Next Article in Journal
Transcriptome Sequencing of Agave amaniensis Reveals Shoot-Related Expression Patterns of Expansin A Genes in Agave
Next Article in Special Issue
Arbuscular Mycorrhizal Fungi as an Important Factor Enabling the Adaptation of Anthyllis vulneraria L. to Zn-Pb-Polluted Tailings
Previous Article in Journal
Inoculation with Actinobacteria spp. Isolated from a Hyper-Arid Environment Enhances Tolerance to Salinity in Lettuce Plants (Lactuca sativa L.)
Previous Article in Special Issue
Impact of Polyploidy Induction for Salinity Stress Mitigation in Soybean (Glycine max L. Merrill)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Molecular Aspects and Integrated Omics Strategies for Understanding the Abiotic Stress Tolerance of Rice

1
Graduate School of Green Green-Bio Science and Crop Biotech Institute, Kyung Hee University, Yongin 17104, Republic of Korea
2
Research Center for Plant Plasticity, Kyung Hee University, Yongin 17104, Republic of Korea
*
Author to whom correspondence should be addressed.
Plants 2023, 12(10), 2019; https://doi.org/10.3390/plants12102019
Submission received: 24 April 2023 / Revised: 11 May 2023 / Accepted: 17 May 2023 / Published: 18 May 2023
(This article belongs to the Special Issue Plant Adaptation to Abiotic Stresses)

Abstract

:
Rice is an important staple food crop for over half of the world’s population. However, abiotic stresses seriously threaten rice yield improvement and sustainable production. Breeding and planting rice varieties with high environmental stress tolerance are the most cost-effective, safe, healthy, and environmentally friendly strategies. In-depth research on the molecular mechanism of rice plants in response to different stresses can provide an important theoretical basis for breeding rice varieties with higher stress resistance. This review presents the molecular mechanisms and the effects of various abiotic stresses on rice growth and development and explains the signal perception mode and transduction pathways. Meanwhile, the regulatory mechanisms of critical transcription factors in regulating gene expression and important downstream factors in coordinating stress tolerance are outlined. Finally, the utilization of omics approaches to retrieve hub genes and an outlook on future research are prospected, focusing on the regulatory mechanisms of multi-signaling network modules and sustainable rice production.

1. Introduction

Rice (Oryza sativa L.) is a major food crop and staple food source for half of the world’s population [1]. According to the report published by the Department of Economic and Social Affairs of the United Nations (https://www.un.org/en/desa, accessed on 3 March 2023), the world population is projected to reach almost 10 billion in 2050 and 11.2 billion in 2100. According to the Global Agricultural Productivity Index 2022 (http://www.globalagriculturalproductivity.org/, accessed on 3 March 2023), global agricultural productivity has fallen below the level needed for the sustainable growth of agricultural output. The average annual Total Factor Productivity growth rate declined from 1.99% in 2001–2010 to 1.12% in 2011–2020. As per the world’s population growth rate, rice production will need to increase by 160 million tons from the current yield (104 million tons) [2,3]. In various aspects, rice significantly promotes social stability, economic development, and food security. By providing a staple food source, rice can help stabilize communities and societies that depend on it for nutrition. Secondly, the cultivation and production of rice can provide employment opportunities for people within the agricultural sector, thereby promoting economic development and reducing poverty. For 70% of the low-income people in Asia, rice is not just a food crop but a way of life, as it is often their primary source of income. We will need about twice as much total food production in the predicted scenario to feed our fast-growing population. In past decades, breakthroughs in rice hybridization and dwarf breeding substantially enhanced grain yield and provided a roadmap for future breeding programs [4]. Climate change, including frequent extreme temperatures, rising carbon dioxide and ozone levels, and uneven rainfall, aggravates the degree of drought or salinization of agricultural lands and changes the growing environment of crops. Despite the significant increase in rice yield, plant breeders face challenges amid the fast-growing world population and the effects of climate change. The acceleration of global climate change has led to an increased incidence of abiotic stress factors, leading to a loss of yields in approximately 50% of rice-growing land [5,6]. In Asia alone, around 42 million hectares of rice-growing areas are affected by drought stress, while globally, 45 million hectares of irrigated land and 32 million hectares of rain-fed dryland are impaired by salinity stress [6]. Additionally, flooding affects over 35% of global rice-growing land [7], and rice productivity is notably impacted in rain-fed areas (which represent more than 50% of total rice-growing land) due to the prevalence of floods, water deficits, and salt stresses [8,9]. As per estimation, countries (China, India, the United States, and Indonesia) where more than 70% of the global population resides may experience increased flood risk due to changing seawater levels and extreme weather events [10]. For instance, a four-degree increase in temperature could lead to a decline in crop yields of 15–35% in Africa and Asia, with estimates of a 25–35% yield reduction in the Middle East [11]. It is reported that elevated salinity levels affect 7% of the world’s land area, leading to an annual economic loss of USD 27.2 billion due to adverse environmental impacts [12]. In addition, it is projected that salinity stress will impact over 30% of cultivable lands globally by 2050 [13], with the potential for further increases due to changes in seawater levels. Therefore, cloning important genes related to abiotic stress tolerance, analyzing the molecular mechanisms of stress responses, and further mining their breeding application value hold important theoretical and practical significance for rice breeding and food security.
The fine-mapping and cloning of several genes that are responsible for tolerance to various abiotic stresses has laid the foundation for yield improvement and enriched the genetic resources of rice [14,15]. Additionally, integrative omics approaches have efficiently determined the genetic variability of the available germplasm resources and provided the basis for rice breeding. The regulatory genes that are responsible for important agronomic traits have been reported through the efficient utilization of genome-wide association studies (GWAS) and integrative omics approaches, which is significant in rice breeding [16,17]. Mainly, the massive utilization of multi-omics methods has provided innovative insights into several biological topics and has been utilized for the genetic improvements to rice [18,19]. Their use is particularly vital for reducing agricultural costs and providing environmental protection.
In short, addressing food security is an endless and challenging task. Here, we review the key achievements made in rice genetics and breeding for environmental stress tolerance in past years, which will help provide systematic and comprehensive information and theoretical support for developing rice with improved yield and tolerance to multiple stresses.

2. Recognition and Signaling of Abiotic Stress

Abiotic stresses severely harm agricultural production and cause extreme deterioration of the ecological environment. When plants are stressed, molecular and morphological changes occur, directly or indirectly affecting plant growth and development. Abiotic stress decreases the final yield by severely affecting the physiology and morphology of crops at the vegetative and reproductive stages. Adversities, such as extreme temperatures, drought, and salinity, are often interconnected. In the past decade, abiotic stresses have occurred more commonly because of global climate change [20,21]. These stresses hamper water uptake and nutrient absorption and disturb plant growth and development, significantly decreasing the germination percentage, growth rate, leaf size, and productive tillers [22]. At the reproductive stage, rice displays more susceptibility to extreme temperatures, and both low and high temperatures lead to spikelet sterility and decreased grain filling [23,24]. Since plants are sessile organisms, they have evolved complex mechanisms to combat stresses by activating cellular machineries, stress avoidance, and recovery from stress [25]. After a stress attack, plants perceive the signals, transduce them to various cellular compartments, and establish a response mechanism to alleviate the effects of episodic stress events.
Plant responses to abiotic stress generally begin with signal perception. After sensing the abiotic stress, the plasma membrane produces secondary signaling molecules, such as reactive oxygen species (ROS) and inositol phospholipids. The secondary messenger stimulates the intracellular membrane by regulating the intracellular Ca2+ level, which then initiates a protein phosphorylation cascade to generate phosphorylated proteins that are directly participating in cytoprotection or regulating the transcription of specific stress-responsive target genes [26]. Some gene products are involved in the production of regulatory factors, such as abscisic acid (ABA) and ethylene, which activate the expression of transcription factors (TFs) [27]. TFs bind to target gene-promoter sequences, further activating or inhibiting downstream functional gene expression, and ultimately play a regulatory role in abiotic stress responses [28]. Revealing abiotic stress response mechanisms and cultivating new varieties of stress-resistant crops will help achieve sustainable agricultural development and ensure food security for the growing world population. Here, we discuss the transcriptional networks that mediate the rice plant’s response to different abiotic stresses, with a particular emphasis on the candidate genes and TFs that are involved in abiotic stress responses and play a critical role in conferring stress tolerance. A list of important candidate genes with their genomic locations is given in Table 1, and their functions are discussed in subsequent sections.

2.1. Low Temperature

In rice, a low temperature below 15 °C directly affects the germination percentage, seedling vigor, tillering, reproduction, and grain maturity [29]. Moreover, many physiological, biochemical, and molecular changes occur during cold acclimation, including the activation of antioxidant systems, the synthesis and accumulation of cryoprotectants, and mechanisms protecting and stabilizing cell membranes [30]. To maintain cell membrane stability, the level of unsaturated phospholipids in the membrane composition increases, and cells accumulate sucrose- and proline-rich osmotic molecules and antifreeze proteins, which trap water molecules by creating hydrogen bonds. Cold stress changes the cell membrane’s fluidity and affects the structure and activity of membrane-localized proteins, triggering the Ca2+ influx, which is essential for inducing the expression of temperature-responsive genes [31]. Studies have confirmed that the rice cold sensor (COLD1) binds to the rice G protein α subunit 1 (RGA1) to mediate cold sensing and low-temperature-induced extracellular Ca2+ influx in rice [14] (Figure 1A). In addition to COLD1, the rice CBL-interacting protein kinase 7 (OsCIPK7) is thought to sense low-temperature signals by regulating the conformation of its kinase domain and Ca2+ influx [32]. However, the mechanism by which COLD1 and OsCIPK7 regulate Ca2+ influx under low-temperature stress remains unclear. Over the past decades, many chilling stress-related genes have been identified in rice using different genetic approaches; however, few have been cloned and characterized [33].
Chlorophyll content and fluorescence are important indicators for evaluating plant stress tolerance [34]. In rice, low temperatures reduce chlorophyll synthesis and chloroplast formation, confirming that chlorophyll-related changes are essential indicators of rice tolerance to low temperatures [35]. Studies have demonstrated that overexpressing OsiSAP8 in rice can significantly improve the plant chlorophyll content and cold tolerance [36]. In rice, the cell membrane is the first to perceive low temperatures and chilling damage, and its physicochemical properties are sensitive to low temperatures, resulting in intracellular electrolyte leakage. Therefore, the electrolyte leakage rate is often an important indicator of whether plants can tolerate low temperatures [37]. It was revealed that the overexpression of OsNAC5 results in decreased electrolyte leakage, thus indicating tolerance to low temperatures [38].
Under cold stress in rice, the intracellular oxygen metabolism is imbalanced, and ROS are generated, triggering membrane lipid peroxidation and resulting in cell membrane system damage [39]. ROS signaling can activate stress-responsive genes and downstream signaling pathways that help plants cope with cold stress, such as antioxidant defense mechanisms, osmolyte accumulation, and membrane remodeling. ROS also promote polyunsaturated fatty acid degradation and malondialdehyde (MDA) production, further damaging plant tissues and cells [40]. The protective mechanism of rice against oxidative stress has two major systems: the enzymatic and the non-enzymatic systems. The enzymatic system includes various antioxidant enzymes catalyzing ROS-scavenging reactions. Among them, superoxide dismutase (SOD) and catalase (CAT) are the two most effective antioxidant enzymes, converting superoxide anion and hydrogen peroxide (H2O2) into water and oxygen molecules, thus reducing ROS damage to cells [41]. Non-enzymatic systems also include various antioxidants, among which reduced glutathione (GSH) and ascorbic acid are the most essential [42,43]. In rice, overexpressing the ascorbate peroxidase gene OsAPXa can increase ascorbate peroxidase activity at low temperatures and reduce lipid peroxidation and MDA levels, improving the seed setting rate of rice at low temperatures [44].
Under cold stress, rice accumulates large amounts of soluble sugars, such as sucrose, hexose, raffinose, glucose, fructose, and trehalose. Soluble sugars can act as osmotic regulators of cells under low temperatures, stabilizing cell membranes and protoplasmic colloids. In addition, sugars also provide a carbon skeleton and energy for synthesizing other organic substances [45,46]. Overexpressing the trehalose synthesis genes OsTPP1, OsTPP2, and OsTPS1 can significantly improve rice tolerance to low temperatures [47].
In rice, when the pollen mother cell meiosis encounters low temperatures, sugars—including sucrose, glucose, and fructose—accumulate in the anthers, and simultaneously, the activity of sucrose-degrading enzymes decreases, and the expression of monosaccharide transporters is downregulated. This results in an insufficient supply of sucrose to the tapetum and pollen grains, causing pollen sterility [48]. A previous study observed that externally applying sucrose could significantly improve the fertility of rice pollen under low temperatures and increase the seed setting rate [49]. Rice also accumulates a large amount of proline at low temperatures. Proline is widely involved in osmotic regulation and carbon and nitrogen metabolism and protects most enzymes from denaturation and inactivation [50]. At the same time, proline also stabilizes polyribosomes and maintains protein synthesis [51]. Under adverse conditions, proline can also remove the excess hydrogen ions (H+) produced by the stress reaction and maintain the optimal pH for aerobic respiration in the cytoplasm [52]. In addition, proline binds to proteins through its hydrophobic group to improve the hydrophilicity of proteins [53]. In rice, the overexpression of OsCOIN can lead to a significantly increased proline content and enhanced low-temperature tolerance [54].
ABA is essential in low-temperature adversity [55]. Under cold stress, maintaining a relatively low level of ABA is beneficial for improving the stress tolerance of rice. Overexpressing the rice ABA metabolism gene OsABA8ox1 reduced ABA levels in rice seedlings and improved the tolerance to low temperatures [56]. The ABA signaling pathway comprises the ABA receptors PYR/PYL/RCAR (pyrabactin resistance/pyrabactin resistance-like/regulatory component of the abscisic acid receptor), negative regulator type 2C protein phosphatase (PP2C), positive regulator four core components, including SNF1-related protein kinase 2 (SNF1-related protein kinase 2, SnRK2), and the TFs AREB/ABF that form a dual negative regulatory system [57,58,59,60].
Under low-temperature conditions, endogenous ABA increases and binds to PYR/PYL/RCAR. Consequently, PYR/PYL/RCAR interacts with PP2C, thus inhibiting its binding to SnRK2. SnRK2 can phosphorylate TFs and activate ABA-responsive gene expressions, improving plant tolerance to low temperatures. However, under normal conditions, the endogenous ABA content is unchanged, and the interaction between PP2C and SnRK2 prevents the latter from phosphorylating downstream substrates, repressing the expression of ABA-responsive genes [61]. The ABA receptor OsPYL9 positively regulates ABA signaling, and its overexpression can significantly improve cold tolerance in rice [62]. In addition to the core component PYL-PP2C-SnRK2-ABF, the components of the ABA signaling pathway include Ca2+, ROS, nitric oxide (NO), phospholipid molecules, and other kinases, such as mitogen-activated protein kinases (MAPK) [26].
The C-repeat-binding factor/dehydration-responsive element-binding factor (CBF/DREB1) is an essential TF part of ABA-independent low-temperature response signaling pathways. CBF belongs to a subfamily of the AP2/ERF (APETALA 2/ethylene responsive) TF family. The AP2/ERF family is divided into four subfamilies: AP2, ERF, DREB, and RAV (related to VP1/ABI3) [63]. CBF contains a conserved AP2 domain, which can bind to the promoter region of low-temperature-responsive genes (CORs) containing the core element CCGAC (also known as CRT, C-repeat) under low temperatures and activate the transcription of CORs. The CBF gene is usually regulated by the bHLH-like TF (ICE1). Therefore, the low-temperature regulation pathway is also called the ICE-CBF-COR pathway [64]. Previous studies have demonstrated that overexpressing OsDREB1D and OsDREB1F of the DREB1 subfamily in rice can improve the tolerance of rice to low temperatures [65].
The mitogen-activated protein kinase OsMAPK3 phosphorylates OsbHLH002/OsICE1, thus reducing its ubiquitination level and promoting the accumulation of active OsbHLH002. Later on, the expression of the downstream gene OsTPP1 is activated by OsbHLH002, which finally increases the content of trehalose and improves rice tolerance to low temperatures [66]. In addition, some CBF genes are induced by ABA, such as OsDREB1F. OsDREB1F is involved in both ABA-independent and -dependent signaling pathways [67], suggesting that the ABA-dependent signaling pathway overlaps with the ICE-CBF-COR pathway [68]. Under cold stress, at the booting stage, the endogenous gibberellins (GA) content in rice decreases. A previous study demonstrated that mutants of sd1 and d35 were sensitive to low temperatures and that external GA treatment could improve the pollen fertility of mutants at low temperatures [49]. The recently cloned HAN1 is crucial for rice cold tolerance at the seedling stage. HAN1 encodes an oxidase that fine-tunes the JA-mediated chilling response by catalyzing the conversion of biologically active jasmonoyl-L-isoleucine (JA-Ile) to the inactive form 12-hydroxy-JA-Ile (12OH-JA-Ile). Functional nucleotide polymorphism in the HAN1 promoter increases putative MYB cis-elements in the allele of temperate japonica rice, enhancing its cold tolerance [69,70].
Several other genes that are involved in cold stress tolerance have been cloned and functionally identified in rice. qLTG3-1 encodes a conserved glycine-rich (GRP) domain, and the three sequence variations in the coding region determine the strength or weakness of seed germination under low temperatures in rice [71]. In a study, it was observed that a single SNP in OsGSTZ2 was responsible for amino-acid differences and was essential for improving low-temperature tolerance in rice at the seedling stage [72]. LTG1 encodes casein kinase I, and the amino acid substitution at position 357 (I357K) in the coding region has important effects on the growth rate, heading date, and rice yield under low temperatures [73]. Ctb1 encodes a protein containing an F-box domain, which interacts with the E3 ubiquitin ligase subunit Skp1 and participates in low-temperature signaling in the ubiquitin–proteasome pathway [74]. CTB4a encodes a conserved leucine-rich repeat receptor-like kinase LRR-RLK (leucine-rich repeat receptor-like kinase), which interacts with the β subunit AtpB of ATP synthase and affects its activity. The three SNPs (−2536, −2511, and −1930 upstream of ATG) in the CTB4a promoter region determine the tolerance of different rice cultivars at low temperatures [33]. OsbZIP73 positively regulates low-temperature tolerance at the seedling stage, and the interaction between OsbZIP71 and OsbZIP73 modulates the ROS and ABA levels in response to cold stress [75]. In rice, NUS1 is mainly expressed in immature leaves and upregulated under cold stress, and nus1 mutants display impaired chloroplast rRNA accumulation and repressed transcription/translation capacity [76]. Other essential genes, such as TEMPERATURE-SENSITIVE VIRESCENT (TSV), WHITE STRIPE LEAF 5 (WSL5), and OsCYP20-2, can protect rice from chilling stress by regulating different chloroplastic and photosynthetic genes [77,78,79]. Significant progress has been made for cold stress tolerance, but our knowledge on plant responses at the single-cell level remains scarce.
Table 1. List of key genes involved in abiotic stress tolerance of rice.
Table 1. List of key genes involved in abiotic stress tolerance of rice.
Gene SymbolGene NameLocus IDPosition 1Position 2References
Cold stress
OsCOLD1Chilling tolerance divergence 1LOC_Os04g5118030311519–3031630330311574–30316221[14]
OsCIPK7CBL-interacting protein kinase 7LOC_Os03g4344024226224–2422793024226372–24227930[32]
OsiSAP8Stress associated protein gene 8LOC_Os06g4101024491979–2449423824491993–24493907[36]
OsNAC5NAC domain transcription factor 5LOC_Os11g082104299149–43017834299277–4301784[38]
OsAPXaAscorbate peroxidase 1LOC_Os03g176909843327–98467479843336–9846670[44]
OsTPP1Trehalose-6-phosphate phosphatase 1LOC_Os02g4423026767603–2677163326767607–2677162[47]
OsCOINCold-inducibleLOC_Os01g01420209771–214229209771–214173[54]
OsPYL9Pyrabactin resistance-like 9LOC_Os06g3667021556404–2155733421556404–21557283[62]
OsDREB1DDehydration responsive element-binding protein 1DLOC_Os06g069703310866–33118223310919–3311822 [65]
OsHAN1Salt overly sensitive 1LOC_Os12g4436027494401–2750885127495775–27508468 [69]
Heat stress
OsHSP26.7Heat shock protein 26LOC_Os03g141807697015–76982847697015–7698027 [80]
OsHSP17.7Small heat shock protein 17.7LOC_Os03g160408838031–88385108837821–8838527 [81]
OsTT1Thermo-tolerance 1LOC_Os03g2697015420148–1542472415420151–15424562 [82]
OsGIRL1Gamma-ray-induced LRR-RLK1LOC_Os02g124406487390–64905776488457–6490577 [83]
OsNSUN2NOP2/SUN domain family member 2LOC_Os09g2963018013221–1802062818013221–18020611 [84]
OsTCM5Thermo-sensitive chlorophyll-deficient mutant 5LOC_Os05g3446020433366–2043793220433366–20437932 [85]
OsEG1Extra glume 1LOC_Os01g6743039177169–3917867639177169–39178676 [86]
OsRBG1Rice big grain 1LOC_Os11g3043017694857–1769604217694769–17696076 [87]
OsANN1Annexin 1LOC_Os02g5175031698161–3170050331698204–31700438 [88]
Drought stress
OsDRO1Deeper rooting 1LOC_Os09g2684016307780–1631083716307780–16310837 [89]
OsASR5Abscisic acid stress and ripening 5LOC_Os11g067203278435–32794253278451–3279419 [90]
OsDSTDrought and salt toleranceLOC_Os03g5724032645456–3264705132645695–32646908 [91]
OsJAZ1Jasmonate ZIM-domain protein 1LOC_Os04g5592033306461–3331023233306468–33310169 [92]
OsEPF1Epidermal patterning factor 1LOC_Os04g5449032414780–3241561332414780–32415613 [93]
OsCPK9Calcium-dependent protein kinase 9LOC_Os03g4827027467403–2747275927467413–27472746 [94]
OsITPK23,4-trisphosphate 5/6-kinase 2LOC_Os03g128406901924–69074096902118–6907409 [95]
Salt stress
OsSAPK4Stress/ABA-activated protein kinase 4LOC_Os01g6497037710241–3771529637710241–37714835 [96]
OsMAPK44Mitogen-activated protein kinase 44LOC_Os05g4914028188762–2819402528188894–28194022 [97]
OsLOL5LSD1-like-5LOC_Os01g4271024292537–2429969724294290–24299500 [98]
OsBADH1Betaine aldehyde dehydrogenase 1LOC_Os04g3902023171426–2317636923171516–23176332 [99]
OsKAT1Shaker potassium channel 1LOC_Os01g5520031761223–3176388731761223–31763887 [100]
OsHAK5High-affinity potassium (K+) transporter 5LOC_Os01g7049040825681–4083030140825678–40830191 [101]
OsVP1Viviparous 1LOC_Os01g6837039723155–3972698839723171–39726984 [102]
Osmotic stress
OsP5CS1Pyrroline-5-carboxylate synthetase 5LOC_Os05g3815022374029–2238103922374029–22380820 [103]
OsPPa6Inorganic pyrophosphatase 6LOC_Os02g5294032374870–3237854632374870–32378165 [104]
OsCCD1Carotenoid-cleavage dioxygenase 1LOC_Os12g4431027464735–2747203627464832–27471667 [105]
OsANN10Annexin 10LOC_Os09g2799016999259–1700137416999461–17001374 [106]
OsCSLD4Curled leaf and dwarf 1LOC_Os12g3689022602880–2260731522602934–22607307 [107]
OsPP65Protein phosphatase 65LOC_Os04g3766022389303–2239383122389359–22394048 [108]
OSCA1Osmolality-sensing ion channel 1LOC_Os01g4527425692717–2570509025696671–25705077 [109]
Submergence stress
OsSUB1BSubmergence 1BLOC_Os09g114806404474–64060396404482–6406039 [110]
OsSUB1CSubmergence 1CLOC_Os09g114606387891–63897896387891–6389789 [110]
OsEIL1Ethylene insensitive3-like gene 1ALOC_Os03g2079011776086–1177800811774230–11778954 [111]
OsACE1Accelerator of internode elongation 1LOC_Os03g2251012929937–1293079712929937–12930797 [112]
The position 1 and position 2 correspond to Rice Genome Annotation Protect (http://rice.uga.edu/ (accessed on 4 April 2023)) and RAP-DB (https://rapdb.dna.affrc.go.jp/index.html (accessed on 4 April 2023)) genome browsers, respectively.

2.2. Heat Stress

High temperatures above 35 °C can negatively impact plant growth, pollen viability, fertilization ability, and grain filling [113]. At the same time, high temperatures inhibit photosynthesis and reduce water content, negatively impacting plant cell division and growth [114]. At present, numerous high-temperature-related genes have been cloned in rice. They can be roughly divided into different categories, including heat shock proteins (HSPs), heat shock TFs (HSTFs), stress-related TFs, and others, and they play a role in temperature sensing and stress responses (Figure 1B).
HSPs have a wide variety and rich content in plants and are major molecular chaperone proteins. HSPs help with the correct folding of proteins and assist their transmembrane transfer under plant stress, thus enhancing the plants’ stress tolerance. About 17 proteins of this type have been reported in rice, and most are reportedly induced by heat shock only. Their specific biological functions are not clear yet [115,116]. Among them, OsHSP26.7 encodes a chloroplastic HSP, which protects chloroplasts from oxidative damage caused by high temperatures and ultraviolet light [80]. OsHSP17.7 was isolated from heat-stressed rice, and HSA32 and HSP101 were proved to interact to form a positive feedback regulation loop with a post-transcriptional role [81].
HSTFs form the regulatory hubs of the heat stress response, responsible for the signal transduction and activation of heat shock protein expression. Among the ~25 HSTFs in rice, ~13 are induced by heat shock; however, research on their functions is still lacking, and the molecular mechanism leading to their expression remains to be revealed [117]. OsHsfA2d encodes two splice isomers, which function under high-temperature stress and help cells establish a protein folding balance [118]. OsHsfA2e is localized in the nucleus and displays C-terminal transcriptional activity. Its higher expression can significantly enhance Arabidopsis environmental stress tolerance [119]. OsHsfA2a has multiple transcripts that are essential for rice growth and stress responses [120].
In addition to HSTFs, other TFs containing stress elements are involved in high-temperature stress. OsDREB1B is an AP2/EREBP TF whose expression is altered by temperature variations [121]. OsWRKY11 encodes a TF containing a WRKY domain, and under the HSP101 promoter, it can significantly improve rice heat and drought tolerance [122]. OsAREB1 and OsbZIP60 encode a bZIP TF, and their expression is affected by high temperatures. The ethylene response factor (HYR) is a key regulator of the direct activation of photosynthesis and can also regulate downstream carbon metabolism genes and affect morphology–physiology under drought and high-temperature stress, stabilizing the rice yield [123]. SNAC3 is an NAC TF that can significantly enhance high-temperature tolerance by mediating the metabolism of reactive oxygen species (ROS) [124]. In rice, the OsMYB55 TF is induced by high temperatures, and its overexpression can significantly improve high-temperature tolerance and grain yield [125].
Additionally, several enzymes are involved in temperature stress and other biological pathways in rice. GAD3 and OsGSr encode a glutamate decarboxylase and a glutamine synthase, respectively, which are highly expressed after being activated by OsMYB55 to promote the synthesis of stress-related amino acids and have an essential contribution to the high-temperature tolerance of rice [125,126]. OsHTAS encodes a ubiquitin ligase located in the nucleus and cytoplasm. It mediates the hydrogen peroxide-induced stomatal opening and closing and has a positive regulatory effect on heat tolerance in rice [127]. OsTT1 was identified in African rice, which has a leucine-rich repeat receptor encoding a 26S proteasome α2 subunit. It is involved in the ubiquitination and degradation of toxic proteins and is essential to the high-temperature response of rice [82]. OsGIRL1 encodes a leucine-rich repeat receptor protein kinase, and its overexpression confers an enhanced seedling survival rate under heat stress [128]. TOGR1 is a DEAD-box RNA helicase involved in high-temperature responses and the processing of rRNA precursors, which is essential for cell proliferation at high temperatures [83]. AET1 is a histidine tRNA guanylyltransferase regulating the translation process of ribosomal proteins and is essential for maintaining protein translation stability at high temperatures [129]. OsNSUN2 encodes an mRNA m5C methyltransferase, which maintains the normal growth and development of rice under high temperatures by regulating mRNA translation efficiency [84]. PGL encodes chlorophyll a oxidase 1, mainly expressed in green tissues, whose mutants are sensitive to heat stress [130]. TCM5 is a chloroplast-targeted Deg protease involved in chloroplast development under high temperatures and is essential for the functional maintenance of the photosynthetic system domain [85].
EG1 is a high-temperature-mediated mitochondrial lipase that can trigger the expression of floral organ genes under high temperatures and maintain the stability of floral organ development [86]. OsAPX2 is a rice cytoplasmic ascorbate peroxidase gene involved in heat-induced ascorbate peroxidation in rice [131]. OsGSK1 is a homologous protein of BIN2, a key regulator of the Arabidopsis BR signaling pathway, and its knockout mutants display sensitivity to heat stress [132]. GSA1 encodes a rice glycosyltransferase UGT83A1 with a broad-spectrum of glycosyltransferase activity and regulates the synthesis of flavonoid metabolites, rice yield, and stress resistance [133].
The accumulation of soluble sugars such as glucose and fructose, and non-soluble sugars such as starch, is an important mechanism for plants to cope with heat stress [134]. Studies have shown that a high accumulation of sugars can inhibit photosynthesis [135], possibly due to the feedback regulation of sugar accumulation in the Calvin cycle and Rubisco activity [136]. To overcome the inhibitory effect of high sugar accumulation on photosynthesis, several strategies have been proposed. For example, it has been suggested that the application of exogenous sugar-metabolizing enzymes, such as invertases and sucrose synthases, can enhance photosynthesis under heat stress by reducing the sugar accumulation [137]. In addition, the manipulation of sugar transport and allocation pathways, such as the overexpression of sugar transporters or starch synthesis enzymes, has also been proposed to reduce sugar accumulation and maintain photosynthesis under heat stress [138]. In rice, some key genes have been identified that play important roles in sugar accumulation and photosynthesis under heat stress. For example, the expression of a sucrose transporter OsSUT1 is upregulated under heat stress, leading to an increased sugar accumulation and decreased photosynthesis [139]. Two heat-responsive genes, ONAC127 and ONAC129, which encode NAM/ATAF/CUC (NAC) domain TFs were also found to be involved in the grain-filling process under heat stress. It was revealed that ONAC127 and ONAC129 regulate sugar transportation and abiotic stress responses, which are crucial for proper grain filling. The target genes of ONAC127 and ONAC129 in developing rice seeds include the sugar transporter gene OsSWEET4 and monosaccharide transporter gene OsMST6 [140]. Other members of few gene families like trehalose-6-phosphate phosphatase and hexokinases might also be potential targets to study sugar metabolism and photosynthesis under heat stress [47,141]. The functional analysis of these gene may help in understanding the regulation of sugar accumulation and the Calvin cycle in rice, and their manipulation may offer potential means to overcome the negative impact of sugar accumulation on photosynthesis under heat stress.
In addition to the abovementioned gene categories, others related to high temperatures have been reported. FLO2 is a protein containing a 34-peptide repeat sequence that controls the amylose content of rice seeds. The overexpression of FLO2 causes grain enlargement and displays significant expression differences among different rice varieties under high-temperature stress [142]. Fie1 is a fertilization-independent endosperm-autonomous gene that regulates seed size and the nighttime high-temperature sensitivity of seeds during grain filling, which is of great significance for maintaining yield under high temperatures [143]. RBG1, a novel positive regulator of grain size, was reported to enhance tolerance to heat, osmosis, and salt stress in rice through auxin and cytokinin pathways [87]. OsLEA4 and OsLEA5 encode two late embryogenesis proteins, and their expression in Escherichia coli led to significant high-temperature tolerance, suggesting they may contribute to stress tolerance in crop species [144]. ZFP177 encodes an A20/AN1-type zinc finger protein induced by heat stress, and its overexpression in tobacco can improve heat stress tolerance [145]. OsRZFP34 encodes a ring zinc finger protein induced by heat, and its loss of function leads to a smaller stomatal diameter and reduced temperature tolerance [146]. OsANN1 is a calcium-binding protein with ATPase activity, and its knockout mutants displayed sensitivity to heat stress, resulting in severe redox imbalance [88]. OsCNGC14 and OsCNGC16 are two cyclic nucleotide-gated ion channel proteins in the plasma membrane responsible for regulating the influx of cytoplasmic calcium ions; their mutants are sensitive to temperature stress [147]. SLG1 encodes a conserved cytoplasmic tRNA 2-thiolated protein 2 (RCTU2) involved in tRNA modification and is essential in high-temperature resistance in rice seedlings and at reproductive development stages in rice [148]. OsGR-RBP4 is a glycine-rich RNA-binding protein screened from the rice cDNA library under heat shock and constitutively induced by heat stress [149].
Transcriptional modifications are crucial for environmental temperature perception and heat stress responses [150]. Recently, four genes related to post-transcriptional regulation under heat stress have been reported in rice, including RNA helicase TOGR1, tRNA His guanylyltransferase AET1, RNA methyltransferase OsNSUN2, and cytoplasmic tRNA 2-thiolated protein 2 SLG1 [83,84,129,148]. Among them, TOGR1 is a chaperone protein of the small subunit complex in the nucleolus and is involved in processing rRNA precursors at high temperatures [83]. AET1 plays a decisive role in the maturation of the precursor tRNA His on the endoplasmic reticulum and ribosomes and in later protein translation [129]. OsNSUN2 mediates the mRNA methylation modifications of photosynthesis-related detoxification proteins in the nucleus under high temperatures and improves the translation efficiency to ensure the normal growth of rice [84].
In higher plants, the connection between ubiquitin molecules and substrates can mediate various cellular functions through the Ub/26S proteasome system, which is also crucial for plant heat tolerance [151]. Three genes involved in this pathway have been reported in rice—OsHCI1, HTAS, and OsTT1 [82,126,127]. Among them, HTAS and OsHCI1 encode E3 ubiquitin ligases. HTAS has ubiquitin ligase activity in vitro and can interact with ascorbate peroxidase to mediate stomatal opening and closing under high temperatures. OsHCI1 accumulates in large amounts in the nucleus at high temperatures and mediates the rapid export of ubiquitination and degradation of substrate proteins along the cytoskeleton [82,126,127]. OsTT1, as a 26S proteasome α2 subunit, can help degrade toxic proteins more efficiently to maintain a high-temperature response [82].
In plants, the accumulation of some metabolites significantly contributes to high-temperature tolerance. In rice, the MYB55 TF regulates the expression of downstream glutamate dehydrogenase GAD3 and glutamine synthase OsGS1.2 and promotes the accumulation of stress-related amino acids, such as L-glutamic acid and GABA, under high temperatures [125]. In addition, the UDP glucosyltransferase gene GSA1 can promote flavonoid and anthocyanin synthesis under heat, drought, and salt stress and enhance the stress resistance of rice [133]. The lipase gene EG1 is located in mitochondria and plastids and mediates the mitochondrial lipase pathway under high temperatures, regulating lipid metabolism and downstream gene expression [86].
Calcium ions are essential in plant responses to abiotic stress, acting as second messengers; however, the molecular mechanism underlying the upregulation of intracellular calcium ion concentration as a key stress signal remains to be elucidated [152]. In rice, CNGC14 and CNGC16 act as cyclic nucleotide-gated ion channels to regulate calcium influx from the plasma membrane under high temperatures, while OsANN1 regulates the redox balance by increasing the activity of ROS-scavenging enzymes [88,147].
In rice, APX2 and SNAC3 are directly related to ROS scavenging [124,131]. The ascorbate peroxidase gene APX2 is synthesized in large amounts in the cytoplasm at high temperatures, enhancing the ability of rice to scavenge hydrogen peroxide, protecting spikelets from lipid peroxidation, and maintaining fertility [131]. The TF SNAC3 positively regulates the expression of ROS-scavenging genes in the nucleus and enhances high-temperature tolerance in rice [124].
Hormonal signaling is also essential under high-temperature stress, and various hormonal signals interact in rice. OsGSK1 acts as a negative regulator of BR signaling to mediate the response to heat stress in rice [132]. Another plasma membrane receptor kinase, OsGIRL1, regulates heat stress by responding to ABA or SA signaling [128]. The heat stress-related TF OsAREB1 is a positive regulator of ABA signaling. Finally, OsRZFP34 is involved in ABA-mediated stomatal opening and closing under heat stress [146].
Several high-temperature-related genes have been reported, and the signaling pathways they are involved in have gradually become clear. However, the molecular mechanism and regulatory network of high-temperature sensing, signaling, and transduction to downstream elements remain poorly known and must be the main topic for future research.

2.3. Drought Stress

Drought is an important abiotic factor limiting rice growth and production. The agricultural drought threat’s frequency and range are increasing worldwide, and 50% of global rice production is severely affected by it. Breeding drought-resistant rice varieties is a practical option to fight against drought stress [153]. However, drought tolerance in rice is a complex trait controlled by multiple genes and significantly interacting with the environment, which makes the drought tolerance mechanism in rice more complicated. Drought tolerance-related genes are divided into three major categories: transcriptional control, stress signaling, and membrane transport [154,155]. These genes mainly influence the molecular, physiological, and biochemical mechanisms after plants are subjected to drought stress [156]. Many work in ABA-dependent and -independent regulatory systems [157].
The root system is responsible for absorbing and translocating nutrients and water and is essential for crop drought-escaping strategies. Drought-tolerant rice varieties have a well-developed root system. By increasing the root-shoot ratio and enhancing root penetration, the plant can maintain a higher water potential under drought conditions, efficiently absorbing water in the soil and forming a stable internal environment for the normal growth of plants [155]. Under drought stress, the root system enhances the overall drought resistance of rice by improving cuticle resistance and increasing the number of root hairs, density, and depth [158]. Many genes involved in root traits in rice are also involved in drought tolerance. A major QTL Deeper rooting 1 (DRO1) controls root tip cell elongation, asymmetric growth, and gravity-based downward bending of the root tip (Figure 1C). The transformation of shallow-rooted rice varieties with DRO1 resulted in drought tolerance by developing deep roots [89]. Other QTLs, including DRO2 and DRO3, also control deep rooting under normal conditions [159]. In rice, qRL6.1 and qRL7 are associated with root length under hydroponic conditions [160,161]. Overexpressing OsDREB2B, CYP735A, and OsDREB1F improves root morphological adaptations under drought conditions [155]. More than 100 association loci were identified using a GWAS study of 529 rice accessions, providing a genetic basis for drought tolerance improvement [162]. Several genes associated with osmoregulation and late embryogenesis were also identified with a positive role under water deficit conditions [156]. Other genes, such as EcNAC67 and OsPYL/RCAR5, induce a higher root and shoot mass and delay leaf rolling under drought stress [163,164].
Controlling the stomatal aperture is an efficient strategy for developing drought-tolerant plants. The opening of the stomata in response to light and under normal water conditions allows water evaporation via transpiration and the entrance of CO2 into leaves for photosynthesis [93,165]. Under drought stress, plants close their stomata to reduce water loss and improve the water-use efficiency and survival rate. In plant species, ABA regulates stomatal movement to reduce transpiration during drought stress [57,58,166]. In rice, PYR1 and PYL proteins function as ABA receptors. The multiple mutants of group I OsPYL1, OsPYL6, and OsPYL12 display defects in the stomatal movement and transpiration control but promote grain productivity [70]. Overexpressing OsPYL3, OsPYL5, OsPYL9, or OsPYL11 improves drought tolerance but reduces yield under normal conditions [60,62,163]. Several studies reported that H2O2 is essential in stomatal conductance under drought stress in ABA-independent and -dependent manners. For example, ABSCISIC ACID STRESS AND RIPENING5 (OsASR5) improves drought tolerance through a stomatal closure pathway associated with H2O2 and ABA signaling [90]. In addition, DROUGHT AND SALT TOLERANCE (DST) and OsSRO1 work independently of ABA to regulate stomatal closure by modulating H2O2 accumulation [91,167]. OsJAZ1 improves plant growth under drought stress and induces ABA signaling [92]. Stomatal density, size, and index are essential for improving drought tolerance. Plants overexpressing EPIDERMAL PATTERNING FACTOR1 (OsEPF1) reduced their stomatal density, index, and size, improving growth, yield, and drought tolerance [93]. Previously, it was reported that OsDRAP1 (DREB2-like) confers drought tolerance [168]. Introducing OsLEA3-1, OsNAC5, OsWRKY47, OsbZIP71, OsNAC10, or OsbZIP46 in rice improves drought tolerance [169,170,171,172,173,174]. Introducing AtDREB1A, EDT1/HDG11, OsTPS1, or OsMIOX in rice improves the water-use efficiency, antioxidant activity, photosynthesis, and accumulation of osmolytes [175,176,177,178]. The transgenic plants overexpressing OsCPK9 display improved drought tolerance through better osmoregulation and enhanced stomatal closure [94]. Overexpressing OsDREB2A improves the survival rate of rice plants under severe salt and drought conditions [179]. Several CDPKs (CDPK7 and CIPK03/CIPK12) are involved in signal transduction pathways in the response to drought stress [180,181]. Under drought stress, reduced inositol triphosphate levels and ROS homeostasis are observed in plants overexpressing OsITPK2 [95]. The genes from the WRKY F family are essential in rice growth in response to drought stress [182]. Multiple TFs, such as OsMUTE, OsSPCHs, OsICEs, and OsFAMA, also control stomatal movement and development in rice [183]. However, the detailed mechanism of drought tolerance by these TFs remains unknown. Stomatal development is the key parameter to control drought stress and deserves more attention to develop drought-resistant rice varieties.

2.4. Salt Stress

Salt stress affects rice growth and grain quality, directly affecting the market popularity and economic value of rice. Excessive soil salinity reduces the soil’s water potential, makes it difficult for plants to absorb water, and causes physiological drought. High external salt concentrations increase cell membrane permeability and electrolyte extravasation. Plants generally reduce the cellular water potential by regionalizing intracellular salts, absorbing exogeneous inorganic ions, and synthesizing organic osmotic regulators such as soluble sugar, betaine, and proline, thus enhancing plant water absorption capacity and relieving physiological drought. Under salt stress, absorbing external inorganic ions regulates cellular osmolarity in rice. Salt-tolerant rice mainly maintains cell osmotic regulation by increasing the absorption of potassium ions and reducing that of sodium ions [184].
When rice is subjected to salt stress, the excessive accumulation of Na+ in cells reduces the absorption of K+ and Mg2+. The cell is likely to lack phosphorus and Ca2+ (Figure 2A). The antagonism of salt stress on nutrient elements disrupts the growth of the apical meristem and affects chlorophyll synthesis, causing physiological disorders and hindering normal metabolism. Salt stress relatively reduces the content of K+, leading to cell membrane hyperpolarization and a decrease in sugar transport and osmotic pressure and affecting cell extension, growth, and development of shoots and roots. Salt stress inhibits the absorption and utilization of Ca2+, resulting in the blockage of cell wall formation, the inhibition of cell division, and the decline of the membrane system [185,186].
TFs positively or negatively regulate salt tolerance, and 80 TFs are upregulated under salt stress. For example, the TFs OsDREB2A, OsCOIN, OsbZIP71, OsMYB2, and OsbZIP23 can lead to various changes in rice, including an accumulation of antioxidants and osmoprotectants and increased Na+ and K+ transporter activity [187]. Overexpressing these salt-responsive TFs enables efficient osmotic regulation, minimal oxidative damage, and increased seedling survival rates in rice [171,188]. Among the negative regulatory TFs, OsWRKY13 delays the growth and development of plants by inhibiting the expression of the salt-responsive genes SNAC1 and ERD1 [189]. Overexpressing OsWRKY45-2 resulted in a significant decrease in the survival rate of rice plants under salt stress, which was caused by the repressed expression of genes, such as SNAC1, DREB1B, NCED4, and Rab16D, suggesting that OsWRKY45-2 may be a transcriptional repressor of these genes. The regulation of OsWRKY45-2 expression by OsWRKY13 indicated that these two TFs might work together in the same pathway to regulate rice salt tolerance jointly. In addition, rice plants displayed obvious chlorosis and dryness after overexpressing the TFs OsABI5 and OsABI5 [190,191]. These studies indicate that these TFs regulate the expression of different types of genes in rice through various pathways in response to salt stress. Cis-acting elements are the binding sites of TFs, and they can regulate the accurate initiation and transcription efficiency of gene transcription by binding TFs. Many cis-acting elements have been found in rice, and their specific functions have been verified with the corresponding salt-responsive TFs. For example, the promoters of most ABA-inducible genes in rice contain the ABA-responsive element ABRE (ACGTGGC), which regulates their expression under salt stress by recruiting zinc finger TFs, such as OsBZ8 and OsABI5 [192]. ABRE was initially discovered in the rice rab16B promoter, and the binding of OsBZ8 and OsABI5 to ABRE was confirmed by a gel electrophoresis mobility shift and yeast one-hybrid assays, respectively. The presence of the cis-element DRE in the ABA-responsive gene OsbZIP23 and of ABRE in the OsNAC6 and ZFP179 genes indicate a crosstalk between ABA-dependent and -independent signaling pathways [193]. In addition, the cis-element AH2 (CAAT (C/G) ATTG) interacts with the salt stress-related TF Oshox22 to improve rice salt tolerance [194]. In addition to the various cis-acting elements identified above, similar-acting elements are found in the promoters of salt-responsive genes. For example, in rice, the promoter of the C3HC4 ring finger gene contains cis-elements that are induced by salinity and other abiotic stresses [195]. At the same time, several cis-elements related to rice salt tolerance have been identified from different gene interaction networks [196]. All of these cis-elements may be involved in regulating salt-responsive gene expression.
In rice, several genes have been identified as having a role in sensing and transducing salt stress signals and an ability to enhance salt tolerance. Under high-salt conditions, overexpressing the protein kinase SAPK4 can increase the seed germination rate, maintain intracellular ion balance, and improve photosynthesis and growth parameters by upregulating the H+-ATPase genes, NHX1 and OsCLC1 [96]. Overexpressing the calcium-dependent protein kinase OsCPK12 resulted in enhanced salt tolerance due to reduced H2O2 accumulation in leaves and increased root biomass [197]. The MAPK gene OsMAPK44 reduces plant damage induced by salt stress by maintaining a stable ion concentration in plant cells [97]. In addition, transgenic rice plants that were overexpressing another MAPK gene, OsMAPK33, displayed a reduced accumulation of harmful substances in cells and K+/Na+ ratio through the expression of ion transport-related genes or downstream salt-responsive positive regulators in the MAPK pathway [198]. OsSRK1 encodes a typical S-type receptor kinase, and its overexpression induces the expression of OsMyb4, OsDREB1A, ZOS3, and OsWRKY08 to improve rice salt tolerance [199]. Under salt stress, G protein, small G protein, and channel protein levels increase, triggering the expression of salt tolerance-related genes [200].
Transgenic plants overexpressing OVP1 can increase the activities of tonoplast pyrophosphatase and ATPase, providing a proton driving force for the antiporter MHX and sequestering Na+ in the vacuole, reducing the damage of Na+ to the cytoplasm, and improving rice salt tolerance [201]. SKC1 (OsHKT8), a member of the OsHKT family, is a Na+-specific transporter related to Na+ long-distance transport and regulates the above-ground Na+/K+ balance in rice under salt stress [202]. CDPKs directly bind to Ca2+, mediate calcium signal channels, and regulate ABA synthesis and response to salt stress [203]. Overexpressing OsLOL5 increases rice salt tolerance and enhances the expressions of the oxidative stress-related genes OsAPX2, OsCAT, OsCu/Zn-SOD, and OsRGRC2, indicating that OsLOL5 may enhance ROS scavenging to improve rice salt tolerance [98]. In rice, the zinc finger protein DST is downregulated under salt stress, resulting in the downregulated expression of ROS-scavenging genes such as catalase and peroxidase and leading to the accumulation of H2O2, stomatal closure, reduced water loss, and Na+ entrance, thus improving the plant’s salt stress tolerance [91].
Osmoprotectant accumulation is critical for alleviating the intracellular osmotic imbalance in plants under salt stress. In rice, the trehalose pathway genes, OsTPP1 and OsTPS1, the proline biosynthesis pathway gene OsP5CS, and the glycine betaine biosynthesis gene OsCMO promote the overall growth rate of rice, attaining a high-salt tolerance [204,205]. Similarly, under salt stress, high light conditions and CO2 enrichment can enhance the expression of OsBADH1 [99]. These two conditions can affect the photosynthetic efficiency of plants, and OsBADH1 may protect the photosynthetic reaction from salt stress through various mechanisms. Therefore, the biosynthesis and expression of osmoprotectant-related genes are crucial for rice salt stress tolerance.
Homeostasis is essential in determining whether plants can overcome salt stress damage. For example, some protein transport receptors, such as ion pumps or ion channels, maintain the K+/Na+ ion balance in rice under salt stress. Overexpressing several ion balance-related genes, such as OsKAT1, OsHAK5, and OsVP1, can enhance salt tolerance, mainly by increasing the growth rate, photosynthesis, and root biomass. In addition, as a chloride channel-encoding gene, OsCLC1 is upregulated under salt stress and maintains yeast growth under salt stress [100,101,102,206]. The decreased expression of OsTPS1 results in Ca2+ deposition, limiting Na+ entry into the apoplast [207]. Therefore, three genes, OsCNGC1, OsCAX, and OsTPS1, may negatively regulate rice salt tolerance.
In addition, the genes associated with ROS are involved in regulating salt tolerance in rice. Transgenic rice overexpressing OsECS, OsVTE1, and OsMSRA4.1 displayed high-salt tolerance [208,209]. Under salt stress, the exogenous expression of the rice dehydroascorbate reductase gene DHAR increases the germination and seedling growth rate of transgenic Arabidopsis by accumulating high levels of ascorbic acid [210]. In addition, overexpressing the cytoplasmic peroxidase genes, OsAPXa and OsAPXb, improves salt tolerance [211].
Under salt stress, the ectopic expression of the HSP genes OsHsp17.0 and OsHsp23.7 can improve cell membrane stability, germination ability, free proline content, and seedling survival rate while effectively preventing electrolyte leakage and reducing MDA content [212]. In addition, overexpressing the cyclophilin gene OsCYP2 confers a high photochemical efficiency, reduced MDA levels, and increased antioxidant enzyme activity in rice under high-salt conditions, suggesting a key role for this gene in salt stress signaling pathways. The expression of OsRab16A under high-salinity conditions can lead to rapid rice growth, the accumulation of osmotic substances, enhanced antioxidant capacity, and the balance of trace elements [213]. In Arabidopsis, overexpressing OsLEA3-2 increased the seed germination rate under salt stress, presumably acting as a molecular chaperone to maintain enzyme activity during dehydration and effectively preventing destructive protein aggregation [214]. In summary, HSPs and molecular chaperone genes can improve rice salt tolerance by maintaining the stability of protein properties and structures. Most importantly, the roles of sodium transporters in regulating ion homeostasis and the salt stress response require further attention.

2.5. Osmotic Stress

When plants are subjected to low temperatures, drought, and salinity, the concentration of the ion transport changes, leading to an increase in a cell’s permeability to water. This causes the accumulation of many small organic molecules, such as sugars, proline, and betaine, which have osmoprotective effects. Furthermore, there is an increase in the concentration of ABA, which triggers a series of physiological and biochemical reactions, leading to changes in the protein composition [215,216]. To maintain an osmotic pressure balance, plants regulate ion absorption and compartmentation, facilitate water intake, enhance antioxidant defense systems, and alter photosynthetic pathways while permeabilizing the biosynthesis of compatible solutes. Thus, some cellular responses arise from primary stress signals, while others arise primarily from secondary signals. Drought stress and salt stress have independent and some common signal transduction mechanisms. In addition, cold stress reduces the root’s water uptake in rice, but prolonged exposure to low root temperatures induces a gradual increase in the root’s osmotic hydraulic conductivity [217]. To avoid osmotic stress, plant cells reduce the intracellular water potential and maintain intracellular water through osmotic adjustment, thereby ensuring the normal physiological activities of plant cells. Substances involved in osmotic regulation are called osmotic regulators, mainly including inorganic ions, such as Na+, K +, Cl-, and small organic molecules, such as polyols (mannose), nitrogen-containing compounds (proline and betaine), sugars, organic acids (malic acid), and their derivatives. Proline is the most water-soluble amino acid, and its functions include being an intracellular osmotic regulator, reducing agent, energy source, hydroxyl radical scavenger, intracellular enzyme protection agent, N-element storage substance, and regulator of redox potential. When rice seedlings are subjected to osmotic stress, the proline content in leaves and roots is significantly involved in cell osmotic regulation [218].
δ-Pyrroline-5-carboxylate synthase (P5CS) is the rate-limiting enzyme that catalyzes the first two steps in proline biosynthesis from glutamate, and is regulated by the transcriptional level of P5CS [219]. Overexpression of the gene OsP5CS1 can significantly increase the proline content and improve the osmotic tolerance, thereby reducing the oxidative damage of rice under salt stress [103]. The wheat gene Ta-UnP regulates the expression of P5CS, and the overexpression of Ta-UnP in rice can highly induce the expression of OsP5CS1, resulting in increased proline content, and improving the osmotic and salt stress tolerance of rice [220]. The wheat gene TaPUB15 encodes a U-box E3 ubiquitin ligase, which is highly expressed in seedling roots and induced by salt stress. The overexpression of wheat TaPUB15 in rice can significantly increase the transcription level of OsP5CS1, and at the same time, the transgenic lines have more roots and an improved salt tolerance [221]. The Alfin-like (AL) family is a group of small plant-specific TFs involved in abiotic stresses in dicotyledons. In an early study, it was found that an AL gene in rice was associated with the grain yield under drought stress. It was also found that Hap1 of OsAL7.1 and Hap7 of OsAL11 were favorable haplotypes of seed weight and germination under osmotic stress. Additionally, during the germination stage, the osal7.1 and osal11 mutants have larger seeds and are more susceptible to mannitol and abscisic acid. In contrast, the overexpression of OsAL7.1 and OsAL11 reduced the stress tolerance at the adult stage [222]. Wang et al. [104] used the CRISPR/Cas9 system to obtain a mutant of the inorganic pyrophosphatase OsPPa6 gene. In the mutant, inorganic phosphorus, ATP, chlorophyll, sucrose, starch, net photosynthetic rate, soluble sugar, and proline were significantly reduced, while the MDA, osmotic potential, and Na+/K+ ratio significantly increased, indicating that the OsPPa6 gene is an important osmoregulatory factor in rice. Calcium-binding proteins are important in signal transduction for growth and stress response. In rice, OsCCD1, a novel small calcium-binding protein with a centrin-like domain, was characterized. OsCCD1 binds Ca2+, and its expression is induced by osmotic and salt stress and positively regulates tolerance to these stresses through involvement with genes such as OsDREB2B, OsAPX1, and OsP5CS (Figure 2B). It was observed that osmotic and salt stress dramatically increased the expression of OsCCD1 via the calcium-mediated ABA signal [105].
Annexin is a multi-gene family of calcium-dependent phospholipid-binding proteins that are found in plants and other organisms. In plants, annexins play various roles such as membrane trafficking, abiotic stress responses, and signal transduction [223]. Interestingly, it was found that OsANN10, a putative annexin gene in rice, negatively regulated plant responses to osmotic stress. Knocking down OsANN10 significantly decreased the content of H2O2 by increasing POD and CAT activities, suggesting a negative regulation of OsANN10 in protecting the cell membrane against oxidative damage via scavenging ROS under osmotic stress [106]. Proteins that regulate cell wall polysaccharide synthesis play a crucial role in osmotic stress tolerance in plants. During osmotic stress, these proteins control the deposition and reorientation of cell wall components, which helps maintain cell integrity and prevent water loss. In a study, the role of the protein OsCSLD4, involved in regulating cell wall polysaccharide synthesis was evaluated in response to osmotic stress. The study shows that OsCSLD4 plays a positive role in osmotic stress tolerance in rice by regulating the ABA content. This study also conducted a transcriptomic analysis to investigate the genes involved in the response to osmotic stress and their regulation by OsCSLD4. The findings suggest that OsCSLD4 plays a crucial role in regulating the expression of genes involved in ABA biosynthesis and signaling pathways, which are essential for the response to osmotic stress [107].
The balance between hormonal signaling and osmotic stress tolerance in plants plays a crucial role in promoting plant growth and development and regulating stomatal closure and the accumulation of compatible solutes. It was found that the OsGA2ox8 is induced by various abiotic stresses and phytohormones, and its overexpression enhances osmotic stress tolerance in rice by increasing osmotic regulators and antioxidants. It was observed that OsGA20ox8 is preferentially expressed in shoots and roots under osmotic stress and is also involved in regulating genes that are associated with anthocyanin and flavonoid biosynthesis, as well as JA and ABA biosynthesis pathways [224]. Recently, OsNF-YA3, a rice TF, was found to regulate this GA and ABA balance by activating GA biosynthetic genes and enhancing GA content while repressing the ABA response to stabilize plant growth and osmotic stress tolerance. It was observed that OsNF-YA3 negatively regulates plant osmotic stress tolerance by binding to the promoters of ABA catabolic genes (OsABA8ox1 and OsABA8ox3) and reducing ABA levels. The activity of OsNF-YA3 is inhibited by the DELLA protein SLR1, and OSMOTIC STRESS/ABA-ACTIVATED PROTEIN KINASE 9 also interacts with OsNF-YA3 and mediates its degradation [225]. The specific roles of type 2C protein phosphatases (PP2Cs) in rice abiotic stress tolerance are not well understood. In rice, 90 PP2C genes have been predicted, but very few members have been functionally characterized [226]. In a previous study, the function of OsPP65 in osmotic and salt stress tolerance in rice was investigated. It was observed that OsPP65 is expressed in rice seedlings and leaves and induced by multiple stresses. The knockout of OsPP65 increased osmotic and salt stress tolerance in rice plants through independent regulation of JA and ABA signaling pathways. Metabolomics analysis indicated modulation of raffinose family oligosaccharide metabolism pathway in rice. OsPP65 is a potential target for improving rice stress tolerance using gene editing [108].
In rice, the HSP90 family gene, OsHSP50.2, was found to be ubiquitously expressed, and its expression was induced by heat and osmotic stress treatments. The OsHSP50.2 overexpression lines showed reduced water loss and enhanced or improved tolerance to osmotic and drought stresses. Overexpression lines exhibited significantly lower levels and less electrolyte leakage but higher SOD activity. It was also observed that the OsHSP50.2-overexpressing plants accumulated a significantly higher proline content and improved osmotic adjustment to drought stress damage [227]. In Arabidopsis, a hyperosmolality senser, OSCA1 (osmolality-sensing ion channel 1), was identified. The homolog of OSCA1 from rice (OsOSCA1.2), consisting of 11 transmembrane (TM) helices and a cytosolic soluble domain that has homology to RNA recognition proteins, also mediates hyperosmolality sensing, transport pathway gating, and balances the intracellular Ca2+ concentration. The TM domain is similar to the TMEM16 family and has a unique structure with extended helical arms. These arms may detect tension on the lipid bilayer caused by turgor pressure and open a transport pathway, thus enhancing osmotic stress tolerance [109]. Receptor-like cytoplasmic kinases (RLCKs) are mainly involved in growth regulation and pathogen responses in plants; however, their role in abiotic stress tolerance remains elusive. A study found that OsRLCK241 is not only induced by salt and drought stresses but also by ROS detoxification by accumulating more compatible osmolytes and enhancing the activities of ROS scavengers to alleviate the osmotic stress evoked by drought and stress [228]. The rice plants have evolved various mechanisms to tolerate osmotic stress, including the regulation of gene expression, the synthesis of compatible solutes, and hormonal regulation. However, knowledge about the complex signaling pathways involved in osmotic stress responses in rice, as well as the interactions between different stress response pathways, is still very limited compared to other stresses.

2.6. Submergence Stress

Submergence or flooding can affect oxygen, CO2, light, and nutrient uptake, thus inhibiting photosynthesis, accelerating energy consumption, and causing plant growth atrophy or death [229,230]. To adapt to excess water conditions, rice has developed specialized anatomical and morphological traits, such as aerenchyma, radial oxygen loss barriers, adventitious roots, and the ability to form a leaf gas film. In addition, plants produce ventilated tissues and ethylene to facilitate gas exchanges and the programmed cell death of cortical and epidermal cells [231,232]. Moreover, the mechanical force generated by the growth of adventitious roots regulates epidermal cell death [233]. However, these strategies are insufficient for survival under continuous and complete submergence, resulting in stunted growth or complete death because of photosynthesis inhibition and the fast consumption of energy reserves. The submergence conditions rapidly trigger GA accumulation, resulting in rapid internode elongation [234]. When submerged, rice plants can form a protective wall around their roots called a radial oxygen loss (ROL) barrier from the root base to the root tip and outside the aerenchyma to prevent oxygen loss that has reached the roots [235]. Gene expression and mutant analyses suggested that suberin is the main component of the ROL barrier in rice roots, but the detailed mechanism of barrier formation remains to be clarified [236,237]. To overcome prolonged submergence, some Asian rice varieties have developed additional traits, such as aerobic germination, quiescence of leaf elongation in response to flash floods, and internode elongation under periodic flooding. Some rice cultivars tolerate submergence for ~15 days by restricting the consumption of carbohydrates, degradation of chlorophyll, and elongation growth [110,238].
Rice adaptation to submergence stress includes submergence escape and tolerance, which are controlled by SNORKEL (SK) and SUBMERGENCE-1 (Sub1), respectively, belonging to the ETHYLENE RESPONSIVE FACTOR (ERF) class of TF genes [234,238]. The functional analysis of ERF-type TF genes revealed their involvement in controlling various physiological and morphological responses under submergence conditions.
When rice germinates under anaerobic conditions, radicle elongation is suppressed, while the coleoptile elongates and exposes its tip to the water surface to secure oxygen [239,240]. During anaerobic germination, the expression of the amylase alcohol dehydrogenase (ADH) genes is induced, leading to stored starch decomposition and ethanol fermentation to secure energy for elongation and promote anaerobic respiration. In rice, OsTPP7 (trehalose6-phosphate phosphatase 7) is involved in anaerobic respiration, promoting energy supply to the coleoptile by increasing the metabolism of trehalose 6-phosphate (T6P) in the embryo [241].
Submergence-tolerant rice exhibits growth atrophy, which is beneficial to reduce energy consumption and ensure rice survival for several weeks under flooding stress and sufficient energy under continuous submergence. A major QTL (Sub1) on chromosome 9 of the submergence-tolerant indica cultivar ‘FR13A’ explains 70% of the phenotypic variation [242]. The Sub1 region of submergence-tolerant varieties contains clusters of three similar genes, SUB1A, SUB1B, and SUB1C, encoding ERF-like TFs, among which SUB1A is the most studied. SUB1B and SUB1C are invariably present in the Sub1 region of all the rice accessions analyzed. Comprehensive genetic studies revealed that SUB1A introgression with SUB1B and SUB1C confers a robust tolerance to submergence without affecting rice grain yield and quality [110,238,243,244]. In addition, SUB1A is present in a limited number of indica and aus cultivars, whereas SUB1B and SUB1C are found in all rice accessions [238]. SUB1A reduces energy consumption by negatively regulating the expression of genes that encode enzymes that degrade starch and sucrose and positively regulating ADH and pyruvate under submergence conditions [110]. In addition, SUB1A inhibits ethylene synthesis and the expression of cell wall loosening and expansion proteins under flooding stress, maintaining high chlorophyll a and b contents (Figure 2C). Through magnetic resonance, metabolites controlled by SUB1A under flooding stress were analyzed. As a result, SUB1A was involved in carbohydrate consumption, amino acid accumulation, and aerial part elongation [245]. In rice, two ERF multigenic SUBMERGENCE-1 (SUB1) and SNORKEL (SK) loci govern the quiescence versus escape antithetical adaptive responses [110,234,238,246]. The detailed molecular and physiological studies of SUB1A and SK genes indicated that these ERF VII genes respond to submergence through the same hormonal pathways and regulate antithetical growth. SUB1A and SKs display higher transcriptional accumulation in response to ethylene under submerged conditions [110,234]. SUB1A positively regulates adaptability to flash flooding, as plants are re-exposed to atmospheric oxygen and post-anoxic injury because of oxidative stress [247]. In addition, SUB1A improves oxidative stress tolerance by regulating gene-encoding enzymes related to ROS detoxification [248]. Interestingly, SUB1A induces the transcription and protein accumulation of SLR1 (slender rice1), a suppressor of GA signaling, causing elongation suppression in submerged conditions [246]. Rice plants that are sensitive to flooding lack the genomic region containing the SUB1A gene or have mutations that cause amino acid substitutions in the Sub1A protein [238]. The amino acid residue involved in this mutation is essential for the phosphorylation of Sub1A by MPK3, an MAP kinase [249]. Phosphorylated Sub1A induces the transcription of MPK3, forming a positive feedback loop, and the transcription of ERF66 and ERF67. Those genes induce the transcription of genes that contribute to flood tolerance, such as ADH [250].
Two key genes, SNORKEL1 and SNORKEL2, were cloned and studied. As a result, the study indicated that they respond to flooding stress by encoding response factors involved in ethylene signaling [251]. Under submergence, ethylene accumulates in rice and induces the expression of SNORKEL1 and SNORKEL2, finally promoting internode elongation through GA. An attempt was made to identify the genetic factors for floating rice; they were identified on chromosomes 1, 3, and 12 [252,253,254,255]. A subsequent linkage analysis, using ‘Taichung No. 65’ (non-floating rice) and ‘C9285’ (floating rice from Bangladesh), revealed that the QTL located on chromosome 12 was the same ethylene responsive gene as SUB1A. SNORKEL1 and SNORKEL2 were identified in the genomic region deleted in non-floating rice. The OsEIL1 protein is stabilized in response to ethylene and regulates downstream gene expression [111]. In floating rice, the expression of SNORKEL1/2 is induced by the OsEIL1 protein stabilized by ethylene accumulation under deep-water stimulation, thus promoting internode elongation through the expression of downstream genes. Interestingly, SNORKEL1/2 belongs to the same subgroup of the ERF family as SUB1A and ERF66/67 [234,250]. Elucidating the detailed signaling mechanisms downstream of SNORKEL1/2 may provide a molecular evolutionary answer to explain why similar TFs function differently in flood adaptation.
To search for new factors that are involved in floating rice, a GWAS was performed on 68 rice cultivars collected from various Asian countries with various floating rice qualities and different internodal lengths [256]. As a result, a region on chromosome 1 was strongly correlated with the internode length. A high-density linkage analysis was performed, and GA20oxidase2 and SD1 (GA 20 oxidase 2; SEMIDWARF1) involved in GA synthesis were identified in the region controlling internode length in floating rice. Comparing different high-floating rice cultivars revealed 17 polymorphic mutations in the promoter and intron regions commonly found in rice [257]. The SD1 allele, SD1DW, containing these mutations significantly affected floating rice in the presence of SNORKEL1/2. SD1DW produces the active GA hormone, GA4, rather than GA1, which has a stronger effect on internode elongation. Later, it was found that OsEIL1 directly binds to a specific region of the SD1DW promoter, revealing that SD1DW is under the control of OsEIL1 and SNORKEL1/2 [258].
Using GA-dependent internode elongation as an index, another gene, ACCELERATOR OF INTERNODE ELONGATION 1 (ACE1), was identified as being induced by GA in a deep-water environment and ultimately inducing internode elongation. However, overexpressing ACE1 alone does not result in internode elongation; initiating internode elongation in floating rice requires ACE1 and independent GA signals. Another QTL, DEC1, encoding a C2H2 zinc finger-type TF and regulating GA-dependent internode elongation, was identified on chromosome 12. Overexpressing DEC1 inhibited internode elongation, while dec1 mutants displayed internode elongation. DEC1 was expressed near the inter-meristem, where cell division was activated in the dec1 mutant. In floating rice, GA administration and deep-water conditions decreased the expression of DEC1, suggesting that this decrease leads to the release of cell division inhibition during internode elongation. Five ACE1-like (ACL) sequences are found in rice, and overexpressing ACL1 induces internode elongation when administered with GA. ACE1 has lost its function in rice, while ACL1 retained it; its expression is induced during reproductive growth [112].

3. Utilization of Integrative Omics Analyses to Identify Potential Candidate Genes Associated with Abiotic Stress

Recent advances in transcriptomic, proteomic, and metabolomic technologies now offer opportunities for candidate gene identification for subsequent utilization in plant breeding and potential improvements in global food security. The efficacy of breeding for increased environmental stress tolerance could be significantly improved through candidate gene identification and their utilization in marker-assisted selection. Several candidate genes underlying rice abiotic stress tolerance have been identified using functional genomics platforms. This section describes some examples from the literature about such candidate genes. Several stress responsive key genes, proteins, and metabolites have been uncovered through integrative omics studies and can be functionally characterized to utilize in breeding programs (Table 2).

3.1. Transcriptomic Data

The RNA-seq analyses upgraded our knowledge of gene expression in rice under different environmental conditions. Recently, RNA-seq has become more advanced, with increased detection sensitivity and sequencing depth. To date, transcriptome analysis has identified candidate genes linked to several abiotic stress responses in rice plants. A transcriptome study identified several DEGs for chilling stress, and by combining QTL fine-mapping and functional identification, the co-localized DEGs may provide insights for elucidating the mechanism of chilling tolerance and the basis for gene cloning. The study identified 154 highly expressed genes encoding oxidoreductases, thioredoxins, and glutathione S-transferases, among others. These genes help maintain cell redox homeostasis and promote chilling tolerance. Some stress-responsive genes encoding death-associated protein kinase 1, calcium homeostasis regulator CHoR1, four LRR-containing proteins, and four F-box domain-containing proteins were also part of this group. Another group of 27 genes, including 2 genes encoding short-chain dehydrogenase/reductase family proteins, 6 genes encoding TFs, and 3 stress-responsive genes encoding senescence-associated protein DIN1, polygalacturonase inhibitor 1, and hairpin-induced protein 1, were identified, indicating their positive role in chilling stress responsiveness. Several genes with a negative role in redox regulation, metabolism, stress response, and transport were identified, including genes encoding peroxidases, wall-associated kinase 3, oxidoreductases, and receptor-like kinase ARK1AS. Interestingly, two HSPs, an HSF, a photosystem II D2 protein, and a set of protein kinase domain-containing proteins were also discovered. Although HSPs and HSFs are known for their role in heat shock response in plants and other organisms, their function in chilling stress responsiveness in plants remains to be analyzed [259].
In an integrated transcriptome-based study of heat-resistant and -tolerant indica rice varieties, three candidate genes were reported as essential for high-temperature stress tolerance at the seedling stage. First, the study identified genes related to ROS (OsCML4), plant hormone signal transduction, metabolic pathways, cysteine and methionine metabolism pathways (LOC_Os01g09450, LOC_Os03g59040, and LOC_Os12g42980), stomatal conductance (LOC_Os03g59040), and survival rate (LOC_Os02g12890) under heat stress. Finally, LOC_Os01g09450, LOC_Os03g59040, and LOC_Os12g42980 were potential candidate genes for high-temperature stress tolerance in seedlings [260]. Another study utilized GWAS and a transcriptome analysis and identified 11 genes that were associated with the rice heat stress response by examining the SNP variation and gene expression. LOC_Os03g16460 encodes a putative uncharacterized protein, and LOC_Os05g07050 encodes a putative pre-mRNA processing splicing factor 8. By analyzing the 221 rice accessions, seven mutation sites classified into five categories and three mutation sites classified into two categories were found for LOC_Os03g16460 and LOC_Os05g07050, respectively. Amino acid changes were observed in the encoded proteins due to these mutations. For LOC_Os03g16460, plants with mutations at all seven sites displayed the highest heat tolerance, whereas plants carrying a single mutation displayed a relatively low heat tolerance—a low survival rate. Similarly, for LOC_Os03g16460, plants with a single mutation were less tolerant to heat than plants with all three mutations. Interestingly, plants with more mutation sites displayed higher heat tolerance in indica rice accessions, whereas fewer mutation sites in japonica accessions tended to have lower heat tolerance [261].
Subsequently, several heat-responsive genes mainly involved in stress response, transport, transcriptional regulation, protein binding, and antioxidant activities were identified through transcriptomic analyses. The time-dependent expression pattern of genes under heat stress included several heat shock TFs (Hsf) and Hsp candidate genes. Hsf (Os02g0527300, Os03g0133000, and Os06g0662200), Hsp (Os03g0218500, Os03g0224700, and Os04g0107900), glycolysis (Os01g0160100, Os05g0187100, and Os05g0524400), and ubiquitin–proteasome system (Os03g0116700, Os05g0160200, and Os06g0173100) gene expression was generally enhanced under heat stress, whereas genes related to specific secondary metabolisms (Os06g0656500, Os08g0498400, and Os09g0492700) and light reaction (Os01g0720500, Os01g0938100, and Os07g0105600) were repressed by heat stress [262]. It is worth mentioning that the whole transcriptome identified 37 novel heat-responsive genes, including many TFs. Remarkably, only 22 out of the 37 genes were heat-responsive and part of a tissue-independent general response; they were also identified in another RNA-seq study. In addition, 46 genes representing HSFs (OsHSFA2f and HSFB2c), molecular chaperones, and co-chaperone families were highly regulated during anthesis under heat stress. Interestingly, a small HSP (LOC_Os02g52150) was also highly upregulated in the transcriptomic data [263]. Another transcriptomic investigation suggested 589 candidate genes related to heat stress responses in rice seedlings. Six heat stress-responsive genes were identified that are associated with a broad range of abiotic stress responses, such as drought, salt, cold, and submergence. These genes are potential candidates for fighting combined stresses. In addition, OsNAC6/SNAC2 (Os03g60080) was upregulated under heat and drought stress, indicating a strong correlation between the two. Furthermore, five HSP10/CPN10 and thirteen HSP20 genes were associated with prolonged heat stress responses, but none of the HSP10/CPN10 or HSP20 genes were associated with an early heat response. Among them, eight HSP20 genes were significantly upregulated in response to multiple abiotic stresses, especially heat and drought. On the other hand, HSP10/CPN10 were not stimulated by other abiotic stresses. These findings suggest that the HSP20 family proteins have evolved to cope with multiple abiotic stress challenges and prolonged heat stress, whereas the HSP10/CPN10 family is specialized to overcome prolonged heat stress [264]. In addition to heat stress, several transcriptome studies have been utilized and successfully identified candidate genes involved in other stresses.
Another transcriptome analysis was performed on 4-week-old rice seedling roots subjected to drought stress for two and three days. As a result, 1098 genes were upregulated in response to 3-day drought stress, which caused severe damage to root development after recovery, unlike the 2-day stress. Interestingly, 68% of candidate genes were not identified in previous transcriptomic studies. Further bioinformatics analysis revealed that RING-box E3 ligases in the ubiquitin–proteasome pathways are significantly stimulated by drought. Finally, the study analyzed the functions of 66 candidate genes and found 29 genes directly involved in drought tolerance. The identified genes were involved in metabolic pathways, protein modification, and protein degradation [265]. In the transcriptomic study, several NAC (CUC, NAM, and ATAF) family genes were upregulated under salt stress, and among them, 14 were reported in previous studies to be associated with abiotic stress responses. Additionally, two HD-ZIP and two MYB genes were highly upregulated, making them candidates for salt stress tolerance in rice [266]. Under salt stress, a transcriptome study of rice roots found that several candidate genes related to multiple stresses. A further analysis identified 13 candidate genes that were already functionally characterized and involved in phytohormone signaling and salt stress tolerance [267].
Another study identified non-coding RNA—miRNAs and lncRNAs—and several candidate genes involved in salinity and submergence stress responses through comparative transcriptome analyses of wild rice under different stress conditions [268]. RNA sequencing of two contrasting recombinant inbred lines (RILs) retrieved five candidate genes, including LOC_Os01g04430, LOC_Os01g04530, and LOC_Os03g22720 for qCL-1.1 and qCL-3.1. These genes are responsible for different metabolic processes, and the RILs expressing these genes were submergence-tolerant [269]. After utilizing the deep transcriptomes of rice leaves and roots under drought, cold, and salt stresses, many upregulated and downregulated genes specifically and commonly expressed under those stress conditions were identified. Data mining identified the expression of key functional and regulatory genes and conserved cis-regulatory elements in the promoter of highly presented genes involved in different abiotic stress signaling pathways [270].
Although several transcriptome studies have reported candidate genes and TFs involved in the response to abiotic stress in rice, the expression landscape of these genes at a single-cell resolution remains poorly understood. From the perspective of transcriptomic work, the importance of individual genes and gene sequences can be further confirmed in genome editing experiments.

3.2. Proteomics Data

A proteomics analysis provides a broad picture of plant stress responses at the protein level. Because of advancements in two-dimensional polyacrylamide gel electrophoresis (2DE), protein detection and quantification, fingerprinting, and partial sequencing by mass spectrometry, the sensitivity, and power of proteomic approaches have increased in recent years.
Some candidate components, including a protein similar to WAK1, a putative armadillo repeat-containing protein, and a protein phosphatase 2C-like protein involved in signal transduction under cold stress—RAB2A—were identified [271]. In a proteomics analysis of cold-treated anthers, 441 DEPs were identified, and following a bioinformatics analysis, 30 upregulated and 4 downregulated were considered candidate genes. Most were related to glycine-rich proteins and C2 domain proteins [272]. An iTRAQ-based proteomics approach identified 85 proteins related to low temperature stress, enriched mainly in transport, photosynthesis, precursor metabolism, and energy production, as well as histones and vitamin B [273].
The proteomic analysis of rice under drought stress alone and In combination with heat stress identified proteins favoring pollen germination. A comparative iTRAQ-based proteomics study compared the proteomes of heat-tolerant and -sensitive rice lines at an early milky stage. A total of 38 DEPs were identified, among which 32 were functionally annotated in NCBI and UniProt databases. These proteins were related to the defense response, transport, energy metabolism, signal transduction, transcript regulation, oxidation, and biosynthesis. A further analysis identified downstream genes/proteins involved in the high light temperature response [274]. Another proteomics study identified 54 candidate proteins involved in high-temperature stress during the first half of the ripening period of heat-sensitive rice lines [275].
Furthermore, two small HSPs (16.9 and 17.4 kDa) were strongly expressed under combined stress at the protein and transcript levels. Comparing the protein and transcript expression patterns for these two genes suggested differential gene regulation at the transcriptional and translational phase in response to heat and drought stress [276]. Proteomics analysis was performed to elucidate the molecular responses of Thai jasmine rice under drought, and 623 DEPs were identified. After a comprehensive analysis, a candidate protein, DEAD/DEAH-box RNA helicase, was identified. It is involved in DNA repair and was reported as a novel protein for promoting plant stress tolerance [277,278]. In another work, iTRAQ-based proteomics identified 1221 DEPs in rice grown under control and light conditions, and a candidate gene-encoding rice β subunit of glyceraldehyde-3-phosphate dehydrogenase (OsGAPB) was identified and functionally characterized [279].
In a proteomic study, the response of rice suspension cells to salt stress using the iTRAQ technique was studied, and significant changes in the carbohydrate and energy metabolism pathways, redox signaling pathways, auxin pathways, and biosynthetic pathways of osmolytes were observed [280]. Another study used iTRAQ techniques to study the proteome changes in rice under salt stress and found 56 proteins showing significantly altered expression, 16 of which were mainly enriched in the processes of photosynthesis, anti-oxidation, and oxidative phosphorylation, which play an important role in maintaining energy balance and generating active oxygen under stress conditions [281]. In another proteomic study, the rice seedlings of spotted leaf 1 (spl1) were completely submerged for five days to initiate cell death and then exposed to air, and the samples were collected at different time points. Finally, PBZ1, as a putative cell death marker, was identified, which was highly inducible in spl1 [282]. In short, using proteomics to identify candidate proteins—and genes—will help improve rice abiotic stress tolerance, which is reflected in rice yield and quality.

3.3. Metabolomics Data

Metabolic compositional variations reflect the changes in gene expression occurring in stress responses. Studying metabolic pathways and their interconnections in the context of systems biology is progressively becoming common for identifying candidate genes. To reveal the candidate genes, integrating transcriptomics, metabolomics, and proteomics, the data provide a comprehensive understanding of the molecular mechanisms underlying stress tolerance in rice. Stress tolerance in rice requires a fully functional metabolism throughout the stress, but the regulation of and the link between associated metabolites and transcripts remain largely unknown.
The comparative metabolomics analysis was performed for two subspecies of rice with significant divergence in chilling tolerance. Finally, 9 metabolites (glutathione, putrescine, asparagine, β-Alanine, γ-Glutamylleucine, oxalate, mannose-6-phosphate, isocitrate, and nicotinamide mononucleotide) with obvious changes in relative abundance under chilling treatment [283]. The metabolomics analysis analyzed heat-treated single mature pollen grains in two contrasting rice spikelet fertility studies under heat stress. Finally, contrasting varietal differences in phosphatidylinositol (PI) (34:3) were detected in mature pollen, together with other 106 metabolites. The major categories were linked to lipid-related metabolites, organic acids, amino acids, and related metabolites, carbohydrates, cell wall-related metabolites, plant hormones, and cluster ions [284].
In rice, integrative transcriptomics and metabolomics analyses of drought-tolerant and susceptible cultivars explored transcript and metabolic responses under drought stress. A study found 4059 and 2677 DEGs and 69 and 47 differential metabolites under drought and normal conditions, respectively. A correlation analysis suggested several candidate genes based on DEGs involved in the associated metabolites’ pathways. The upregulation of 4-hydroxycinnamic acid and ferulic acid correlated with the performance of photosynthesis-related DEGs during the early stages of drought. The identified metabolites, 4-hydroxycinnamic acid and ferulic acid, were reported to be critical for rice drought tolerance, and the DEGs involved in the pathways of these metabolites were suggested as promising candidate genes for improving drought tolerance [285]. The potential 46 candidate genes identified from a transcriptomics study were confirmed through metabolite profiling in 21 cultivars under drought conditions. The study revealed that, during drought stress, the levels of most metabolites had a negative correlation with the performance parameters. However, eight metabolites showed a positive correlation, including allantoin, galactaric acid, gluconic acid, glucose, salicylic acid glucopyranoside, and three unidentified metabolites. Furthermore, 28 genes were found to have a significant correlation between expression level and performance under drought, with predominantly negative associations. Among those showing a significant positive correlation was a gene responsible for encoding a cytosolic fructose-1,6-bisphosphatase enzyme, which plays a key role in regulating C-metabolism [286]. Another transcriptome and metabolome analysis of two contrasting genotypes under salt stress revealed the repression of genes involved in metabolite degradation and the upregulation of genes involved in several pathways underlying salt tolerance. Combined analyses identified 43 candidate genes related to salt stress. With more precise analysis, the metabolic pathways leading to the primary metabolites examined were investigated, revealing two significant findings. Firstly, the accumulation of metabolites induced by salt stress in the shots of both genotypes was primarily limited to a few sugars, including glucose, fructose, and fructofuranans, as well as proline, all of which were ABA-mediated. On the other hand, salt-induced metabolic accumulation in the roots of both genotypes involved products of N metabolism, such as allantoin, urea, and glutamine, and trehalose was the only sugar with a relatively high accumulation in the roots of both genotypes. Finally, changes in gene expression led to the coordinated accumulation of key primary metabolites in the shoots or roots of the salt-sensitive rice line and salt-tolerant progeny, with 12 key genes identified that are related to the identified primary metabolites [287].
Metabolic profiling demonstrated the presence of SUB1A in the crossbred line M202 (Sub1), which has a higher tolerance than wild-type M202 under deep flood conditions. The presence of SUB1A was found to be associated with a reduction in carbohydrate metabolism in shoot tissue, which is consistent with its role in limiting starch catalysts for elongation growth. On the other hand, the absence of SUB1A resulted in an increase in sucrose consumption and the accumulation of amino acids synthesized from glycolysis intermediates and pyruvate. Alanine, produced through pyruvate metabolism, showed the largest difference between the two contrasting varieties under submergence conditions. However, elevated levels of glutamine, glutamate, leucine, isoleucine, threonine, and valine were also higher in the absence of SUB1A. Additionally, the compound alanylglycine (AlaGly) was identified and characterized, and its levels under submergence conditions were assessed [288]. In most cases, metabolomics has been utilized to study the expression of metabolites in response to different stresses, but very few studies have focused on identifying candidate genes. However, it is clear from the abovementioned studies that combining metabolomics with other omics approaches may enable the identification of the intended candidate genes.
Table 2. List of some abiotic stress related key genes, proteins, and metabolites identified through omics studies.
Table 2. List of some abiotic stress related key genes, proteins, and metabolites identified through omics studies.
Stress TypeNo. of Key FactorsFunctional Categories or Names of FactorsReferences
Transcriptomics
Cold154Oxidoreductases, thioredoxins, glutathione S-transferases, and cell redox homeostasis[259]
Heat3Plant hormone signal transduction, metabolic pathways, cysteine, and methionine metabolism pathways[260]
11Pre-mRNA processing splicing [261]
18Transcriptional regulation, transport, protein binding, antioxidant, and stress response[262]
22Heat response, molecular chaperone, and co-chaperone [263]
Drought29Metabolic pathways, protein modification, and protein degradation[265]
salt13Phytohormone signaling and salt stress[267]
Submergence5Metabolic processes[269]
Proteomics
Cold stress34Glycine-rich proteins and C2 domain proteins[272]
85Transport, photosynthesis, precursor metabolism, and energy production, as well as histones and vitamin[273]
Heat32Defense response, transport, energy metabolism, signal transduction, transcript regulation, oxidation, etc.[274]
Drought2Heat shock proteins[276]
1DNA repair[277]
Salt58Carbohydrate and energy metabolism pathways, redox signaling pathways, auxin pathways, etc.[280]
16Photosynthesis, anti-oxidation, and oxidative phosphorylation[281]
Submergence1Programmed cell death[282]
Metabolomics
Cold9Glutathione, putrescine, asparagine, β-Alanine, γ-Glutamylleucine, oxalate, mannose-6-phosphate, etc.[283]
Heat109Lipid-related metabolites, organic acids, amino acids and related metabolites, carbohydrates, etc.[284]
Drought2Metabolites, 4-hydroxycinnamic acid, and ferulic acid[285]
8Allantoin, galactaric and gluconic acid, glucose and salicylic acid glucopyranoside, etc.[286]
Salt16Fructofuranose, fructose, glucose, proline, urea, and allantoin [287]
Submergence1Alanylglycine[288]

4. Conclusions and Perspectives

The molecular mechanism of response to abiotic stress has gradually become more precise in rice. In recent years, researchers have conducted many studies on the signal transduction and gene expression regulatory pathways during the response to major abiotic stresses, and many hormones’ and TFs’ roles in plant stress tolerance have also been revealed. However, most of the research progress at this stage focuses on discovering unknown TFs, editing promoter binding sites, and working on the biochemical level. New research directions, tools, and practices must be developed. The current research has identified many TFs that bind to cis-acting elements of target gene promoters to activate the expression of specific downstream genes. In rice, the binding sites of these TFs have been analyzed, and their binding mechanism has been studied in detail. However, further research on multiple TFs synergistically acting as negative feedback regulators during stress signaling is necessary. In addition, the research conducted on the exploration of alternative approaches such as epigenetic modifications or small RNA regulations is still very limited. So far, many studies have shown that epigenetic regulation is involved in the response process of plant abiotic stress, and this new understanding has also introduced new research content and directions for epigenetic research. Epigenetic regulation under abiotic stress is a dynamic process, so it is very difficult to study the epigenetic mechanisms of plants under stress. There are still research gaps to reveal the outcomes of epimutations of cloned genes and their regulatory networks related to abiotic stress responses and epigenetic information. The epigenetic modifiers and specific epigenetic modification sites of stress responses and the corresponding stress-response target genes have not been fully understood. Under abiotic stress, the interaction between epigenetic regulation and plant endogenous hormones at different levels needs to be further studied. A series of epigenetic regulatory factors have been identified to participate in plant abiotic stress responses; how to apply them to crop breeding and improve agricultural production needs to be further explored.
Big data analysis and a variety of new methods are helpful to quickly find key players for broad-spectrum stress tolerance and can be further cloned to study their regulatory mechanisms. Analysis of the structure of key proteins, especially receptor proteins, and identification of important protein domains and key amino acid sites facilitate target designs for gene editing and other technical means. In addition, revealing the mechanism of plant abiotic stress responses requires comprehensively utilizing genomics, transcriptomics, and proteomics methods at the molecular level. Identifying a novel gene network that can sense stresses, transduce signals, and activate TFs and downstream genes directly related to stress tolerance will effectively support conventional and molecular breeding programs and genetic engineering strategies. Co-expression networks are an attractive framework for gene interaction analysis and offer various applications, from functional gene annotation to the comparison of co-expression networks across species. This approach will further contribute to the elucidation of important biological processes and provide a valuable predictive tool for modern molecular breeding and crop engineering strategies. At the cytological level, with the rapid development of microscopy techniques, cryo-transmission electron microscopy or high-resolution fluorescence microscopy to observe changes in plant cell membranes, organelles, and cytoskeletons will help better understand and analyze the cellular responses to different abiotic stresses. In short, cutting-edge technologies, such as whole genome sequencing platforms, high-throughput integrative omics techniques and resources, targeted genome editing technologies, and synthetic techniques, are growing and require utilizing basic research in advanced breeding methodologies. These latest improvements in precision breeding and data analytics, combined with stress-tolerant plants, can produce rice for the global market and achieve global food targets.

Author Contributions

Conceptualization, B.U. and K.-H.J.; writing—original draft preparation, B.U. and B.D.; writing—review and editing, B.U. and K.-H.J.; funding acquisition, K.-H.J.; validation, B.U. and K.-H.J.; visualization; K.-H.J.; supervision; K.-H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (2021K1A3A1A61002988 and 2021R1A2C2010448 to K.-H.J.), and a grant from Cooperative Research Program for Agriculture Science and Technology Development (PJ01703502 to K.-H.J.) of the Rural Development Administration, Republic of Korea.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Gynheung An for establishing Crop Biotech Institute at Kyung Hee University. We sincerely thank Mohsin Niaz for his guidance in utilizing online resources to identify key genes involved in abiotic stresses.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sasaki, T.; Burr, B. International Rice Genome Sequencing Project: The effort to completely sequence the rice genome. Curr. Opin. Plant Biol. 2000, 3, 138–142. [Google Scholar] [CrossRef]
  2. Muthu, V.; Abbai, R.; Nallathambi, J.; Rahman, H.; Ramasamy, S.; Kambale, R.; Thulasinathan, T.; Ayyenar, B.; Muthurajan, R. Pyramiding QTLs controlling tolerance against drought, salinity and submergence in rice through marker assisted breeding. PLoS ONE 2020, 15, e0227421. [Google Scholar] [CrossRef] [PubMed]
  3. Tripathi, A.D.; Mishra, R.; Maurya, K.K.; Singh, R.B.; Wilson, D.W. Estimates for world population and global food availability for global health. In The Role of Functional Food Security in Global Health; Academic Press: Cambridge, MA, USA, 2019; pp. 3–24. [Google Scholar]
  4. Bai, S.; Yu, H.; Wang, B.; Li, J. Retrospective and perspective of rice breeding in China. J. Genet. Genom. 2018, 45, 603–612. [Google Scholar] [CrossRef]
  5. Sheehy, J.E.; Ferrer, A.B.; Mitchell, P. Harnessing photosynthesis in tomorrow’s world: Humans, crop production and poverty alleviation. In Photosynthesis. Energy from the Sun: 14th International Congress on Photosynthesis; Springer: Berlin/Heidelberg, Germany, 2008; pp. 1237–1242. [Google Scholar]
  6. Singh, R.; Singh, Y.; Xalaxo, S.; Verulkar, S.; Yadav, N.; Singh, S.; Singh, N.; Prasad, K.; Kondayya, K.; Rao, P.R. From QTL to variety-harnessing the benefits of QTLs for drought, flood and salt tolerance in mega rice varieties of India through a multi-institutional network. Plant Sci. 2016, 242, 278–287. [Google Scholar] [CrossRef] [PubMed]
  7. Panda, D.; Barik, J.; Sarkar, R.K. Recent advances of genetic resources, genes and genetic approaches for flooding tolerance in rice. Curr. Genom. 2021, 22, 41–58. [Google Scholar] [CrossRef] [PubMed]
  8. Panda, D.; Mohanty, B.; Behera, P.K.; Barik, J.; Mishra, S.S. Harnessing leaf photosynthetic traits and antioxidant defence for multiple stress tolerance in three premium indigenous rice landraces of Jeypore tract of Odisha, India. Funct. Plant Biol. 2019, 47, 99–111. [Google Scholar] [CrossRef] [PubMed]
  9. Singh, A.; Septiningsih, E.M.; Balyan, H.S.; Singh, N.K.; Rai, V. Genetics, physiological mechanisms and breeding of flood-tolerant rice (Oryza sativa L.). Plant Cell Physiol. 2017, 58, 185–197. [Google Scholar]
  10. Sri Shalini, S.; Palanivelu, K.; Ramachandran, A.; Raghavan, V. Biochar from biomass waste as a renewable carbon material for climate change mitigation in reducing greenhouse gas emissions—A review. Biomass Conver. Biorefinery 2021, 11, 2247–2267. [Google Scholar] [CrossRef]
  11. Munir, N.; Hasnain, M.; Roessner, U.; Abideen, Z. Strategies in improving plant salinity resistance and use of salinity resistant plants for economic sustainability. Crit. Rev. Environ. Sci. Technol. 2022, 52, 2150–2196. [Google Scholar] [CrossRef]
  12. Yang, C.; Sun, J. Soil salinity drives the distribution patterns and ecological functions of fungi in saline-alkali land in the Yellow River Delta, China. Front. Microbiol. 2020, 11, 594284. [Google Scholar] [CrossRef] [PubMed]
  13. Hasnain, M.; Abideen, Z.; Anthony Dias, D.; Naz, S.; Munir, N. Utilization of Saline Water Enhances Lipid Accumulation in Green Microalgae for the Sustainable Production of Biodiesel. BioEnergy Res. 2022, 16, 1026–1039. [Google Scholar] [CrossRef]
  14. Ma, Y.; Dai, X.; Xu, Y.; Luo, W.; Zheng, X.; Zeng, D.; Pan, Y.; Lin, X.; Liu, H.; Zhang, D. COLD1 confers chilling tolerance in rice. Cell 2015, 160, 1209–1221. [Google Scholar] [CrossRef] [PubMed]
  15. Takasaki, H.; Maruyama, K.; Kidokoro, S.; Ito, Y.; Fujita, Y.; Shinozaki, K.; Yamaguchi-Shinozaki, K.; Nakashima, K. The abiotic stress-responsive NAC-type transcription factor OsNAC5 regulates stress-inducible genes and stress tolerance in rice. Mol. Genet. Genom. 2010, 284, 173–183. [Google Scholar] [CrossRef]
  16. Chen, W.; Wang, W.; Peng, M.; Gong, L.; Gao, Y.; Wan, J.; Wang, S.; Shi, L.; Zhou, B.; Li, Z. Comparative and parallel genome-wide association studies for metabolic and agronomic traits in cereals. Nat. Commun. 2016, 7, 12767. [Google Scholar] [CrossRef]
  17. Huang, X.; Sang, T.; Zhao, Q.; Feng, Q.; Zhao, Y.; Li, C.; Zhu, C.; Lu, T.; Zhang, Z.; Li, M. Genome-wide association studies of 14 agronomic traits in rice landraces. Nat. Genet. 2010, 42, 961–967. [Google Scholar] [CrossRef] [PubMed]
  18. Wu, L.; Han, L.; Li, Q.; Wang, G.; Zhang, H.; Li, L. Using interactome big data to crack genetic mysteries and enhance future crop breeding. Mol. Plant 2021, 14, 77–94. [Google Scholar] [CrossRef]
  19. Zeng, D.; Tian, Z.; Rao, Y.; Dong, G.; Yang, Y.; Huang, L.; Leng, Y.; Xu, J.; Sun, C.; Zhang, G. Rational design of high-yield and superior-quality rice. Nat. Plants 2017, 3, 17031. [Google Scholar] [CrossRef]
  20. Gourdji, S.M.; Sibley, A.M.; Lobell, D.B. Global crop exposure to critical high temperatures in the reproductive period: Historical trends and future projections. Environ. Res. Lett. 2013, 8, 024041. [Google Scholar] [CrossRef]
  21. Pryor, S.; Barthelmie, R.; Schoof, J. High-resolution projections of climate-related risks for the Midwestern USA. Clim. Res. 2013, 56, 61–79. [Google Scholar] [CrossRef]
  22. Keyster, M.; Niekerk, L.-A.; Basson, G.; Carelse, M.; Bakare, O.; Ludidi, N.; Klein, A.; Mekuto, L.; Gokul, A. Decoding heavy metal stress signalling in plants: Towards improved food security and safety. Plants 2020, 9, 1781. [Google Scholar] [CrossRef]
  23. Krishnan, P.; Ramakrishnan, B.; Reddy, K.R.; Reddy, V. High-temperature effects on rice growth, yield, and grain quality. Adv. Agron. 2011, 111, 87–206. [Google Scholar]
  24. Kim, J.; Shon, J.; Lee, C.-K.; Yang, W.; Yoon, Y.; Yang, W.-H.; Kim, Y.-G.; Lee, B.-W. Relationship between grain filling duration and leaf senescence of temperate rice under high temperature. Field Crops Res. 2011, 122, 207–213. [Google Scholar] [CrossRef]
  25. Lamers, J.; Van Der Meer, T.; Testerink, C. How plants sense and respond to stressful environments. Plant Physiol. 2020, 182, 1624–1635. [Google Scholar] [CrossRef]
  26. Zhu, J.-K. Abiotic stress signaling and responses in plants. Cell 2016, 167, 313–324. [Google Scholar] [CrossRef] [PubMed]
  27. Kazan, K. Diverse roles of jasmonates and ethylene in abiotic stress tolerance. Trends Plant Sci. 2015, 20, 219–229. [Google Scholar] [CrossRef] [PubMed]
  28. Guo, M.; Liu, J.-H.; Ma, X.; Luo, D.-X.; Gong, Z.-H.; Lu, M.-H. The plant heat stress transcription factors (HSFs): Structure, regulation, and function in response to abiotic stresses. Front. Plant Sci. 2016, 7, 114. [Google Scholar] [CrossRef]
  29. Lone, J.; Khan, M.; Bhat, M.; Shikari, A.; Wani, S.; Sofi, N.; Khan, M.; Lone, R. Cold tolerance at germination and seedling stages of rice: Methods of evaluation and characterization of thirty rice genotypes under stress conditions. Int. J. Curr. Microbiol. Appl. Sci. 2018, 7, 1103–1109. [Google Scholar] [CrossRef]
  30. Janmohammadi, M.; Zolla, L.; Rinalducci, S. Low temperature tolerance in plants: Changes at the protein level. Phytochemistry 2015, 117, 76–89. [Google Scholar] [CrossRef]
  31. Dikilitas, M.; Karakas, S.; Simsek, E.; Yadav, A.N. Microbes from cold deserts and their applications in mitigation of cold stress in plants. In Microbiomes of Extreme Environments: Biodiversity and Biotechnological Applications; CRC Press: Boca Raton, FL, USA, 2021; pp. 126–152. [Google Scholar]
  32. Deng, X.; Hu, W.; Wei, S.; Zhou, S.; Zhang, F.; Han, J.; Chen, L.; Li, Y.; Feng, J.; Fang, B. TaCIPK29, a CBL-interacting protein kinase gene from wheat, confers salt stress tolerance in transgenic tobacco. PLoS ONE 2013, 8, e69881. [Google Scholar] [CrossRef]
  33. Zhang, Z.; Li, J.; Pan, Y.; Li, J.; Shi, H.; Zeng, Y.; Guo, H.; Yang, S.; Zheng, W.; Yu, J. Natural variation in CTB4a enhances rice adaptation to cold habitats. Nat. Commun. 2017, 8, 14788. [Google Scholar] [CrossRef]
  34. Sharma, P.; Sharma, N.; Deswal, R. The molecular biology of the low-temperature response in plants. Bioessays 2005, 27, 1048–1059. [Google Scholar] [CrossRef] [PubMed]
  35. Hsu, C.H.; Hsu, Y.T. Biochemical responses of rice roots to cold stress. Bot. Stud. 2019, 60, 14. [Google Scholar] [CrossRef] [PubMed]
  36. Kanneganti, V.; Gupta, A.K. Overexpression of OsiSAP8, a member of stress associated protein (SAP) gene family of rice confers tolerance to salt, drought and cold stress in transgenic tobacco and rice. Plant Mol. Biol. 2008, 66, 445–462. [Google Scholar] [CrossRef]
  37. Los, D.A.; Murata, N. Membrane fluidity and its roles in the perception of environmental signals. Biochim. Biophys. Acta Biomembr. 2004, 1666, 142–157. [Google Scholar] [CrossRef] [PubMed]
  38. Song, S.-Y.; Chen, Y.; Chen, J.; Dai, X.-Y.; Zhang, W.-H. Physiological mechanisms underlying OsNAC5-dependent tolerance of rice plants to abiotic stress. Planta 2011, 234, 331–345. [Google Scholar] [CrossRef] [PubMed]
  39. Mittal, D.; Madhyastha, D.A.; Grover, A. Genome-wide transcriptional profiles during temperature and oxidative stress reveal coordinated expression patterns and overlapping regulons in rice. PLoS ONE 2012, 7, e40899. [Google Scholar] [CrossRef]
  40. Pamplona, R. Advanced lipoxidation end-products. Chem. Biol. Interact. 2011, 192, 14–20. [Google Scholar] [CrossRef]
  41. Bonnecarrère, V.; Borsani, O.; Díaz, P.; Capdevielle, F.; Blanco, P.; Monza, J. Response to photoxidative stress induced by cold in japonica rice is genotype dependent. Plant Sci. 2011, 180, 726–732. [Google Scholar] [CrossRef]
  42. Kim, S.-I.; Tai, T.H. Evaluation of seedling cold tolerance in rice cultivars: A comparison of visual ratings and quantitative indicators of physiological changes. Euphytica 2011, 178, 437–447. [Google Scholar] [CrossRef]
  43. Xie, G.; Kato, H.; Sasaki, K.; Imai, R. A cold-induced thioredoxin h of rice, OsTrx23, negatively regulates kinase activities of OsMPK3 and OsMPK6 in vitro. FEBS Lett. 2009, 583, 2734–2738. [Google Scholar] [CrossRef]
  44. Sato, Y.; Masuta, Y.; Saito, K.; Murayama, S.; Ozawa, K. Enhanced chilling tolerance at the booting stage in rice by transgenic overexpression of the ascorbate peroxidase gene, OsAPXa. Plant Cell Rep. 2011, 30, 399–406. [Google Scholar] [CrossRef]
  45. Wang, Y.; Cui, Y.; Hu, G.; Wang, X.; Chen, H.; Shi, Q.; Xiang, J.; Zhang, Y.; Zhu, D.; Zhang, Y. Reduced bioactive gibberellin content in rice seeds under low temperature leads to decreased sugar consumption and low seed germination rates. Plant Physiol. Biochem. 2018, 133, 1–10. [Google Scholar] [CrossRef] [PubMed]
  46. Yuanyuan, M.; Yali, Z.; Jiang, L.; Hongbo, S. Roles of plant soluble sugars and their responses to plant cold stress. Afr. J. Biotechnol. 2009, 8, 2004–2010. [Google Scholar]
  47. Shima, S.; Matsui, H.; Tahara, S.; Imai, R. Biochemical characterization of rice trehalose-6-phosphate phosphatases supports distinctive functions of these plant enzymes. FEBS J. 2007, 274, 1192–1201. [Google Scholar] [CrossRef] [PubMed]
  48. Oliver, S.N.; Dennis, E.S.; Dolferus, R. ABA regulates apoplastic sugar transport and is a potential signal for cold-induced pollen sterility in rice. Plant Cell Physiol. 2007, 48, 1319–1330. [Google Scholar] [CrossRef] [PubMed]
  49. Sakata, T.; Oda, S.; Tsunaga, Y.; Shomura, H.; Kawagishi-Kobayashi, M.; Aya, K.; Saeki, K.; Endo, T.; Nagano, K.; Kojima, M. Reduction of gibberellin by low temperature disrupts pollen development in rice. Plant Physiol. 2014, 164, 2011–2019. [Google Scholar] [CrossRef] [PubMed]
  50. Kishor, P.K.; Sangam, S.; Amrutha, R.; Laxmi, P.S.; Naidu, K.; Rao, K.S.; Rao, S.; Reddy, K.; Theriappan, P.; Sreenivasulu, N. Regulation of proline biosynthesis, degradation, uptake and transport in higher plants: Its implications in plant growth and abiotic stress tolerance. Curr. Sci. 2005, 88, 424–438. [Google Scholar]
  51. Kandpal, R.P.; Rao, N.A. Alterations in the biosynthesis of proteins and nucleic acids in finger millet (Eleucine coracana) seedlings during water stress and the effect of proline on protein biosynthesis. Plant Sci. 1985, 40, 73–79. [Google Scholar] [CrossRef]
  52. Venekamp, J. Regulation of cytosol acidity in plants under conditions of drought. Physiol. Plant. 1989, 76, 112–117. [Google Scholar] [CrossRef]
  53. Schobert, B.; Tschesche, H. Unusual solution properties of proline and its interaction with proteins. Biochim. Biophys. Acta Gen. Subj. 1978, 541, 270–277. [Google Scholar] [CrossRef]
  54. Liu, K.; Wang, L.; Xu, Y.; Chen, N.; Ma, Q.; Li, F.; Chong, K. Overexpression of OsCOIN, a putative cold inducible zinc finger protein, increased tolerance to chilling, salt and drought, and enhanced proline level in rice. Planta 2007, 226, 1007–1016. [Google Scholar] [CrossRef] [PubMed]
  55. Mittler, R.; Blumwald, E. The roles of ROS and ABA in systemic acquired acclimation. Plant Cell 2015, 27, 64–70. [Google Scholar] [CrossRef]
  56. Mega, R.; Meguro-Maoka, A.; Endo, A.; Shimosaka, E.; Murayama, S.; Nambara, E.; Seo, M.; Kanno, Y.; Abrams, S.R.; Sato, Y. Sustained low abscisic acid levels increase seedling vigor under cold stress in rice (Oryza sativa L.). Sci. Rep. 2015, 5, 13819. [Google Scholar] [CrossRef] [PubMed]
  57. Ma, Y.; Szostkiewicz, I.; Korte, A.; Moes, D.; Yang, Y.; Christmann, A.; Grill, E. Regulators of PP2C phosphatase activity function as abscisic acid sensors. Science 2009, 324, 1064–1068. [Google Scholar] [CrossRef] [PubMed]
  58. Park, S.-Y.; Fung, P.; Nishimura, N.; Jensen, D.R.; Fujii, H.; Zhao, Y.; Lumba, S.; Santiago, J.; Rodrigues, A.; Chow, T.-F.; et al. Abscisic acid inhibits type 2C protein phosphatases via the PYR/PYL family of START proteins. Science 2009, 324, 1068–1071. [Google Scholar] [CrossRef] [PubMed]
  59. Fujii, H.; Chinnusamy, V.; Rodrigues, A.; Rubio, S.; Antoni, R.; Park, S.-Y.; Cutler, S.R.; Sheen, J.; Rodriguez, P.L.; Zhu, J.-K. In vitro reconstitution of an abscisic acid signalling pathway. Nature 2009, 462, 660–664. [Google Scholar] [CrossRef]
  60. Kim, H.; Hwang, H.; Hong, J.-W.; Lee, Y.-N.; Ahn, I.P.; Yoon, I.S.; Yoo, S.-D.; Lee, S.; Lee, S.C.; Kim, B.-G. A rice orthologue of the ABA receptor, OsPYL/RCAR5, is a positive regulator of the ABA signal transduction pathway in seed germination and early seedling growth. J. Exp. Bot. 2012, 63, 1013–1024. [Google Scholar] [CrossRef]
  61. Sah, S.K.; Reddy, K.R.; Li, J. Abscisic acid and abiotic stress tolerance in crop plants. Front. Plant Sci. 2016, 7, 571. [Google Scholar] [CrossRef]
  62. Tian, X.; Wang, Z.; Li, X.; Lv, T.; Liu, H.; Wang, L.; Niu, H.; Bu, Q. Characterization and functional analysis of pyrabactin resistance-like abscisic acid receptor family in rice. Rice 2015, 8, 28. [Google Scholar] [CrossRef]
  63. Rashid, M.; Guangyuan, H.; Guangxiao, Y.; Hussain, J.; Xu, Y. AP2/ERF transcription factor in rice: Genome-wide canvas and syntenic relationships between monocots and eudicots. Evol. Bioinform. 2012, 8, 321–355. [Google Scholar] [CrossRef]
  64. Chinnusamy, V.; Zhu, J.; Zhu, J.-K. Cold stress regulation of gene expression in plants. Trends Plant Sci. 2007, 12, 444–451. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, Y.; Chen, C.; Jin, X.-F.; Xiong, A.-S.; Peng, R.-H.; Hong, Y.-H.; Yao, Q.-H.; Chen, J.-M. Expression of a rice DREB1 gene, OsDREB1D, enhances cold and high-salt tolerance in transgenic Arabidopsis. BMB Rep. 2009, 42, 486–492. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, Z.; Li, J.; Li, F.; Liu, H.; Yang, W.; Chong, K.; Xu, Y. OsMAPK3 phosphorylates OsbHLH002/OsICE1 and inhibits its ubiquitination to activate OsTPP1 and enhances rice chilling tolerance. Dev. Cell 2017, 43, 731–743.e5. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, Q.; Guan, Y.; Wu, Y.; Chen, H.; Chen, F.; Chu, C. Overexpression of a rice OsDREB1F gene increases salt, drought, and low temperature tolerance in both Arabidopsis and rice. Plant Mol. Biol. 2008, 67, 589–602. [Google Scholar] [CrossRef] [PubMed]
  68. Wang, D.-Z.; Jin, Y.-N.; Ding, X.-H.; Wang, W.-J.; Zhai, S.-S.; Bai, L.-P.; Guo, Z.-F. Gene regulation and signal transduction in the ICE–CBF–COR signaling pathway during cold stress in plants. Biochemistry 2017, 82, 1103–1117. [Google Scholar] [CrossRef]
  69. Mao, D.; Xin, Y.; Tan, Y.; Hu, X.; Bai, J.; Liu, Z.-y.; Yu, Y.; Li, L.; Peng, C.; Fan, T. Natural variation in the HAN1 gene confers chilling tolerance in rice and allowed adaptation to a temperate climate. Proc. Natl. Acad. Sci. USA 2019, 116, 3494–3501. [Google Scholar] [CrossRef]
  70. Miao, C.; Xiao, L.; Hua, K.; Zou, C.; Zhao, Y.; Bressan, R.A.; Zhu, J.-K. Mutations in a subfamily of abscisic acid receptor genes promote rice growth and productivity. Proc. Natl. Acad. Sci. USA 2018, 115, 6058–6063. [Google Scholar] [CrossRef]
  71. Fujino, K.; Sekiguchi, H.; Matsuda, Y.; Sugimoto, K.; Ono, K.; Yano, M. Molecular identification of a major quantitative trait locus, qLTG3–1, controlling low-temperature germinability in rice. Proc. Natl. Acad. Sci. USA 2008, 105, 12623–12628. [Google Scholar] [CrossRef]
  72. Kim, S.-I.; Andaya, V.C.; Tai, T.H. Cold sensitivity in rice (Oryza sativa L.) is strongly correlated with a naturally occurring I99V mutation in the multifunctional glutathione transferase isoenzyme GSTZ2. Biochem. J. 2011, 435, 373–380. [Google Scholar] [CrossRef]
  73. Lu, G.; Wu, F.Q.; Wu, W.; Wang, H.J.; Zheng, X.M.; Zhang, Y.; Chen, X.; Zhou, K.; Jin, M.; Cheng, Z. Rice LTG 1 is involved in adaptive growth and fitness under low ambient temperature. Plant J. 2014, 78, 468–480. [Google Scholar] [CrossRef]
  74. Saito, K.; Hayano-Saito, Y.; Kuroki, M.; Sato, Y. Map-based cloning of the rice cold tolerance gene Ctb1. Plant Sci. 2010, 179, 97–102. [Google Scholar] [CrossRef]
  75. Liu, C.; Ou, S.; Mao, B.; Tang, J.; Wang, W.; Wang, H.; Cao, S.; Schläppi, M.R.; Zhao, B.; Xiao, G. Early selection of bZIP73 facilitated adaptation of japonica rice to cold climates. Nat. Commun. 2018, 9, 3302. [Google Scholar] [CrossRef] [PubMed]
  76. Kusumi, K.; Sakata, C.; Nakamura, T.; Kawasaki, S.; Yoshimura, A.; Iba, K. A plastid protein NUS1 is essential for build-up of the genetic system for early chloroplast development under cold stress conditions. Plant J. 2011, 68, 1039–1050. [Google Scholar] [CrossRef] [PubMed]
  77. Sun, J.; Zheng, T.; Yu, J.; Wu, T.; Wang, X.; Chen, G.; Tian, Y.; Zhang, H.; Wang, Y.; Terzaghi, W. TSV, a putative plastidic oxidoreductase, protects rice chloroplasts from cold stress during development by interacting with plastidic thioredoxin Z. New Phytol. 2017, 215, 240–255. [Google Scholar] [CrossRef]
  78. Liu, X.; Lan, J.; Huang, Y.; Cao, P.; Zhou, C.; Ren, Y.; He, N.; Liu, S.; Tian, Y.; Nguyen, T. WSL5, a pentatricopeptide repeat protein, is essential for chloroplast biogenesis in rice under cold stress. J. Exp. Bot. 2018, 69, 3949–3961. [Google Scholar] [CrossRef]
  79. Ge, Y.; Han, J.; Zhou, G.; Xu, Y.; Ding, Y.; Shi, M.; Guo, C.; Wu, G. Silencing of miR156 confers enhanced resistance to brown planthopper in rice. Planta 2018, 248, 813–826. [Google Scholar] [CrossRef]
  80. Lee, B.-H.; Won, S.-H.; Lee, H.-S.; Miyao, M.; Chung, W.-I.; Kim, I.-J.; Jo, J. Expression of the chloroplast-localized small heat shock protein by oxidative stress in rice. Gene 2000, 245, 283–290. [Google Scholar] [CrossRef]
  81. Lin, M.-y.; Chai, K.-h.; Ko, S.-s.; Kuang, L.-y.; Lur, H.-S.; Charng, Y.-y. A positive feedback loop between HEAT SHOCK PROTEIN101 and HEAT STRESS-ASSOCIATED 32-KD PROTEIN modulates long-term acquired thermotolerance illustrating diverse heat stress responses in rice varieties. Plant Physiol. 2014, 164, 2045–2053. [Google Scholar] [CrossRef]
  82. Li, X.-M.; Chao, D.-Y.; Wu, Y.; Huang, X.; Chen, K.; Cui, L.-G.; Su, L.; Ye, W.-W.; Chen, H.; Chen, H.-C. Natural alleles of a proteasome α2 subunit gene contribute to thermotolerance and adaptation of African rice. Nat. Genet. 2015, 47, 827–833. [Google Scholar] [CrossRef]
  83. Wang, D.; Qin, B.; Li, X.; Tang, D.; Zhang, Y.e.; Cheng, Z.; Xue, Y. Nucleolar DEAD-box RNA helicase TOGR1 regulates thermotolerant growth as a pre-rRNA chaperone in rice. PLoS Genet. 2016, 12, e1005844. [Google Scholar] [CrossRef]
  84. Tang, Y.; Gao, C.-C.; Gao, Y.; Yang, Y.; Shi, B.; Yu, J.-L.; Lyu, C.; Sun, B.-F.; Wang, H.-L.; Xu, Y. OsNSUN2-mediated 5-methylcytosine mRNA modification enhances rice adaptation to high temperature. Develop. Cell 2020, 53, 272–286.e277. [Google Scholar] [CrossRef] [PubMed]
  85. Zheng, K.; Zhao, J.; Lin, D.; Chen, J.; Xu, J.; Zhou, H.; Teng, S.; Dong, Y. The rice TCM5 gene encoding a novel Deg protease protein is essential for chloroplast development under high temperatures. Rice 2016, 9, 13. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, B.; Wu, S.; Zhang, Y.e.; Xu, T.; Guo, F.; Tang, H.; Li, X.; Wang, P.; Qian, W.; Xue, Y. A high temperature-dependent mitochondrial lipase EXTRA GLUME1 promotes floral phenotypic robustness against temperature fluctuation in rice (Oryza sativa L.). PLoS Genet. 2016, 12, e1006152. [Google Scholar] [CrossRef] [PubMed]
  87. Lo, S.F.; Cheng, M.L.; Hsing, Y.i.C.; Chen, Y.S.; Lee, K.W.; Hong, Y.F.; Hsiao, Y.; Hsiao, A.S.; Chen, P.J.; Wong, L.I. Rice Big Grain 1 promotes cell division to enhance organ development, stress tolerance and grain yield. Plant Biotechnol. J. 2020, 18, 1969–1983. [Google Scholar] [CrossRef] [PubMed]
  88. Qiao, B.; Zhang, Q.; Liu, D.; Wang, H.; Yin, J.; Wang, R.; He, M.; Cui, M.; Shang, Z.; Wang, D. A calcium-binding protein, rice annexin OsANN1, enhances heat stress tolerance by modulating the production of H2O2. J. Exp. Bot. 2015, 66, 5853–5866. [Google Scholar] [CrossRef]
  89. Uga, Y.; Sugimoto, K.; Ogawa, S.; Rane, J.; Ishitani, M.; Hara, N.; Kitomi, Y.; Inukai, Y.; Ono, K.; Kanno, N. Control of root system architecture by DEEPER ROOTING 1 increases rice yield under drought conditions. Nat. Genet. 2013, 45, 1097–1102. [Google Scholar] [CrossRef]
  90. Li, J.; Li, Y.; Yin, Z.; Jiang, J.; Zhang, M.; Guo, X.; Ye, Z.; Zhao, Y.; Xiong, H.; Zhang, Z. Os ASR 5 enhances drought tolerance through a stomatal closure pathway associated with ABA and H2O2 signalling in rice. Plant Biotechnol. J. 2017, 15, 183–196. [Google Scholar] [CrossRef] [PubMed]
  91. Huang, X.-Y.; Chao, D.-Y.; Gao, J.-P.; Zhu, M.-Z.; Shi, M.; Lin, H.-X. A previously unknown zinc finger protein, DST, regulates drought and salt tolerance in rice via stomatal aperture control. Genes Dev. 2009, 23, 1805–1817. [Google Scholar] [CrossRef]
  92. Fu, J.; Wu, H.; Ma, S.; Xiang, D.; Liu, R.; Xiong, L. OsJAZ1 attenuates drought resistance by regulating JA and ABA signaling in rice. Front. Plant Sci. 2017, 8, 2108. [Google Scholar] [CrossRef]
  93. Caine, R.S.; Yin, X.; Sloan, J.; Harrison, E.L.; Mohammed, U.; Fulton, T.; Biswal, A.K.; Dionora, J.; Chater, C.C.; Coe, R.A. Rice with reduced stomatal density conserves water and has improved drought tolerance under future climate conditions. New Phytol. 2019, 221, 371–384. [Google Scholar] [CrossRef]
  94. Wei, S.; Hu, W.; Deng, X.; Zhang, Y.; Liu, X.; Zhao, X.; Luo, Q.; Jin, Z.; Li, Y.; Zhou, S. A rice calcium-dependent protein kinase OsCPK9 positively regulates drought stress tolerance and spikelet fertility. BMC Plant Biol. 2014, 14, 133. [Google Scholar] [CrossRef] [PubMed]
  95. Du, H.; Liu, L.; You, L.; Yang, M.; He, Y.; Li, X.; Xiong, L. Characterization of an inositol 1, 3, 4-trisphosphate 5/6-kinase gene that is essential for drought and salt stress responses in rice. Plant Mol. Biol. 2011, 77, 547–563. [Google Scholar] [CrossRef] [PubMed]
  96. Diédhiou, C.J.; Popova, O.V.; Dietz, K.-J.; Golldack, D. The SNF1-type serine-threonine protein kinase SAPK4 regulates stress-responsive gene expression in rice. BMC Plant Biol. 2008, 8, 49. [Google Scholar] [CrossRef] [PubMed]
  97. Jeong, M.-J.; Lee, S.-K.; Kim, B.-G.; Kwon, T.-R.; Cho, W.-S.; Park, Y.-T.; Lee, J.-O.; Kwon, H.-B.; Byun, M.-O.; Park, S.-C. A rice (Oryza sativa L.) MAP kinase gene, OsMAPK44, is involved in response to abiotic stresses. Plant Cell Tissue Organ Cult. 2006, 85, 151–160. [Google Scholar] [CrossRef]
  98. Guan, Q.; Ma, H.; Wang, Z.; Wang, Z.; Bu, Q.; Liu, S. A rice LSD1-like-type ZFP gene OsLOL5 enhances saline-alkaline tolerance in transgenic Arabidopsis thaliana, yeast and rice. BMC Genom. 2016, 17, 142. [Google Scholar] [CrossRef]
  99. Hasthanasombut, S.; Ntui, V.; Supaibulwatana, K.; Mii, M.; Nakamura, I. Expression of Indica rice OsBADH1 gene under salinity stress in transgenic tobacco. Plant Biotechnol. Rep. 2010, 4, 75–83. [Google Scholar] [CrossRef]
  100. Obata, T.; Kitamoto, H.K.; Nakamura, A.; Fukuda, A.; Tanaka, Y. Rice shaker potassium channel OsKAT1 confers tolerance to salinity stress on yeast and rice cells. Plant Physiol. 2007, 144, 1978–1985. [Google Scholar] [CrossRef]
  101. Horie, T.; Sugawara, M.; Okada, T.; Taira, K.; Kaothien-Nakayama, P.; Katsuhara, M.; Shinmyo, A.; Nakayama, H. Rice sodium-insensitive potassium transporter, OsHAK5, confers increased salt tolerance in tobacco BY2 cells. J. Biosci. Bioeng. 2011, 111, 346–356. [Google Scholar] [CrossRef]
  102. Liu, S.; Zheng, L.; Xue, Y.; Zhang, Q.; Wang, L.; Shou, H. Overexpression of OsVP1 and OsNHX1 increases tolerance to drought and salinity in rice. J. Plant Biol. 2010, 53, 444–452. [Google Scholar] [CrossRef]
  103. Dikilitas, M.; Simsek, E.; Roychoudhury, A. Role of proline and glycine betaine in overcoming abiotic stresses. In Protective Chemical Agents in the Amelioration of Plant Abiotic Stress: Biochemical and Molecular Perspectives; John Wiley & Sons: Hoboken, NJ, USA, 2020; pp. 1–23. [Google Scholar]
  104. Wang, B.; Xie, G.; Liu, Z.; He, R.; Han, J.; Huang, S.; Liu, L.; Cheng, X. Mutagenesis reveals that the OsPPa6 gene is required for enhancing the alkaline tolerance in rice. Front. Plant Sci. 2019, 10, 759. [Google Scholar] [CrossRef]
  105. Jing, P.; Zou, J.; Kong, L.; Hu, S.; Wang, B.; Yang, J.; Xie, G. OsCCD1, a novel small calcium-binding protein with one EF-hand motif, positively regulates osmotic and salt tolerance in rice. Plant Sci. 2016, 247, 104–114. [Google Scholar] [CrossRef] [PubMed]
  106. Gao, S.; Song, T.; Han, J.; He, M.; Zhang, Q.; Zhu, Y.; Zhu, Z. A calcium-dependent lipid binding protein, OsANN10, is a negative regulator of osmotic stress tolerance in rice. Plant Sci. 2020, 293, 110420. [Google Scholar] [CrossRef]
  107. Zhao, H.; Li, Z.; Wang, Y.; Wang, J.; Xiao, M.; Liu, H.; Quan, R.; Zhang, H.; Huang, R.; Zhu, L. Cellulose synthase-like protein OsCSLD4 plays an important role in the response of rice to salt stress by mediating abscisic acid biosynthesis to regulate osmotic stress tolerance. Plant Biotechnol. J. 2022, 20, 468–484. [Google Scholar] [CrossRef] [PubMed]
  108. Liu, Q.; Ding, J.; Huang, W.; Yu, H.; Wu, S.; Li, W.; Mao, X.; Chen, W.; Xing, J.; Li, C. OsPP65 negatively regulates osmotic and salt stress responses through regulating phytohormone and raffinose family oligosaccharide metabolic pathways in rice. Rice 2022, 15, 34. [Google Scholar] [CrossRef] [PubMed]
  109. Maity, K.; Heumann, J.M.; McGrath, A.P.; Kopcho, N.J.; Hsu, P.-K.; Lee, C.-W.; Mapes, J.H.; Garza, D.; Krishnan, S.; Morgan, G.P. Cryo-EM structure of OSCA1. 2 from Oryza sativa elucidates the mechanical basis of potential membrane hyperosmolality gating. Proc. Natl. Acad. Sci. USA 2019, 116, 14309–14318. [Google Scholar] [CrossRef]
  110. Fukao, T.; Xu, K.; 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] [PubMed]
  111. Yang, C.; Ma, B.; He, S.-J.; Xiong, Q.; Duan, K.-X.; Yin, C.-C.; Chen, H.; Lu, X.; Chen, S.-Y.; Zhang, J.-S. MAOHUZI6/ETHYLENE INSENSITIVE3-LIKE1 and ETHYLENE INSENSITIVE3-LIKE2 regulate ethylene response of roots and coleoptiles and negatively affect salt tolerance in rice. Plant Physiol. 2015, 169, 148–165. [Google Scholar] [CrossRef]
  112. Nagai, K.; Mori, Y.; Ishikawa, S.; Furuta, T.; Gamuyao, R.; Niimi, Y.; Hobo, T.; Fukuda, M.; Kojima, M.; Takebayashi, Y. Antagonistic regulation of the gibberellic acid response during stem growth in rice. Nature 2020, 584, 109–114. [Google Scholar] [CrossRef]
  113. Kobata, T.; Uemuki, N. High temperatures during the grain-filling period do not reduce the potential grain dry matter increase of rice. Agron. J. 2004, 96, 406–414. [Google Scholar] [CrossRef]
  114. Tashiro, T.; Wardlaw, I. The effect of high temperature on the accumulation of dry matter, carbon and nitrogen in the kernel of rice. Funct. Plant Biol. 1991, 18, 259–265. [Google Scholar] [CrossRef]
  115. Zou, J.; Liu, A.; Chen, X.; Zhou, X.; Gao, G.; Wang, W.; Zhang, X. Expression analysis of nine rice heat shock protein genes under abiotic stresses and ABA treatment. J. Plant Physiol. 2009, 166, 851–861. [Google Scholar] [CrossRef] [PubMed]
  116. Singh, A.; Singh, U.; Mittal, D.; Grover, A. Genome-wide analysis of rice ClpB/HSP100, ClpC and ClpD genes. BMC Genom. 2010, 11, 95. [Google Scholar] [CrossRef] [PubMed]
  117. Wang, C.; Zhang, Q.; Shou, H.-X. Identification and expression analysis of OsHsfs in rice. J. Zhejiang Univ. Sci. B 2009, 10, 291–300. [Google Scholar] [CrossRef] [PubMed]
  118. Cheng, Q.; Zhou, Y.; Liu, Z.; Zhang, L.; Song, G.; Guo, Z.; Wang, W.; Qu, X.; Zhu, Y.; Yang, D. An alternatively spliced heat shock transcription factor, Os HSFA 2dI, functions in the heat stress-induced unfolded protein response in rice. Plant Biol. 2015, 17, 419–429. [Google Scholar] [CrossRef]
  119. Yokotani, N.; Ichikawa, T.; Kondou, Y.; Matsui, M.; Hirochika, H.; Iwabuchi, M.; Oda, K. Expression of rice heat stress transcription factor OsHsfA2e enhances tolerance to environmental stresses in transgenic Arabidopsis. Planta 2008, 227, 957–967. [Google Scholar] [CrossRef]
  120. Wang, H.; Bian, M.; Yang, Z.; Lin, C.; Shi, W. Preliminary functional analysis of the isoforms of OsHsfA2a (Oryza sativa L.) generated by alternative splicing. Plant Mol. Biol. Rep. 2013, 31, 38–46. [Google Scholar] [CrossRef]
  121. Qin, Q.-l.; Liu, J.-g.; Zhang, Z.; Peng, R.-h.; Xiong, A.-s.; Yao, Q.-h.; Chen, J.-m. Isolation, optimization, and functional analysis of the cDNA encoding transcription factor OsDREB1B in Oryza Sativa L. Mol. Breed. 2007, 19, 329–340. [Google Scholar] [CrossRef]
  122. Wu, X.; Shiroto, Y.; Kishitani, S.; Ito, Y.; Toriyama, K. Enhanced heat and drought tolerance in transgenic rice seedlings overexpressing OsWRKY11 under the control of HSP101 promoter. Plant Cell Rep. 2009, 28, 21–30. [Google Scholar] [CrossRef]
  123. Ambavaram, M.M.; Basu, S.; Krishnan, A.; Ramegowda, V.; Batlang, U.; Rahman, L.; Baisakh, N.; Pereira, A. Coordinated regulation of photosynthesis in rice increases yield and tolerance to environmental stress. Nat. Commun. 2014, 5, 5302. [Google Scholar] [CrossRef]
  124. Fang, Y.; Liao, K.; Du, H.; Xu, Y.; Song, H.; Li, X.; Xiong, L. A stress-responsive NAC transcription factor SNAC3 confers heat and drought tolerance through modulation of reactive oxygen species in rice. J. Exp. Bot. 2015, 66, 6803–6817. [Google Scholar] [CrossRef]
  125. El-Kereamy, A.; Bi, Y.-M.; Ranathunge, K.; Beatty, P.H.; Good, A.G.; Rothstein, S.J. The rice R2R3-MYB transcription factor OsMYB55 is involved in the tolerance to high temperature and modulates amino acid metabolism. PLoS ONE 2012, 7, e52030. [Google Scholar] [CrossRef] [PubMed]
  126. Lim, S.D.; Cho, H.Y.; Park, Y.C.; Ham, D.J.; Lee, J.K.; Jang, C.S. The rice RING finger E3 ligase, OsHCI1, drives nuclear export of multiple substrate proteins and its heterogeneous overexpression enhances acquired thermotolerance. J. Exp. Bot. 2013, 64, 2899–2914. [Google Scholar] [CrossRef] [PubMed]
  127. Liu, J.; Zhang, C.; Wei, C.; Liu, X.; Wang, M.; Yu, F.; Xie, Q.; Tu, J. The RING finger ubiquitin E3 ligase OsHTAS enhances heat tolerance by promoting H2O2-induced stomatal closure in rice. Plant Physiol. 2016, 170, 429–443. [Google Scholar] [CrossRef] [PubMed]
  128. Park, S.; Moon, J.-C.; Park, Y.C.; Kim, J.-H.; Kim, D.S.; Jang, C.S. Molecular dissection of the response of a rice leucine-rich repeat receptor-like kinase (LRR-RLK) gene to abiotic stresses. J. Plant Physiol. 2014, 171, 1645–1653. [Google Scholar] [CrossRef]
  129. Chen, K.; Guo, T.; Li, X.-M.; Zhang, Y.-M.; Yang, Y.-B.; Ye, W.-W.; Dong, N.-Q.; Shi, C.-L.; Kan, Y.; Xiang, Y.-H. Translational regulation of plant response to high temperature by a dual-function tRNAHis guanylyltransferase in rice. Mol. Plant 2019, 12, 1123–1142. [Google Scholar] [CrossRef]
  130. Yang, Y.; Xu, J.; Huang, L.; Leng, Y.; Dai, L.; Rao, Y.; Chen, L.; Wang, Y.; Tu, Z.; Hu, J. PGL, encoding chlorophyllide a oxygenase 1, impacts leaf senescence and indirectly affects grain yield and quality in rice. J. Exp. Bot. 2016, 67, 1297–1310. [Google Scholar] [CrossRef]
  131. Chou, T.-S.; Chao, Y.-Y.; Kao, C.H. Involvement of hydrogen peroxide in heat shock-and cadmium-induced expression of ascorbate peroxidase and glutathione reductase in leaves of rice seedlings. J. Plant Physiol. 2012, 169, 478–486. [Google Scholar] [CrossRef]
  132. Koh, S.; Lee, S.-C.; Kim, M.-K.; Koh, J.H.; Lee, S.; An, G.; Choe, S.; Kim, S.-R. T-DNA tagged knockout mutation of rice OsGSK1, an orthologue of Arabidopsis BIN2, with enhanced tolerance to various abiotic stresses. Plant Mol. Biol. 2007, 65, 453–466. [Google Scholar] [CrossRef]
  133. Dong, N.-Q.; Sun, Y.; Guo, T.; Shi, C.-L.; Zhang, Y.-M.; Kan, Y.; Xiang, Y.-H.; Zhang, H.; Yang, Y.-B.; Li, Y.-C. UDP-glucosyltransferase regulates grain size and abiotic stress tolerance associated with metabolic flux redirection in rice. Nat. Commun. 2020, 11, 2629. [Google Scholar] [CrossRef]
  134. Jeandet, P.; Formela-Luboińska, M.; Labudda, M.; Morkunas, I. The role of sugars in plant responses to stress and their regulatory function during development. Int. J. Mol. Sci. 2022, 23, 5161. [Google Scholar] [CrossRef]
  135. Dallagnol, L.; Rodrigues, F.; Chaves, A.d.M.; Vale, F.; DaMatta, F. Photosynthesis and sugar concentration are impaired by the defective active silicon uptake in rice plants infected with Bipolaris oryzae. Plant Pathol. 2013, 62, 120–129. [Google Scholar] [CrossRef]
  136. Paul, M.J.; Foyer, C.H. Sink regulation of photosynthesis. J. Exp. Bot. 2001, 52, 1383–1400. [Google Scholar] [CrossRef] [PubMed]
  137. Xalxo, R.; Yadu, B.; Chandra, J.; Chandrakar, V.; Keshavkant, S. Alteration in carbohydrate metabolism modulates thermotolerance of plant under heat stress. In Heat Stress Tolerance in Plants: Physiological, Molecular and Genetic Perspectives; John Wiley & Sons: Hoboken, NJ, USA, 2020; pp. 77–115. [Google Scholar]
  138. Kuhn, C.; Hajirezaei, M.-R.; Fernie, A.R.; Roessner-Tunali, U.; Czechowski, T.; Hirner, B.; Frommer, W.B. The sucrose transporter StSUT1 localizes to sieve elements in potato tuber phloem and influences tuber physiology and development. Plant Physiol. 2003, 131, 102–113. [Google Scholar] [CrossRef] [PubMed]
  139. Miyazaki, M.; Araki, M.; Okamura, K.; Ishibashi, Y.; Yuasa, T.; Iwaya-Inoue, M. Assimilate translocation and expression of sucrose transporter, OsSUT1, contribute to high-performance ripening under heat stress in the heat-tolerant rice cultivar Genkitsukushi. J. Plant Physiol. 2013, 170, 1579–1584. [Google Scholar] [CrossRef] [PubMed]
  140. Ren, Y.; Huang, Z.; Jiang, H.; Wang, Z.; Wu, F.; Xiong, Y.; Yao, J. A heat stress responsive NAC transcription factor heterodimer plays key roles in rice grain filling. J. Exp. Bot. 2021, 72, 2947–2964. [Google Scholar] [CrossRef]
  141. Stephen, K.; Beena, R.; Kiran, A.; Shanija, S.; Saravanan, R. Changes in physiological traits and expression of key genes involved in sugar signaling pathway in rice under high temperature stress. 3 Biotech 2022, 12, 183. [Google Scholar] [CrossRef]
  142. She, K.-C.; Kusano, H.; Koizumi, K.; Yamakawa, H.; Hakata, M.; Imamura, T.; Fukuda, M.; Naito, N.; Tsurumaki, Y.; Yaeshima, M. A novel factor FLOURY ENDOSPERM2 is involved in regulation of rice grain size and starch quality. Plant Cell 2010, 22, 3280–3294. [Google Scholar] [CrossRef]
  143. Dhatt, B.K.; Paul, P.; Sandhu, J.; Hussain, W.; Irvin, L.; Zhu, F.; Adviento-Borbe, M.A.; Lorence, A.; Staswick, P.; Yu, H. Allelic variation in rice Fertilization Independent Endosperm 1 contributes to grain width under high night temperature stress. New Phytol. 2021, 229, 335–350. [Google Scholar] [CrossRef]
  144. He, S.; Tan, L.; Hu, Z.; Chen, G.; Wang, G.; Hu, T. Molecular characterization and functional analysis by heterologous expression in E. coli under diverse abiotic stresses for OsLEA5, the atypical hydrophobic LEA protein from Oryza sativa L. Mol. Genet. Genom. 2012, 287, 39–54. [Google Scholar] [CrossRef]
  145. Huang, J.; Wang, M.-M.; Jiang, Y.; Bao, Y.-M.; Huang, X.; Sun, H.; Xu, D.-Q.; Lan, H.-X.; Zhang, H.-S. Expression analysis of rice A20/AN1-type zinc finger genes and characterization of ZFP177 that contributes to temperature stress tolerance. Gene 2008, 420, 135–144. [Google Scholar] [CrossRef]
  146. Hsu, K.-H.; Liu, C.-C.; Wu, S.-J.; Kuo, Y.-Y.; Lu, C.-A.; Wu, C.-R.; Lian, P.-J.; Hong, C.-Y.; Ke, Y.-T.; Huang, J.-H. Expression of a gene encoding a rice RING zinc-finger protein, OsRZFP34, enhances stomata opening. Plant Mol. Biol. 2014, 86, 125–137. [Google Scholar] [CrossRef]
  147. Cui, Y.; Lu, S.; Li, Z.; Cheng, J.; Hu, P.; Zhu, T.; Wang, X.; Jin, M.; Wang, X.; Li, L. CYCLIC NUCLEOTIDE-GATED ION CHANNELs 14 and 16 promote tolerance to heat and chilling in rice. Plant Physiol. 2020, 183, 1794–1808. [Google Scholar] [CrossRef] [PubMed]
  148. Xu, Y.; Zhang, L.; Ou, S.; Wang, R.; Wang, Y.; Chu, C.; Yao, S. Natural variations of SLG1 confer high-temperature tolerance in indica rice. Nat. Commun. 2020, 11, 5441. [Google Scholar] [CrossRef] [PubMed]
  149. Sahi, C.; Agarwal, M.; Singh, A.; Grover, A. Molecular characterization of a novel isoform of rice (Oryza sativa L.) glycine rich-RNA binding protein and evidence for its involvement in high temperature stress response. Plant Sci. 2007, 173, 144–155. [Google Scholar] [CrossRef]
  150. Cortijo, S.; Charoensawan, V.; Brestovitsky, A.; Buning, R.; Ravarani, C.; Rhodes, D.; van Noort, J.; Jaeger, K.E.; Wigge, P.A. Transcriptional regulation of the ambient temperature response by H2A. Z nucleosomes and HSF1 transcription factors in Arabidopsis. Mol. Plant 2017, 10, 1258–1273. [Google Scholar] [CrossRef]
  151. Kurepa, J.; Walker, J.M.; Smalle, J.; Gosink, M.M.; Davis, S.J.; Durham, T.L.; Sung, D.-Y.; Vierstra, R.D. The small ubiquitin-like modifier (SUMO) protein modification system in Arabidopsis: Accumulation of SUMO1 and-2 conjugates is increased by stress. J. Biol. Chem. 2003, 278, 6862–6872. [Google Scholar] [CrossRef] [PubMed]
  152. Li, B.; Gao, K.; Ren, H.; Tang, W. Molecular mechanisms governing plant responses to high temperatures. J. Integr. Plant Biol. 2018, 60, 757–779. [Google Scholar] [CrossRef] [PubMed]
  153. Bouman, B.; Peng, S.; Castaneda, A.; Visperas, R. Yield and water use of irrigated tropical aerobic rice systems. Agric. Water Manag. 2005, 74, 87–105. [Google Scholar] [CrossRef]
  154. Hasanuzzaman, M.; Fujita, M.; Nahar, K.; Biswas, J.K. Advances in Rice Research for Abiotic Stress Tolerance; Woodhead Publishing: Duxford, UK, 2018. [Google Scholar]
  155. Kim, Y.; Chung, Y.S.; Lee, E.; Tripathi, P.; Heo, S.; Kim, K.-H. Root response to drought stress in rice (Oryza sativa L.). Int. J. Mol. Sci. 2020, 21, 1513. [Google Scholar] [CrossRef]
  156. Dash, P.K.; Rai, R.; Rai, V.; Pasupalak, S. Drought induced signaling in rice: Delineating canonical and non-canonical pathways. Front. Chem. 2018, 6, 264. [Google Scholar] [CrossRef]
  157. Du, H.; Huang, F.; Wu, N.; Li, X.; Hu, H.; Xiong, L. Integrative regulation of drought escape through ABA-dependent and-independent pathways in rice. Mol. Plant 2018, 11, 584–597. [Google Scholar] [CrossRef] [PubMed]
  158. Blum, A. Effective use of water (EUW) and not water-use efficiency (WUE) is the target of crop yield improvement under drought stress. Field Crops Res. 2009, 112, 119–123. [Google Scholar] [CrossRef]
  159. Uga, Y.; Yamamoto, E.; Kanno, N.; Kawai, S.; Mizubayashi, T.; Fukuoka, S. A major QTL controlling deep rooting on rice chromosome 4. Sci. Rep. 2013, 3, 3040. [Google Scholar] [CrossRef] [PubMed]
  160. Obara, M.; Tamura, W.; Ebitani, T.; Yano, M.; Sato, T.; Yamaya, T. Fine-mapping of qRL6. 1, a major QTL for root length of rice seedlings grown under a wide range of NH4+ concentrations in hydroponic conditions. Theor. Appl. Genet. 2010, 121, 535–547. [Google Scholar] [CrossRef]
  161. Wang, H.; Xu, X.; Zhan, X.; Zhai, R.; Wu, W.; Shen, X.; Dai, G.; Cao, L.; Cheng, S. Identification of qRL7, a major quantitative trait locus associated with rice root length in hydroponic conditions. Breed. Sci. 2013, 63, 267–274. [Google Scholar] [CrossRef]
  162. Li, X.; Guo, Z.; Lv, Y.; Cen, X.; Ding, X.; Wu, H.; Li, X.; Huang, J.; Xiong, L. Genetic control of the root system in rice under normal and drought stress conditions by genome-wide association study. PLoS Genet. 2017, 13, e1006889. [Google Scholar] [CrossRef]
  163. Kim, H.; Lee, K.; Hwang, H.; Bhatnagar, N.; Kim, D.-Y.; Yoon, I.S.; Byun, M.-O.; Kim, S.T.; Jung, K.-H.; Kim, B.-G. Overexpression of PYL5 in rice enhances drought tolerance, inhibits growth, and modulates gene expression. J. Exp. Bot. 2014, 65, 453–464. [Google Scholar] [CrossRef]
  164. Rahman, H.; Ramanathan, V.; Nallathambi, J.; Duraialagaraja, S.; Muthurajan, R. Over-expression of a NAC 67 transcription factor from finger millet (Eleusine coracana L.) confers tolerance against salinity and drought stress in rice. BMC Biotechnol. 2016, 16, 7–20. [Google Scholar] [CrossRef]
  165. Qi, J.; Song, C.P.; Wang, B.; Zhou, J.; Kangasjärvi, J.; Zhu, J.K.; Gong, Z. Reactive oxygen species signaling and stomatal movement in plant responses to drought stress and pathogen attack. J. Integr. Plant Biol. 2018, 60, 805–826. [Google Scholar] [CrossRef]
  166. Chen, X.; Ding, Y.; Yang, Y.; Song, C.; Wang, B.; Yang, S.; Guo, Y.; Gong, Z. Protein kinases in plant responses to drought, salt, and cold stress. J. Integr. Plant Biol. 2021, 63, 53–78. [Google Scholar] [CrossRef]
  167. You, J.; Zong, W.; Li, X.; Ning, J.; Hu, H.; Li, X.; Xiao, J.; Xiong, L. The SNAC1-targeted gene OsSRO1c modulates stomatal closure and oxidative stress tolerance by regulating hydrogen peroxide in rice. J. Exp. Bot. 2013, 64, 569–583. [Google Scholar] [CrossRef] [PubMed]
  168. Huang, L.; Wang, Y.; Wang, W.; Zhao, X.; Qin, Q.; Sun, F.; Hu, F.; Zhao, Y.; Li, Z.; Fu, B. Characterization of transcription factor gene OsDRAP1 conferring drought tolerance in rice. Front. Plant Sci. 2018, 9, 94. [Google Scholar] [CrossRef]
  169. Hu, H.; Dai, M.; Yao, J.; Xiao, B.; Li, X.; Zhang, Q.; Xiong, L. Overexpressing a NAM, ATAF, and CUC (NAC) transcription factor enhances drought resistance and salt tolerance in rice. Proc. Natl. Acad. Sci. USA 2006, 103, 12987–12992. [Google Scholar] [CrossRef] [PubMed]
  170. Xiao, B.; Huang, Y.; Tang, N.; Xiong, L. Over-expression of a LEA gene in rice improves drought resistance under the field conditions. Theor. Appl. Genet. 2007, 115, 35–46. [Google Scholar] [CrossRef] [PubMed]
  171. Liu, C.; Mao, B.; Ou, S.; Wang, W.; Liu, L.; Wu, Y.; Chu, C.; Wang, X. OsbZIP71, a bZIP transcription factor, confers salinity and drought tolerance in rice. Plant Mol. Biol. 2014, 84, 19–36. [Google Scholar] [CrossRef]
  172. Raineri, J.; Wang, S.; Peleg, Z.; Blumwald, E.; Chan, R.L. The rice transcription factor OsWRKY47 is a positive regulator of the response to water deficit stress. Plant Mol. Biol. 2015, 88, 401–413. [Google Scholar] [CrossRef] [PubMed]
  173. Tang, N.; Zhang, H.; Li, X.; Xiao, J.; Xiong, L. Constitutive activation of transcription factor OsbZIP46 improves drought tolerance in rice. Plant Physiol. 2012, 158, 1755–1768. [Google Scholar] [CrossRef]
  174. Jeong, J.S.; Kim, Y.S.; Baek, K.H.; Jung, H.; Ha, S.-H.; Do Choi, Y.; Kim, M.; Reuzeau, C.; Kim, J.-K. Root-specific expression of OsNAC10 improves drought tolerance and grain yield in rice under field drought conditions. Plant Physiol. 2010, 153, 185–197. [Google Scholar] [CrossRef]
  175. Yu, L.; Chen, X.; Wang, Z.; Wang, S.; Wang, Y.; Zhu, Q.; Li, S.; Xiang, C. Arabidopsis enhanced drought tolerance1/HOMEODOMAIN GLABROUS11 confers drought tolerance in transgenic rice without yield penalty. Plant Physiol. 2013, 162, 1378–1391. [Google Scholar] [CrossRef]
  176. Ravikumar, G.; Manimaran, P.; Voleti, S.; Subrahmanyam, D.; Sundaram, R.; Bansal, K.; Viraktamath, B.; Balachandran, S. Stress-inducible expression of AtDREB1A transcription factor greatly improves drought stress tolerance in transgenic indica rice. Transgenic Res. 2014, 23, 421–439. [Google Scholar] [CrossRef]
  177. Duan, J.; Zhang, M.; Zhang, H.; Xiong, H.; Liu, P.; Ali, J.; Li, J.; Li, Z. OsMIOX, a myo-inositol oxygenase gene, improves drought tolerance through scavenging of reactive oxygen species in rice (Oryza sativa L.). Plant Sci. 2012, 196, 143–151. [Google Scholar] [CrossRef]
  178. Li, H.-W.; Zang, B.-S.; Deng, X.-W.; Wang, X.-P. Overexpression of the trehalose-6-phosphate synthase gene OsTPS1 enhances abiotic stress tolerance in rice. Planta 2011, 234, 1007–1018. [Google Scholar] [CrossRef]
  179. Cui, M.; Zhang, W.; Zhang, Q.; Xu, Z.; Zhu, Z.; Duan, F.; Wu, R. Induced over-expression of the transcription factor OsDREB2A improves drought tolerance in rice. Plant Physiol. Biochem. 2011, 49, 1384–1391. [Google Scholar] [CrossRef]
  180. Saijo, Y.; Hata, S.; Kyozuka, J.; Shimamoto, K.; Izui, K. Over-expression of a single Ca2+-dependent protein kinase confers both cold and salt/drought tolerance on rice plants. Plant J. 2000, 23, 319–327. [Google Scholar] [CrossRef]
  181. Xiang, Y.; Huang, Y.; Xiong, L. Characterization of stress-responsive CIPK genes in rice for stress tolerance improvement. Plant Physiol. 2007, 144, 1416–1428. [Google Scholar] [CrossRef]
  182. Sahebi, M.; Hanafi, M.M.; Rafii, M.; Mahmud, T.; Azizi, P.; Osman, M.; Abiri, R.; Taheri, S.; Kalhori, N.; Shabanimofrad, M. Improvement of drought tolerance in rice (Oryza sativa L.): Genetics, genomic tools, and the WRKY gene family. BioMed Res. Int. 2018, 2018, 3158474. [Google Scholar] [CrossRef]
  183. Wu, Z.; Chen, L.; Yu, Q.; Zhou, W.; Gou, X.; Li, J.; Hou, S. Multiple transcriptional factors control stomata development in rice. New Phytol. 2019, 223, 220–232. [Google Scholar] [CrossRef]
  184. Takahashi, R.; Nishio, T.; Ichizen, N.; Takano, T. Cloning and functional analysis of the K+ transporter, PhaHAK2, from salt-sensitive and salt-tolerant reed plants. Biotechnol. Lett. 2007, 29, 501–506. [Google Scholar] [CrossRef]
  185. Nieves-Cordones, M.; Miller, A.J.; Alemán, F.; Martínez, V.; Rubio, F. A putative role for the plasma membrane potential in the control of the expression of the gene encoding the tomato high-affinity potassium transporter HAK5. Plant Mol. Biol. 2008, 68, 521–532. [Google Scholar] [CrossRef]
  186. Cai, J.; Chen, L.; Qu, H.; Lian, J.; Liu, W.; Hu, Y.; Xu, G. Alteration of nutrient allocation and transporter genes expression in rice under N, P, K, and Mg deficiencies. Acta Physiol. Plant. 2012, 34, 939–946. [Google Scholar] [CrossRef]
  187. Gumi, A.M.; Guha, P.K.; Mazumder, A.; Jayaswal, P.; Mondal, T.K. Characterization of OglDREB2A gene from African rice (Oryza glaberrima), comparative analysis and its transcriptional regulation under salinity stress. 3 Biotech 2018, 8, 91. [Google Scholar] [CrossRef] [PubMed]
  188. Jan, A.; Maruyama, K.; Todaka, D.; Kidokoro, S.; Abo, M.; Yoshimura, E.; Shinozaki, K.; Nakashima, K.; Yamaguchi-Shinozaki, K. OsTZF1, a CCCH-tandem zinc finger protein, confers delayed senescence and stress tolerance in rice by regulating stress-related genes. Plant Physiol. 2013, 161, 1202–1216. [Google Scholar] [CrossRef]
  189. Qiu, D.; Xiao, J.; Xie, W.; Liu, H.; Li, X.; Xiong, L.; Wang, S. Rice gene network inferred from expression profiling of plants overexpressing OsWRKY13, a positive regulator of disease resistance. Mol. Plant 2008, 1, 538–551. [Google Scholar] [PubMed]
  190. Qiu, D.; Xiao, J.; Ding, X.; Xiong, M.; Cai, M.; Cao, Y.; Li, X.; Xu, C.; Wang, S. OsWRKY13 mediates rice disease resistance by regulating defense-related genes in salicylate-and jasmonate-dependent signaling. Mol. Plant Microbe Interact. 2007, 20, 492–499. [Google Scholar] [CrossRef] [PubMed]
  191. Zou, M.; Guan, Y.; Ren, H.; Zhang, F.; Chen, F. A bZIP transcription factor, OsABI5, is involved in rice fertility and stress tolerance. Plant Mol. Biol. 2008, 66, 675–683. [Google Scholar] [CrossRef]
  192. Mukherjee, K.; Choudhury, A.R.; Gupta, B.; Gupta, S.; Sengupta, D.N. An ABRE-binding factor, OSBZ8, is highly expressed in salt tolerant cultivars than in salt sensitive cultivars of indica rice. BMC Plant Biol. 2006, 6, 18. [Google Scholar] [CrossRef]
  193. Sun, S.-J.; Guo, S.-Q.; Yang, X.; Bao, Y.-M.; Tang, H.-J.; Sun, H.; Huang, J.; Zhang, H.-S. Functional analysis of a novel Cys2/His2-type zinc finger protein involved in salt tolerance in rice. J. Exp. Bot. 2010, 61, 2807–2818. [Google Scholar] [CrossRef]
  194. Zhang, S.; Haider, I.; Kohlen, W.; Jiang, L.; Bouwmeester, H.; Meijer, A.H.; Schluepmann, H.; Liu, C.-M.; Ouwerkerk, P.B. Function of the HD-Zip I gene Oshox22 in ABA-mediated drought and salt tolerances in rice. Plant Mol. Biol. 2012, 80, 571–585. [Google Scholar] [CrossRef]
  195. Ma, K.; Xiao, J.; Li, X.; Zhang, Q.; Lian, X. Sequence and expression analysis of the C3HC4-type RING finger gene family in rice. Gene 2009, 444, 33–45. [Google Scholar] [CrossRef]
  196. Mishra, P.; Singh, N.; Jain, A.; Jain, N.; Mishra, V. Identification of cis-regulatory elements associated with salinity and drought stress tolerance in rice from co-expressed gene interaction networks. Bioinformation 2018, 14, 123. [Google Scholar] [CrossRef]
  197. Asano, T.; Hayashi, N.; Kobayashi, M.; Aoki, N.; Miyao, A.; Mitsuhara, I.; Ichikawa, H.; Komatsu, S.; Hirochika, H.; Kikuchi, S. A rice calcium-dependent protein kinase OsCPK12 oppositely modulates salt-stress tolerance and blast disease resistance. Plant J. 2012, 69, 26–36. [Google Scholar] [CrossRef] [PubMed]
  198. Lee, S.-K.; Kim, B.-G.; Kwon, T.-R.; Jeong, M.-J.; Park, S.-R.; Lee, J.-W.; Byun, M.-O.; Kwon, H.-B.; Matthews, B.F.; Hong, C.-B. Overexpression of the mitogen-activated protein kinase gene OsMAPK33 enhances sensitivity to salt stress in rice (Oryza sativa L.). J. Biosci. 2011, 36, 139–151. [Google Scholar] [CrossRef] [PubMed]
  199. Jinjun, Z.; Peina, J.; Fang, Z.; Chongke, Z.; Bo, B.; Yaping, L.; Haifeng, W.; Fan, C.; Xianzhi, X. Ossrk1, an atypical s-receptor-like kinase positively regulates leaf width and salt tolerance in rice. Rice Sci. 2020, 27, 133–142. [Google Scholar] [CrossRef]
  200. Cheng, Y.; Qi, Y.; Zhu, Q.; Chen, X.; Wang, N.; Zhao, X.; Chen, H.; Cui, X.; Xu, L.; Zhang, W. New changes in the plasma-membrane-associated proteome of rice roots under salt stress. Proteomics 2009, 9, 3100–3114. [Google Scholar] [CrossRef] [PubMed]
  201. Song, X.-J.; Huang, W.; Shi, M.; Zhu, M.-Z.; Lin, H.-X. A QTL for rice grain width and weight encodes a previously unknown RING-type E3 ubiquitin ligase. Nat. Genet. 2007, 39, 623–630. [Google Scholar] [CrossRef] [PubMed]
  202. Ren, Z.-H.; Gao, J.-P.; Li, L.-G.; Cai, X.-L.; Huang, W.; Chao, D.-Y.; Zhu, M.-Z.; Wang, Z.-Y.; Luan, S.; Lin, H.-X. A rice quantitative trait locus for salt tolerance encodes a sodium transporter. Nat. Genet. 2005, 37, 1141–1146. [Google Scholar] [CrossRef] [PubMed]
  203. Chen, Y.; Zhou, X.; Chang, S.; Chu, Z.; Wang, H.; Han, S.; Wang, Y. Calcium-dependent protein kinase 21 phosphorylates 14-3-3 proteins in response to ABA signaling and salt stress in rice. Biochem. Biophys. Res. Commun. 2017, 493, 1450–1456. [Google Scholar] [CrossRef]
  204. Ge, L.-F.; Chao, D.-Y.; Shi, M.; Zhu, M.-Z.; Gao, J.-P.; Lin, H.-X. Overexpression of the trehalose-6-phosphate phosphatase gene OsTPP1 confers stress tolerance in rice and results in the activation of stress responsive genes. Planta 2008, 228, 191–201. [Google Scholar] [CrossRef]
  205. Luo, D.; Niu, X.; Yu, J.; Yan, J.; Gou, X.; Lu, B.-R.; Liu, Y. Rice choline monooxygenase (OsCMO) protein functions in enhancing glycine betaine biosynthesis in transgenic tobacco but does not accumulate in rice (Oryza sativa L. ssp. japonica). Plant Cell Rep. 2012, 31, 1625–1635. [Google Scholar] [CrossRef]
  206. Nakamura, A.; Fukuda, A.; Sakai, S.; Tanaka, Y. Molecular cloning, functional expression and subcellular localization of two putative vacuolar voltage-gated chloride channels in rice (Oryza sativa L.). Plant Cell Physiol. 2006, 47, 32–42. [Google Scholar] [CrossRef]
  207. Senadheera, P.; Singh, R.; Maathuis, F.J. Differentially expressed membrane transporters in rice roots may contribute to cultivar dependent salt tolerance. J. Exp. Bot. 2009, 60, 2553–2563. [Google Scholar] [CrossRef]
  208. Choe, Y.-H.; Kim, Y.-S.; Kim, I.-S.; Bae, M.-J.; Lee, E.-J.; Kim, Y.-H.; Park, H.-M.; Yoon, H.-S. Homologous expression of γ-glutamylcysteine synthetase increases grain yield and tolerance of transgenic rice plants to environmental stresses. J. Plant Physiol. 2013, 170, 610–618. [Google Scholar] [CrossRef]
  209. Guo, X.; Wu, Y.; Wang, Y.; Chen, Y.; Chu, C. OsMSRA4. 1 and OsMSRB1. 1, two rice plastidial methionine sulfoxide reductases, are involved in abiotic stress responses. Planta 2009, 230, 227–238. [Google Scholar] [CrossRef]
  210. Ushimaru, T.; Nakagawa, T.; Fujioka, Y.; Daicho, K.; Naito, M.; Yamauchi, Y.; Nonaka, H.; Amako, K.; Yamawaki, K.; Murata, N. Transgenic Arabidopsis plants expressing the rice dehydroascorbate reductase gene are resistant to salt stress. J. Plant Physiol. 2006, 163, 1179–1184. [Google Scholar] [CrossRef]
  211. Lu, Z.; Liu, D.; Liu, S. Two rice cytosolic ascorbate peroxidases differentially improve salt tolerance in transgenic Arabidopsis. Plant Cell Rep. 2007, 26, 1909–1917. [Google Scholar] [CrossRef]
  212. Zou, J.; Liu, C.; Liu, A.; Zou, D.; Chen, X. Overexpression of OsHsp17. 0 and OsHsp23. 7 enhances drought and salt tolerance in rice. J. Plant Physiol. 2012, 169, 628–635. [Google Scholar] [CrossRef]
  213. Ganguly, M.; Datta, K.; Roychoudhury, A.; Gayen, D.; Sengupta, D.N.; Datta, S.K. Overexpression of Rab16A gene in indica rice variety for generating enhanced salt tolerance. Plant Signal. Behav. 2012, 7, 502–509. [Google Scholar] [CrossRef]
  214. Battaglia, M.; Covarrubias, A.A. Late Embryogenesis Abundant (LEA) proteins in legumes. Front. Plant Sci. 2013, 4, 190. [Google Scholar] [CrossRef]
  215. Yoshiba, Y.; Kiyosue, T.; Nakashima, K.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Regulation of levels of proline as an osmolyte in plants under water stress. Plant Cell Physiol. 1997, 38, 1095–1102. [Google Scholar] [CrossRef]
  216. Ghosh, U.; Islam, M.; Siddiqui, M.; Cao, X.; Khan, M. Proline, a multifaceted signalling molecule in plant responses to abiotic stress: Understanding the physiological mechanisms. Plant Biol. 2022, 24, 227–239. [Google Scholar] [CrossRef]
  217. Ahamed, A.; Murai-Hatano, M.; Ishikawa-Sakurai, J.; Hayashi, H.; Kawamura, Y.; Uemura, M. Cold stress-induced acclimation in rice is mediated by root-specific aquaporins. Plant Cell Physiol. 2012, 53, 1445–1456. [Google Scholar] [CrossRef] [PubMed]
  218. Castillo, E.G.; Tuong, T.P.; Ismail, A.M.; Inubushi, K. Response to salinity in rice: Comparative effects of osmotic and ionic stresses. Plant Prod. Sci. 2007, 10, 159–170. [Google Scholar] [CrossRef]
  219. Hong, Z.; Lakkineni, K.; Zhang, Z.; Verma, D.P.S. Removal of feedback inhibition of Δ1-pyrroline-5-carboxylate synthetase results in increased proline accumulation and protection of plants from osmotic stress. Plant Physiol. 2000, 122, 1129–1136. [Google Scholar] [CrossRef] [PubMed]
  220. Liang, W.; Cui, W.; Ma, X.; Wang, G.; Huang, Z. Function of wheat Ta-UnP gene in enhancing salt tolerance in transgenic Arabidopsis and rice. Biochem. Biophys. Res. Commun. 2014, 450, 794–801. [Google Scholar] [CrossRef]
  221. Li, Q.; Li, B.; Wang, J.; Chang, X.; Mao, X.; Jing, R. TaPUB15, a U-Box E3 ubiquitin ligase gene from wheat, enhances salt tolerance in rice. Food Energy Secur. 2021, 10, e250. [Google Scholar] [CrossRef]
  222. Yang, Y.; Ma, X.; Xia, H.; Wang, L.; Chen, S.; Xu, K.; Yang, F.; Zou, Y.; Wang, Y.; Zhu, J. Natural variation of Alfin-like family affects seed size and drought tolerance in rice. Plant J. 2022, 112, 1176–1193. [Google Scholar] [CrossRef] [PubMed]
  223. Tenhaken, R. Cell wall remodeling under abiotic stress. Front. Plant Sci. 2015, 5, 771. [Google Scholar] [CrossRef]
  224. Wang, Y.; Du, F.; Wang, J.; Li, Y.; Zhang, Y.; Zhao, X.; Zheng, T.; Li, Z.; Xu, J.; Wang, W. Molecular dissection of the gene OsGA2ox8 conferring osmotic stress tolerance in rice. Int. J. Mol. Sci. 2021, 22, 9107. [Google Scholar] [CrossRef]
  225. Jin, X.; Zhang, Y.; Li, X.; Huang, J. OsNF-YA3 regulates plant growth and osmotic stress tolerance by interacting with SLR1 and SAPK9 in rice. Plant J. 2023. [Google Scholar] [CrossRef]
  226. Singh, A.; Giri, J.; Kapoor, S.; Tyagi, A.K.; Pandey, G.K. Protein phosphatase complement in rice: Genome-wide identification and transcriptional analysis under abiotic stress conditions and reproductive development. BMC Genom. 2010, 11, 435. [Google Scholar] [CrossRef]
  227. Xiang, J.; Chen, X.; Hu, W.; Xiang, Y.; Yan, M.; Wang, J. Overexpressing heat-shock protein OsHSP50. 2 improves drought tolerance in rice. Plant Cell Rep. 2018, 37, 1585–1595. [Google Scholar] [CrossRef] [PubMed]
  228. Zhang, H.; Zhai, N.; Ma, X.; Zhou, H.; Cui, Y.; Wang, C.; Xu, G. Overexpression of OsRLCK241 confers enhanced salt and drought tolerance in transgenic rice (Oryza sativa L.). Gene 2021, 768, 145278. [Google Scholar] [CrossRef] [PubMed]
  229. Bailey-Serres, J.; Chang, R. Sensing and signalling in response to oxygen deprivation in plants and other organisms. Ann. Bot. 2005, 96, 507–518. [Google Scholar] [CrossRef] [PubMed]
  230. Voesenek, L.; Sasidharan, R. Ethylene–and oxygen signalling–drive plant survival during flooding. Plant Biol. 2013, 15, 426–435. [Google Scholar] [CrossRef] [PubMed]
  231. Drew, M.C.; He, C.-J.; Morgan, P.W. Programmed cell death and aerenchyma formation in roots. Trends Plant Sci. 2000, 5, 123–127. [Google Scholar] [CrossRef] [PubMed]
  232. Shiono, K.; Takahashi, H.; Colmer, T.D.; Nakazono, M. Role of ethylene in acclimations to promote oxygen transport in roots of plants in waterlogged soils. Plant Sci. 2008, 175, 52–58. [Google Scholar] [CrossRef]
  233. 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]
  234. Hattori, Y.; Nagai, K.; Furukawa, S.; Song, X.-J.; Kawano, R.; Sakakibara, H.; Wu, J.; Matsumoto, T.; Yoshimura, A.; Kitano, H. The ethylene response factors SNORKEL1 and SNORKEL2 allow rice to adapt to deep water. Nature 2009, 460, 1026–1030. [Google Scholar] [CrossRef]
  235. Colmer, T.; Pedersen, O. Oxygen dynamics in submerged rice (Oryza sativa). New Phytol. 2008, 178, 326–334. [Google Scholar] [CrossRef]
  236. Kulichikhin, K.; Yamauchi, T.; Watanabe, K.; Nakazono, M. Biochemical and molecular characterization of rice (O ryza sativa L.) roots forming a barrier to radial oxygen loss. Plant Cell Environ. 2014, 37, 2406–2420. [Google Scholar]
  237. Sasaki, A.; Ashikari, M.; Ueguchi-Tanaka, M.; Itoh, H.; Nishimura, A.; Swapan, D.; Ishiyama, K.; Saito, T.; Kobayashi, M.; Khush, G.S. A mutant gibberellin-synthesis gene in rice. Nature 2002, 416, 701–702. [Google Scholar] [CrossRef] [PubMed]
  238. Xu, K.; 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]
  239. Ismail, A.M.; Ella, E.S.; Vergara, G.V.; Mackill, D.J. Mechanisms associated with tolerance to flooding during germination and early seedling growth in rice (Oryza sativa). Ann. Bot. 2009, 103, 197–209. [Google Scholar] [CrossRef] [PubMed]
  240. Magneschi, L.; Perata, P. Rice germination and seedling growth in the absence of oxygen. Ann. Bot. 2009, 103, 181–196. [Google Scholar] [CrossRef]
  241. Kretzschmar, T.; Pelayo, M.A.F.; Trijatmiko, K.R.; Gabunada, L.F.M.; Alam, R.; Jimenez, R.; Mendioro, M.S.; Slamet-Loedin, I.H.; Sreenivasulu, N.; Bailey-Serres, J. A trehalose-6-phosphate phosphatase enhances anaerobic germination tolerance in rice. Nat. Plants 2015, 1, 15124. [Google Scholar] [CrossRef]
  242. Xu, K.; Mackill, D.J. A major locus for submergence tolerance mapped on rice chromosome 9. Mol. Breed. 1996, 2, 219–224. [Google Scholar] [CrossRef]
  243. Septiningsih, E.M.; Pamplona, A.M.; Sanchez, D.L.; Neeraja, C.N.; Vergara, G.V.; Heuer, S.; Ismail, A.M.; Mackill, D.J. Development of submergence-tolerant rice cultivars: The Sub1 locus and beyond. Ann. Bot. 2009, 103, 151–160. [Google Scholar] [CrossRef]
  244. 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]
  245. Fukao, T.; Yeung, E.; Bailey-Serres, J. The submergence tolerance gene SUB1A delays leaf senescence under prolonged darkness through hormonal regulation in rice. Plant Physiol. 2012, 160, 1795–1807. [Google Scholar] [CrossRef]
  246. 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]
  247. Drew, M.C. Oxygen deficiency and root metabolism: Injury and acclimation under hypoxia and anoxia. Annu. Rev. Plant Biol. 1997, 48, 223–250. [Google Scholar] [CrossRef] [PubMed]
  248. Fukao, T.; Yeung, E.; Bailey-Serres, J. The submergence tolerance regulator SUB1A mediates crosstalk between submergence and drought tolerance in rice. Plant Cell 2011, 23, 412–427. [Google Scholar] [CrossRef] [PubMed]
  249. Singh, P.; Sinha, A.K. A positive feedback loop governed by SUB1A1 interaction with MITOGEN-ACTIVATED PROTEIN KINASE3 imparts submergence tolerance in rice. Plant Cell 2016, 28, 1127–1143. [Google Scholar] [CrossRef] [PubMed]
  250. Reynoso, M.A.; Kajala, K.; Bajic, M.; West, D.A.; Pauluzzi, G.; Yao, A.I.; Hatch, K.; Zumstein, K.; Woodhouse, M.; Rodriguez-Medina, J. Evolutionary flexibility in flooding response circuitry in angiosperms. Science 2019, 365, 1291–1295. [Google Scholar] [CrossRef] [PubMed]
  251. Nagai, K.; Kurokawa, Y.; Mori, Y.; Minami, A.; Reuscher, S.; Wu, J.; Matsumoto, T.; Ashikari, M. SNORKEL genes relating to flood tolerance were pseudogenized in normal cultivated rice. Plants 2022, 11, 376. [Google Scholar] [CrossRef]
  252. Nemoto, K.; Ukai, Y.; Tang, D.-Q.; Kasai, Y.; Morita, M. Inheritance of early elongation ability in floating rice revealed by diallel and QTL analyses. Theor. Appl. Genet. 2004, 109, 42–47. [Google Scholar] [CrossRef]
  253. Tang, D.-Q.; Kasai, Y.; Miyamoto, N.; Ukai, Y.; Nemoto, K. Comparison of QTLs for early elongation ability between two floating rice cultivars with a different phylogenetic origin. Breed. Sci. 2005, 55, 1–5. [Google Scholar] [CrossRef]
  254. Hattori, Y.; Miura, K.; Asano, K.; Yamamoto, E.; Mori, H.; Kitano, H.; Matsuoka, M.; Ashikari, M. A major QTL confers rapid internode elongation in response to water rise in deepwater rice. Breed. Sci. 2007, 57, 305–314. [Google Scholar] [CrossRef]
  255. Kawano, R.; Doi, K.; Yasui, H.; Mochizuki, T.; Yoshimura, A. Mapping of QTLs for floating ability in rice. Breed. Sci. 2008, 58, 47–53. [Google Scholar] [CrossRef]
  256. Kuroha, T.; Ashikari, M. Molecular mechanisms and future improvement of submergence tolerance in rice. Mol. Breed. 2020, 40, 41. [Google Scholar] [CrossRef]
  257. Kuroha, T.; Nagai, K.; Gamuyao, R.; Wang, D.R.; Furuta, T.; Nakamori, M.; Kitaoka, T.; Adachi, K.; Minami, A.; Mori, Y. Ethylene-gibberellin signaling underlies adaptation of rice to periodic flooding. Science 2018, 361, 181–186. [Google Scholar] [CrossRef] [PubMed]
  258. Yamauchi, T.; Yoshioka, M.; Fukazawa, A.; Mori, H.; Nishizawa, N.K.; Tsutsumi, N.; Yoshioka, H.; Nakazono, M. An NADPH oxidase RBOH functions in rice roots during lysigenous aerenchyma formation under oxygen-deficient conditions. Plant Cell 2017, 29, 775–790. [Google Scholar] [CrossRef] [PubMed]
  259. Zhang, T.; Zhao, X.; Wang, W.; Pan, Y.; Huang, L.; Liu, X.; Zong, Y.; Zhu, L.; Yang, D.; Fu, B. Comparative transcriptome profiling of chilling stress responsiveness in two contrasting rice genotypes. PLoS ONE 2012, 7, e43274. [Google Scholar] [CrossRef] [PubMed]
  260. Wei, Z.; Yuan, Q.; Lin, H.; Li, X.; Zhang, C.; Gao, H.; Zhang, B.; He, H.; Liu, T.; Jie, Z. Linkage analysis, GWAS, transcriptome analysis to identify candidate genes for rice seedlings in response to high temperature stress. BMC Plant Biol. 2021, 21, 85. [Google Scholar] [CrossRef]
  261. Yang, Y.; Zhang, C.; Zhu, D.; He, H.; Wei, Z.; Yuan, Q.; Li, X.; Gao, X.; Zhang, B.; Gao, H. Identifying candidate genes and patterns of heat-stress response in rice using a genome-wide association study and transcriptome analyses. Crop J. 2022, 10, 1633–1643. [Google Scholar] [CrossRef]
  262. Zhang, X.; Rerksiri, W.; Liu, A.; Zhou, X.; Xiong, H.; Xiang, J.; Chen, X.; Xiong, X. Transcriptome profile reveals heat response mechanism at molecular and metabolic levels in rice flag leaf. Gene 2013, 530, 185–192. [Google Scholar] [CrossRef]
  263. González-Schain, N.; Dreni, L.; Lawas, L.M.; Galbiati, M.; Colombo, L.; Heuer, S.; Jagadish, K.S.; Kater, M.M. Genome-wide transcriptome analysis during anthesis reveals new insights into the molecular basis of heat stress responses in tolerant and sensitive rice varieties. Plant Cell Physiol. 2016, 57, 57–68. [Google Scholar] [CrossRef]
  264. Jung, K.-H.; Ko, H.-J.; Nguyen, M.X.; Kim, S.-R.; Ronald, P.; An, G. Genome-wide identification and analysis of early heat stress responsive genes in rice. J. Plant Biol. 2012, 55, 458–468. [Google Scholar] [CrossRef]
  265. Yoo, Y.-H.; Nalini Chandran, A.K.; Park, J.-C.; Gho, Y.-S.; Lee, S.-W.; An, G.; Jung, K.-H. OsPhyB-mediating novel regulatory pathway for drought tolerance in rice root identified by a global RNA-Seq transcriptome analysis of rice genes in response to water deficiencies. Front. Plant Sci. 2017, 8, 580. [Google Scholar] [CrossRef]
  266. Kong, W.; Zhong, H.; Gong, Z.; Fang, X.; Sun, T.; Deng, X.; Li, Y. Meta-analysis of salt stress transcriptome responses in different rice genotypes at the seedling stage. Plants 2019, 8, 64. [Google Scholar] [CrossRef]
  267. Chandran, A.K.N.; Kim, J.-W.; Yoo, Y.-H.; Park, H.L.; Kim, Y.-J.; Cho, M.-H.; Jung, K.-H. Transcriptome analysis of rice-seedling roots under soil–salt stress using RNA-Seq method. Plant Biotechnol. Rep. 2019, 13, 567–578. [Google Scholar] [CrossRef]
  268. Bansal, J.; Gupta, K.; Rajkumar, M.S.; Garg, R.; Jain, M. Draft genome and transcriptome analyses of halophyte rice Oryza coarctata provide resources for salinity and submergence stress response factors. Physiol. Plant. 2021, 173, 1309–1322. [Google Scholar] [CrossRef] [PubMed]
  269. Kong, W.; Li, S.; Zhang, C.; Qiang, Y.; Li, Y. Combination of quantitative trait locus (QTL) mapping and transcriptome analysis reveals submerged germination QTLs and candidate genes controlling coleoptile length in rice. Food Energy Secur. 2022, 11, e354. [Google Scholar] [CrossRef]
  270. Venu, R.; Sreerekha, M.; Sheshu Madhav, M.; Nobuta, K.; Mohan, K.M.; Chen, S.; Jia, Y.; Meyers, B.C.; Wang, G.-L. Deep transcriptome sequencing reveals the expression of key functional and regulatory genes involved in the abiotic stress signaling pathways in rice. J. Plant Biol. 2013, 56, 216–231. [Google Scholar] [CrossRef]
  271. Yan, S.-P.; Zhang, Q.-Y.; Tang, Z.-C.; Su, W.-A.; Sun, W.-N. Comparative proteomic analysis provides new insights into chilling stress responses in rice. Mol. Cell. Proteom. 2006, 5, 484–496. [Google Scholar] [CrossRef]
  272. Lee, J.; Lee, Y.; Kim, M.; Ham, T.-H.; Jo, S.-M.; Kwon, S.-W. Quantitative shotgun proteomic analysis of cold-stressed mature rice anthers. Plant Biotechnol. Rep. 2017, 11, 417–427. [Google Scholar] [CrossRef]
  273. Neilson, K.A.; Mariani, M.; Haynes, P.A. Quantitative proteomic analysis of cold-responsive proteins in rice. Proteomics 2011, 11, 1696–1706. [Google Scholar] [CrossRef] [PubMed]
  274. Jagadish, S.K.; Muthurajan, R.; Rang, Z.W.; Malo, R.; Heuer, S.; Bennett, J.; Craufurd, P.Q. Spikelet proteomic response to combined water deficit and heat stress in rice (Oryza sativa cv. N22). Rice 2011, 4, 1–11. [Google Scholar] [CrossRef]
  275. Zhang, H.Y.; Lei, G.; Zhou, H.W.; He, C.; Liao, J.L.; Huang, Y.J. Quantitative iTRAQ-based proteomic analysis of rice grains to assess high night temperature stress. Proteomics 2017, 17, 1600365. [Google Scholar] [CrossRef]
  276. Liao, J.-L.; Huang, Y.-J. Evaluation of protocols used in 2-D electrophoresis for proteome analysis of young rice caryopsis. Genom. Proteom. Bioinform. 2011, 9, 229–237. [Google Scholar] [CrossRef]
  277. Vashisht, A.A.; Tuteja, N. Stress responsive DEAD-box helicases: A new pathway to engineer plant stress tolerance. J. Photochem. Photobiol. B Biol. 2006, 84, 150–160. [Google Scholar] [CrossRef] [PubMed]
  278. Maksup, S.; Roytrakul, S.; Supaibulwatana, K. Physiological and comparative proteomic analyses of Thai jasmine rice and two check cultivars in response to drought stress. J. Plant Interact. 2014, 9, 43–55. [Google Scholar] [CrossRef]
  279. Liu, Y.; Pan, T.; Tang, Y.; Zhuang, Y.; Liu, Z.; Li, P.; Li, H.; Huang, W.; Tu, S.; Ren, G. Proteomic analysis of rice subjected to low light stress and overexpression of OsGAPB increases the stress tolerance. Rice 2020, 13, 30. [Google Scholar] [CrossRef] [PubMed]
  280. Liu, D.; Ford, K.L.; Roessner, U.; Natera, S.; Cassin, A.M.; Patterson, J.H.; Bacic, A. Rice suspension cultured cells are evaluated as a model system to study salt responsive networks in plants using a combined proteomic and metabolomic profiling approach. Proteomics 2013, 13, 2046–2062. [Google Scholar] [CrossRef] [PubMed]
  281. Xu, J.; Lan, H.; Fang, H.; Huang, X.; Zhang, H.; Huang, J. Quantitative proteomic analysis of the rice (Oryza sativa L.) salt response. PLoS ONE 2015, 10, e0120978. [Google Scholar] [CrossRef]
  282. Kim, S.T.; Kim, S.G.; Kang, Y.H.; Wang, Y.; Kim, J.-Y.; Yi, N.; Kim, J.-K.; Rakwal, R.; Koh, H.-J.; Kang, K.Y. Proteomics analysis of rice lesion mimic mutant (spl 1) reveals tightly localized probenazole-induced protein (PBZ1) in cells undergoing programmed cell death. J. Proteome Res. 2008, 7, 1750–1760. [Google Scholar] [CrossRef]
  283. Zhang, J.; Luo, W.; Zhao, Y.; Xu, Y.; Song, S.; Chong, K. Comparative metabolomic analysis reveals a reactive oxygen species-dominated dynamic model underlying chilling environment adaptation and tolerance in rice. New Phytol. 2016, 211, 1295–1310. [Google Scholar] [CrossRef]
  284. Wada, H.; Hatakeyama, Y.; Nakashima, T.; Nonami, H.; Erra-Balsells, R.; Hakata, M.; Nakata, K.; Hiraoka, K.; Onda, Y.; Nakano, H. On-site single pollen metabolomics reveals varietal differences in phosphatidylinositol synthesis under heat stress conditions in rice. Sci. Rep. 2020, 10, 2013. [Google Scholar] [CrossRef]
  285. Ma, X.; Xia, H.; Liu, Y.; Wei, H.; Zheng, X.; Song, C.; Chen, L.; Liu, H.; Luo, L. Transcriptomic and metabolomic studies disclose key metabolism pathways contributing to well-maintained photosynthesis under the drought and the consequent drought-tolerance in rice. Front. Plant Sci. 2016, 7, 1886. [Google Scholar] [CrossRef]
  286. Degenkolbe, T.; Do, P.T.; Kopka, J.; Zuther, E.; Hincha, D.K.; Köhl, K.I. Identification of drought tolerance markers in a diverse population of rice cultivars by expression and metabolite profiling. PLoS ONE 2013, 8, e63637. [Google Scholar] [CrossRef]
  287. Wang, W.-S.; Zhao, X.-Q.; Li, M.; Huang, L.-Y.; Xu, J.-L.; Zhang, F.; Cui, Y.-R.; Fu, B.-Y.; Li, Z.-K. Complex molecular mechanisms underlying seedling salt tolerance in rice revealed by comparative transcriptome and metabolomic profiling. J. Exp. Bot. 2016, 67, 405–419. [Google Scholar] [CrossRef] [PubMed]
  288. Barding, G.A., Jr.; Fukao, T.; Béni, S.; Bailey-Serres, J.; Larive, C.K. Differential metabolic regulation governed by the rice SUB1A gene during submergence stress and identification of alanylglycine by 1H NMR spectroscopy. J. Proteome Res. 2012, 11, 320–330. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mechanism of cold, heat, and drought stress sensing and tolerance in rice. (A) Cold stress mechanism and response. Cold stress signals are perceived by COLD1 and CIPK, and different genes related to phytohormones and osmoprotectants are regulated. The upregulation of ABA-responsive genes leads to ABA accumulation and cold tolerance. (B) Heat stress sensing and response. Heat stress signals are perceived by different heat shock transcription factors and proteins. Different genes associated with ROS, lipid metabolism, Ca2+ homeostasis, and phytohormones are regulated. Several ROS and cell homeostasis genes are activated to trigger the heat stress response. (C) Pathway of drought sensing and tolerance. The root system is crucial for drought tolerance, and DRO1 is upregulated under drought stress, leading to deeper roots and improved drought tolerance. Other genes associated with phytohormones, stomatal balance, water-use efficiency, osmotic adjustment, and root and shoot biomass are crucial for drought tolerance.
Figure 1. Mechanism of cold, heat, and drought stress sensing and tolerance in rice. (A) Cold stress mechanism and response. Cold stress signals are perceived by COLD1 and CIPK, and different genes related to phytohormones and osmoprotectants are regulated. The upregulation of ABA-responsive genes leads to ABA accumulation and cold tolerance. (B) Heat stress sensing and response. Heat stress signals are perceived by different heat shock transcription factors and proteins. Different genes associated with ROS, lipid metabolism, Ca2+ homeostasis, and phytohormones are regulated. Several ROS and cell homeostasis genes are activated to trigger the heat stress response. (C) Pathway of drought sensing and tolerance. The root system is crucial for drought tolerance, and DRO1 is upregulated under drought stress, leading to deeper roots and improved drought tolerance. Other genes associated with phytohormones, stomatal balance, water-use efficiency, osmotic adjustment, and root and shoot biomass are crucial for drought tolerance.
Plants 12 02019 g001
Figure 2. Salt stress, osmotic stress, and submergence stress tolerance mechanisms in rice. (A) Salt stress sensing and response. Different proteins play a role in antioxidant and osmoprotectant accumulation, ROS and Na+ homeostasis, MDA accumulation, and electrolyte leakage and are required to trigger the tolerance mechanism. Some WRKY TFs suppress the expression of OsNAC1 and DREB1B, resulting in salt susceptibility. (B) Osmotic stress tolerance mechanism overview in rice. The genes involving ABA, proline, and polysaccharides biosynthesis are mentioned. Further, genes involved in ROS scavenging, preventing electrolyte leakage, and balancing the intracellular Ca2+ concentration are also highlighted. (C) Submergence tolerance response and regulatory mechanism. Rice follows the quiescent strategy to adapt and escape periodic flash flooding. SUB1A is the key regulator for submergence tolerance. It triggers the transcriptional regulation of SLR1 and other ERF response factors. In floating rice, ethylene accumulation under deep-water conditions stabilizes the ethylene signaling factor OsEIL1. OsEIL1 increases gene expression by binding to the promoter of SD1. After that, the accumulated gibberellin increases the expression of ACE1; meanwhile, the expression of DEC1, a factor inhibiting internode elongation, is reduced. On the other hand, OsEIL1 also binds to the promoter regions of SNORKEL1 and SNORKEL2, triggering the expression of other downstream genes.
Figure 2. Salt stress, osmotic stress, and submergence stress tolerance mechanisms in rice. (A) Salt stress sensing and response. Different proteins play a role in antioxidant and osmoprotectant accumulation, ROS and Na+ homeostasis, MDA accumulation, and electrolyte leakage and are required to trigger the tolerance mechanism. Some WRKY TFs suppress the expression of OsNAC1 and DREB1B, resulting in salt susceptibility. (B) Osmotic stress tolerance mechanism overview in rice. The genes involving ABA, proline, and polysaccharides biosynthesis are mentioned. Further, genes involved in ROS scavenging, preventing electrolyte leakage, and balancing the intracellular Ca2+ concentration are also highlighted. (C) Submergence tolerance response and regulatory mechanism. Rice follows the quiescent strategy to adapt and escape periodic flash flooding. SUB1A is the key regulator for submergence tolerance. It triggers the transcriptional regulation of SLR1 and other ERF response factors. In floating rice, ethylene accumulation under deep-water conditions stabilizes the ethylene signaling factor OsEIL1. OsEIL1 increases gene expression by binding to the promoter of SD1. After that, the accumulated gibberellin increases the expression of ACE1; meanwhile, the expression of DEC1, a factor inhibiting internode elongation, is reduced. On the other hand, OsEIL1 also binds to the promoter regions of SNORKEL1 and SNORKEL2, triggering the expression of other downstream genes.
Plants 12 02019 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Usman, B.; Derakhshani, B.; Jung, K.-H. Recent Molecular Aspects and Integrated Omics Strategies for Understanding the Abiotic Stress Tolerance of Rice. Plants 2023, 12, 2019. https://doi.org/10.3390/plants12102019

AMA Style

Usman B, Derakhshani B, Jung K-H. Recent Molecular Aspects and Integrated Omics Strategies for Understanding the Abiotic Stress Tolerance of Rice. Plants. 2023; 12(10):2019. https://doi.org/10.3390/plants12102019

Chicago/Turabian Style

Usman, Babar, Behnam Derakhshani, and Ki-Hong Jung. 2023. "Recent Molecular Aspects and Integrated Omics Strategies for Understanding the Abiotic Stress Tolerance of Rice" Plants 12, no. 10: 2019. https://doi.org/10.3390/plants12102019

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