M.Y.; Zhang, Q.F. Physiological responses of bermudagrass (Cynodon dactylon) to submergence. Acta Physiol. Plant. 2010, 32, 133–140. [Google Scholar] [CrossRef]
Figure 1. Response and the adaptive mechanisms of plants under flooding (hypoxic) stress. (A) Plants morphology under waterlogging and submergence stress. (B) The physiological changes of plants in response to flooding stress. (C) The main anatomical characteristics of plants adaptation to flooding stress. Triangles and pentagrams point to the location of aerenchyma and radial O2 loss (ROL) barriers respectively.
Figure 1. Response and the adaptive mechanisms of plants under flooding (hypoxic) stress. (A) Plants morphology under waterlogging and submergence stress. (B) The physiological changes of plants in response to flooding stress. (C) The main anatomical characteristics of plants adaptation to flooding stress. Triangles and pentagrams point to the location of aerenchyma and radial O2 loss (ROL) barriers respectively.
Ijms 22 01088 g001
Figure 2. The molecular mechanisms of ROL barriers, ARs and aerenchyma formation. Researches about the formation of flooding-induced ROL barriers at the molecular level are relatively deficient, however, suberin has been considered to contribute to the formation of ROL barriers and genes (such as KCS, CYP, GPAT, ABC, LTP, αβHD, POD) involved in suberin biosynthesis may function in this process [48]. For ARs formation, flooding (hypoxic) stress can induce expressions of RBOHB, RBOHF3, some Ethylene Response Factor (ERF) genes and ARN, then affects ARs formation directly or through ROS regulation [25,43,117]. Plants also can influence the development of ARs by mediating auxin polar transport through the activation of some auxin associated genes, such as GNOM1, PIN1 and PIN2 [120,125,126]. Aerenchyma formation depends on the accumulation of ROS. RBOH for the generation of ROS and MT for ROS scavenging play important roles in this process [53]. In addition, genes LSD1, EDS1 and PAD4, which control the generation of ethylene and ROS in the upstream, are also involved in the formation of aerenchyma [62].
Figure 2. The molecular mechanisms of ROL barriers, ARs and aerenchyma formation. Researches about the formation of flooding-induced ROL barriers at the molecular level are relatively deficient, however, suberin has been considered to contribute to the formation of ROL barriers and genes (such as KCS, CYP, GPAT, ABC, LTP, αβHD, POD) involved in suberin biosynthesis may function in this process [48]. For ARs formation, flooding (hypoxic) stress can induce expressions of RBOHB, RBOHF3, some Ethylene Response Factor (ERF) genes and ARN, then affects ARs formation directly or through ROS regulation [25,43,117]. Plants also can influence the development of ARs by mediating auxin polar transport through the activation of some auxin associated genes, such as GNOM1, PIN1 and PIN2 [120,125,126]. Aerenchyma formation depends on the accumulation of ROS. RBOH for the generation of ROS and MT for ROS scavenging play important roles in this process [53]. In addition, genes LSD1, EDS1 and PAD4, which control the generation of ethylene and ROS in the upstream, are also involved in the formation of aerenchyma [62].
Ijms 22 01088 g002
Figure 3. Phylogenetic tree of ERF-VIIs from different plant species. These plant species include model plant Arabidopsis, main crops (rice, maize, wheat, tomato) and other species with ERF-VII members that have been reported. The coding sequences of these ERF-VII genes are listed in Supplemental file 1. The tree was constructed using a neighbor-joining method with 1000 bootstraps in Mega 7 [142]. ERF-VII genes that have been identified to function in flooding-tolerance are marked in red font. Picture shapes with various colors represent ERF-VIIs from different species. Ijms 22 01088 i001, rice; Ijms 22 01088 i002, tomato; Ijms 22 01088 i003, Arabidopsis; Ijms 22 01088 i004, maize; Ijms 22 01088 i005, wheat; others picture shapes, ERF-VIIs that have been studied.
Figure 3. Phylogenetic tree of ERF-VIIs from different plant species. These plant species include model plant Arabidopsis, main crops (rice, maize, wheat, tomato) and other species with ERF-VII members that have been reported. The coding sequences of these ERF-VII genes are listed in Supplemental file 1. The tree was constructed using a neighbor-joining method with 1000 bootstraps in Mega 7 [142]. ERF-VII genes that have been identified to function in flooding-tolerance are marked in red font. Picture shapes with various colors represent ERF-VIIs from different species. Ijms 22 01088 i001, rice; Ijms 22 01088 i002, tomato; Ijms 22 01088 i003, Arabidopsis; Ijms 22 01088 i004, maize; Ijms 22 01088 i005, wheat; others picture shapes, ERF-VIIs that have been studied.
Ijms 22 01088 g003
Figure 4. The molecular mechanisms of plants with the ‘quiescent’ (LOQS) and ‘escape’ (LOES) strategies as exposed to submergence stress. Generally, flooding stress induces the accumulation of ethylene and then leads to different responses according to flooding depth. Under complete submergence, the ethylene response factor SUB1A is induced and represses the gibberellic acid (GA) signal to prevent shoot elongation through SLR1 and SLRL1 [104,153]. Ethylene accumulation induced by partial-submergence contributes to the expression of the SK1 and SK2, which are involved in the process of internode elongation via GA signal [31]. Ethylene-responsive transcription factor EIL1a can activate the expression of SD1 (a gibberellin biosynthesis gene), then increased the synthesis of gibberellins to promote internode elongation [110]. Additionally, an ‘accelerator’ ACE1 and a ‘decelerator’ DEC1 antagonistically regulate internode elongation together with gibberellic acid, expression of ACE1 or downregulation of DEC1 contributes to the internode elongation [18]. The dotted arrow represents that GA may be induced by flooding stress directly without being regulated by ethylene.
Figure 4. The molecular mechanisms of plants with the ‘quiescent’ (LOQS) and ‘escape’ (LOES) strategies as exposed to submergence stress. Generally, flooding stress induces the accumulation of ethylene and then leads to different responses according to flooding depth. Under complete submergence, the ethylene response factor SUB1A is induced and represses the gibberellic acid (GA) signal to prevent shoot elongation through SLR1 and SLRL1 [104,153]. Ethylene accumulation induced by partial-submergence contributes to the expression of the SK1 and SK2, which are involved in the process of internode elongation via GA signal [31]. Ethylene-responsive transcription factor EIL1a can activate the expression of SD1 (a gibberellin biosynthesis gene), then increased the synthesis of gibberellins to promote internode elongation [110]. Additionally, an ‘accelerator’ ACE1 and a ‘decelerator’ DEC1 antagonistically regulate internode elongation together with gibberellic acid, expression of ACE1 or downregulation of DEC1 contributes to the internode elongation [18]. The dotted arrow represents that GA may be induced by flooding stress directly without being regulated by ethylene.
Ijms 22 01088 g004
Table 1. The formation of adventitious roots (ARs), aerenchyma and ROL barriers in plants responding to flooding stress.
Table 1. The formation of adventitious roots (ARs), aerenchyma and ROL barriers in plants responding to flooding stress.
Flooding Stress (Real or Simulated)SpeciesAdventitious RootAerenchymaROL BarrierReferences
Under stagnant deoxygenated conditionsRiceY *YY[45]
Partial submergenceRiceY [39,40]
Under stagnant deoxygenated conditionsRice YY[15]
Under stagnant deoxygenated conditionsRice Y[46]
Partial submergenceRice Y [47]
In aerated, stagnant deoxygenated, or N2-flushed conditionsRice Y[48]
Under stagnant deoxygenated conditionsTeosinteYYY[49]
WaterloggingTeosinte Y [50]
Under stagnant deoxygenated conditionsTeosinteY Y[51]
WaterloggingMaize and teosinteYY [52]
WaterloggingMaize Y [53]
WaterloggingMaizeY [38]
WaterloggingMaizeY [25]
WaterloggingWheat Y [54]
WaterloggingWheat Y [55]
WaterloggingWheat YY[56]
Under stagnant conditionsHordeum marinum, Triticum aestivum, Hordeum marinum × Triticum aestivumYYY[23]
WaterloggingBarleyYY [57]
Flooding (Waterlogging)Rumex Y [58]
WaterloggingRumexY [44,59]
WaterloggingSunflowerY [60]
Waterlogging or hypoxic cultureWillowYY [61]
WaterloggingArabidopsis Y [62]
Flooding (Waterlogging)TomatoY [42]
Partial submergenceBittersweetY [63]
Half-flooding (Partial submergence)TaxodiumYY [64]
WaterloggingCucumberY [43]
WaterloggingSeashore paspalum, Bermudagrass Y [65]
WaterloggingSome Echinochloa species Y[66]
* ‘Y’ represents ‘Yes’ and means plants have the corresponding trait.
Table 2. Changes of ROS, antioxidant enzymes and other physiologic indexes in plant tissues as exposed to flooding stress.
Table 2. Changes of ROS, antioxidant enzymes and other physiologic indexes in plant tissues as exposed to flooding stress.
Plant SpeciesSamplesROSOthersAntioxidant EnzymesReference
BarleyLeaf CAT ↑ *, ascorbate peroxidase (APX) ↑ and POD ↑; SOD ↓ **[73]
BarleyRootH2O2 ↑, O2.− [72]
RiceRoot and shoot SOD ↑ and CAT ↑, APX ↓[74]
WheatLeaf, root and shoot Leaf water potential ↓, stomatal conductance ↓, photosynthesis ↓, chlorophyll content ↓, shoot nitrogen content ↓, shoot and root growth ↓, the diameter of metaxylem and protoxylem vessels of the nodal roots ↓ [54]
WheatFlag leaf POD ↑ in tolerant genotypes, SOD ↓ and CAT ↓[75]
MaizeLeaf SOD ↑, CAT ↑ and APX ↑, POD ↓[76]
CitrusLeaf and root Leaf MDA (early stage) - ***, Leaf MDA (15–27 days after waterlogging) ↑; Root MDA ↑ or -; Leaf proline ↑ or -.SOD ↑[77]
CitrusLeaf SOD ↑, CAT ↑[78]
TobaccoLeaf SOD↑ [79]
LotusShoot SOD ↓, dehydroascorbate reductase (DHAR) ↓, GR ↓, APX ↑, monodehydroascorbate reductase (MDAR) -[80]
Creeping bentgrasssRootH2O2 -MDA -SOD ↑, AR had no consistent trend with different genotypes. GR -, POD -[81]
Pigeon peaRootSuperoxide radical, H2O2Visible yellowing and senescence of leaves; leaf area, dry matter, relative water content and chlorophyll content in leaves, and membrane stability index in roots and leaves decreased.SOD ↑, CAT ↑, GR ↑ and APX ↑[82,83]
BermudagrassLeafH2O2 ↓, O2.− -Proline -POD ↑, GR ↑, CAT ↑[68]
Perennial ryegrassShoot and root Leaf chlorophyll and total carotenoid ↓, water-soluble carbohydrate (shoots and roots) ↓, MDA ↑CAT and POD (shoot) ↑, SOD, CAT, POD and APX (root) ↓[84]
SoybeanLeafH2O2 content increased at 7 days of waterlogging.Activities of lactate dehydrogenase (LDH), alanine aminotransferase (AlaAT), alcohol dehydrogenase (ADH) and pyruvate decarboxylase (PDC) changed inconsistently in different genotypes under waterlogging stress.SOD activity increased after 7 days of waterlogging, APX and CAT activities changed with no consistent trend among different genotypes or conditions.[35]
* ‘↑’ represents ‘increase’; ** ‘↓’ represents ‘decrease’; *** ‘-’ represents ‘no change’.
Table 3. General descriptions of recently reported genes and the relative phytohormones involved in plant response to flooding stress.
Table 3. General descriptions of recently reported genes and the relative phytohormones involved in plant response to flooding stress.
Flooding TypeSpeciesGenes NameFunctionPhytohormoneReferences
WaterloggingArabidopsisLSD1, EDS1 and PAD4These three genes LESION SIMULATING DISEASE1 (LSD1), ENHANCED DISEASE SUSCEPTIBILITY1 (EDS1) and PHYTOALEXIN DEFICIENT4 (PAD4) controlled the formation of lysigenous aerenchyma by regulating the generation of ethylene and ROS, and functioned in plants acclimation to waterlogging (hypoxia) stress.ethylene[62]
Under hypoxic stressArabidopsisAtLDH1AtLDH1 encodes a lactate dehydrogenase, overexpression of AtLDH1 increased the survival of roots and knockout AtLDH1 impaired the growth of roots and shoots under hypoxic stress. [98]
Under hypoxia stressArabidopsisAtRAP2.2AtRAP2.2 encodes an ERF-VII type transcription factor, overexpression of AtRAP2.2 increased the survival rate under hypoxia stress.ethylene[99]
Under anoxia conditionArabidopsisAtHRE1Overexpression of HRE1 in Arabidopsis enhanced its tolerance to hypoxic stress through regulating the expression of anaerobic genes and ethanol metabolism.ethylene[100]
SubmergenceArabidopsisAtRAP2.12AtRAP2.12 encodes an ERF-VII type transcription factor and overexpressing AtRAP2.12 in Arabidopsis showed increased survival rate under anoxia stress.ethylene[101]
Under hypoxia conditionArabidopsisAtPCO1 and AtPCO2Plant cysteine oxidase (PCO) makes the penultimate cysteine of ERF-VIIs oxidized and in turn, the expression of PCO is regulated by ERF-VIIs; overexpressing AtPCO1 and AtPCO2 in Arabidopsis became sensitive to submergence and the survival rate decreased significantly compared to wild type.ethylene[102]
SubmergenceArabidopsisAtRBOH IThe expression of AtRBOH I (respiratory burst oxidase homolog I from Arabidopsis) is induced under hypoxic stress and loss function of AtRBOH I decreased the survival rate compared to wild type under submergence condition.ethylene, auxin[103]
SubmergenceRiceSub1A (SUB1A)Sub1A encodes an ERF-type transcription factor and can enhance submergence-tolerance of rice.ethylene[104]
Partial submergenceDeepwater riceSK1 and SK2SNORKEL1 (SK1) and SNORKE2 (SK2) are ethylene response factors and regulate the elongation of internodes through gibberellin under submergence stress.gibberellin, ethylene[31]
Waterlogging and submergenceRiceCIPK15 and SnRK1ACIPK15 encodes a calcineurin B-like (CBL)-interacting protein kinase that positively regulates the expression of SnRK1A (Snf1-related protein kinase 1), and functions in rice acclimation to flooding stress by affecting sugar and energy production. [105,106]
Under anaerobic conditionRiceAmy3 subfamily genesAmy3 subfamily genes encode α-amylases and their expression can be induced in rice embryos by anaerobic conditions, which contributes rice to survive under submerged conditions. [106]
Under anaerobic conditionRiceOsTPP7OsTPP7 encodes a trehalose-6-phosphate (TP6) phosphatase and facilitates rice germination under anaerobic conditions by regulating local T6P/sucrose ratios. [107]
SubmergenceRiceOsEREBP1OsEREBP1 encodes an ERF transcription factor and may enhance the submergence tolerance through regulating jasmonate and abscisic acid signals.jasmonate, abscisic acid[108]
SubmergenceRiceLGF1LGF1 controlled the formation of leaf gas films and affected the synthesis of C30 primary alcohol, which ultimately improved submergence-tolerance of rice with increased underwater photosynthesis. [109]
SubmergenceDeepwater riceSD1SD1 (SEMIDWARF1) encodes a gibberellin biosynthesis gene, it can promote internodes elongation by increasing synthesis of gibberellins and is regulated by a transcription factor OsEIL1a.gibberellin, ethylene[110]
SubmergenceRiceOsARD1OsARD1 encodes an acireductone dioxygenase (ARD) and mediates the biosynthesis of an initial substrate methionine in the process of ethylene synthesis. Overexpression of OsARD1 increased the production of internal ethylene and submergence-tolerance of transgenic lines was improved simultaneously.ethylene[111]
Waterlogging and submergenceDeepwater riceACE1 and DEC1ACE1 confers the intercalary meristematic cells with the division ability and therefore regulated the elongation of internodes together with GA; DEC1 restricts internodes elongation.gibberellin[18]
SubmergenceMaizeSubtol6Subtol6 is a major QTL that can explain 22% of the phenotypic differences in submergence tolerance within the recombinant inbred lines. [112]
WaterloggingMaizeZmEREB180ZmEREB180 encodes an ERFV-II transcription factor from maize; overexpression of ZmEREB180 increased the survival rate under waterlogging stress through promoting ARs formation and regulating antioxidant activities.ethylene[25]
WaterloggingWheat TaERFVII.1TaERFVII.1 belongs to ERF-VII family and functions in waterlogging-tolerance of wheat, overexpression of TaERFVII.1 increased the survival rate under waterlogging stress.ethylene[17]
WaterloggingBarleyHvERF2.11The expression of HvERF2.11 can be induced by waterlogging and mediated waterlogging-tolerance of plants through improving some antioxidant and ADH enzymes activities.ethylene[113]
WaterloggingActinidia deliciosaAdPDC1AdPDC1 encodes a pyruvate decarboxylase which catalyzes the first step in ethanolic fermentation pathway and it may function in kiwifruit acclimation to waterlogging stress.abscisic acid[114]
WaterloggingActinidia deliciosaAdRAP2.3AdRAP2.3 is an ERF-VII transcription factor and may mediate waterlogging-resistance by regulating PDC and ADH genes.ethylene[115]
WaterloggingChrysanthe-mum morifoliumCmSOS1SOS1 encodes a Na+/H+ antiporter and may interact with CmRCD1 to mediate plant tolerance to waterlogging stress potentially. [116]
Waterlogging CucumberCsARN6.1CsARN6.1 encodes an AAA ATPase, transgenic lines of CsARN6.1 increased the number of ARs via enhanced ATPase activity and further affected waterlogging tolerance. [117]
WaterloggingMentha arvensisMaRAP2-4MaRAP2-4 from Mentha arvensis encodes an ERF-I type transcription factor, overexpression of MaRAP2-4 in Arabidopsis enhanced its tolerance to waterlogging and oxidative stress.ethylene, jasmonic acid[118]
WaterloggingPetuniaPhERF2PhERF2 may regulate the process of programmed cell death and alcoholic fermentation, on the base of which enhances waterlogging-tolerance of petunia.ethylene[119]
SubmergenceArabidopsisWRKY33, WRKY12WRKY33 regulates the expression of RAP2.2 together with WRKY12, in turn, RAP2.2 exhibits a feedback regulation to WRKY33. These three genes mediate submergence-tolerance of Arabidopsis collectively.ethylene[19]
SubmergenceArabidopsisANAC017ANAC017 encodes a NAC transcription factor and mediates differential submergence stress of Arabidopsis in juvenile and adult stages. [20]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jia, W.; Ma, M.; Chen, J.; Wu, S. Plant Morphological, Physiological and Anatomical Adaption to Flooding Stress and the Underlying Molecular Mechanisms. Int. J. Mol. Sci. 2021, 22, 1088. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22031088

AMA Style

Jia W, Ma M, Chen J, Wu S. Plant Morphological, Physiological and Anatomical Adaption to Flooding Stress and the Underlying Molecular Mechanisms. International Journal of Molecular Sciences. 2021; 22(3):1088. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22031088

Chicago/Turabian Style

Jia, Weitao, Maohua Ma, Jilong Chen, and Shengjun Wu. 2021. "Plant Morphological, Physiological and Anatomical Adaption to Flooding Stress and the Underlying Molecular Mechanisms" International Journal of Molecular Sciences 22, no. 3: 1088. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22031088

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