Next Article in Journal
Blockade of TIPE2-Mediated Ferroptosis of Myeloid-Derived Suppressor Cells Achieves the Full Potential of Combinatory Ferroptosis and Anti-PD-L1 Cancer Immunotherapy
Previous Article in Journal
Valproate-Induced Model of Autism in Adult Zebrafish: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plant Coping with Cold Stress: Molecular and Physiological Adaptive Mechanisms with Future Perspectives

1
Henan Collaborative Innovation Centre of Modern Biological Breeding, Henan Institute of Science and Technology, Xinxiang 453003, China
2
State Key Laboratory of Herbage Improvement and Grassland Agroecosystems, College of Ecology, Lanzhou University, Lanzhou 730000, China
3
Department of Agronomy, College of Agriculture, Shandong Agriculture University, Tai’an 271018, China
4
Agricultural Research Station, Office of VP for Research & Graduate Studies, Qatar University, Doha 2713, Qatar
*
Author to whom correspondence should be addressed.
Cells 2025, 14(2), 110; https://doi.org/10.3390/cells14020110
Submission received: 17 December 2024 / Revised: 10 January 2025 / Accepted: 11 January 2025 / Published: 13 January 2025

Abstract

:
Cold stress strongly hinders plant growth and development. However, the molecular and physiological adaptive mechanisms of cold stress tolerance in plants are not well understood. Plants adopt several morpho-physiological changes to withstand cold stress. Plants have evolved various strategies to cope with cold stress. These strategies included changes in cellular membranes and chloroplast structure, regulating cold signals related to phytohormones and plant growth regulators (ABA, JA, GA, IAA, SA, BR, ET, CTK, and MET), reactive oxygen species (ROS), protein kinases, and inorganic ions. This review summarizes the mechanisms of how plants respond to cold stress, covering four main signal transduction pathways, including the abscisic acid (ABA) signal transduction pathway, Ca2+ signal transduction pathway, ROS signal transduction pathway, and mitogen-activated protein kinase (MAPK/MPK) cascade pathway. Some transcription factors, such as AP2/ERF, MYB, WRKY, NAC, and bZIP, not only act as calmodulin-binding proteins during cold perception but can also play important roles in the downstream chilling-signaling pathway. This review also highlights the analysis of those transcription factors such as bHLH, especially bHLH-type transcription factors ICE, and discusses their functions as phytohormone-responsive elements binding proteins in the promoter region under cold stress. In addition, a theoretical framework outlining plant responses to cold stress tolerance has been proposed. This theory aims to guide future research directions and inform agricultural production practices, ultimately enhancing crop resilience to cold stress.

1. Introduction

Crop yields are negatively affected by cold stress, including chilling (0~12 °C) and freezing stress (≤0 °C) resulting in constrained sowing time, damage to plant tissues, and stunted plant growth [1]. Tolerance to cold stress in plants can be divided into chill-susceptible (<12 °C plants will be damaged), chill-tolerant but freezing susceptible (0~12 °C plants can adapt, but once the tissues are frozen, the cells will be damaged), and freeze-tolerant (survival < 0 °C) [2]. Cold stress seriously threatens global crop productivity and yield reduction in temperate climates [3]. The influence of cold stress depends on the status of plant cellular morphology (e.g., membranes and organelles) [4], biochemistry (e.g., enzyme activity) [5], physiological functioning (e.g., gas exchange and water management) [6], and phenology (e.g., developmental stages) [7]. The physio-biochemical adaptive mechanisms of plants have been widely discussed, but less attention has been paid to molecular adaptive mechanisms. Thus, plant responses to cold stress are complex and multidimensional, involving signal perception and transduction.
Plants respond to cold stress through a series of physiological adaptations, including regulation of phytohormones (e.g., abscisic acid, brassinosteroids, and jasmonic acid) [8], osmotic substances (e.g., soluble sugar and betaine) [9], and inorganic ions (Ca2+) [10]. Plants may acquire cold resistance through cold acclimation [11], in which exposure of plants to low temperatures (0~5 °C lower than threshold temperature and above 0 °C) for a certain period (usually several days or weeks) [12] allows them to adapt to cold stress through phytohormone, antioxidants, osmotic regulation substances, and inorganic ions. They act as intercellular signals linking environmental stimuli to plant responses, and these signal transduction pathways are often accomplished through extensive transcriptional regulation. Understanding the changes in morphological, physiological, and molecular metabolism of plants in response to cold stress is critical for cold-tolerant breeding.

2. Morphological Changes in Response to Cold Stress in Plants

Cold tolerance is an environmental adaptation characteristic resulting from various changes in morphological structure, gene expression, and protein expression caused by the interaction between genes and the environment. In this section, we first consider cold stress in the context of the whole plant’s developmental stages, and then highlight the need to identify and evaluate the varying cold tolerance of different plants.
All stages of plant development, from seed germination, flowering, and fruiting to dormancy, are impacted by cold stress. Various levels of cold stress can influence the same plant tissue at distinct growth stages. There are many plants for which cold damage classification is based on leaf damage [13]. Cold stress leads to insufficient leaf [14] and root [15] development (Table 1). Through principal component analysis, root characteristics may aid in identifying corn (Zea mays L.) hybrids that exhibit cold tolerance [16]. In one study, 15 hybrids (45%) had cumulative root length greater than the average total root length under extreme cold treatment [16]. Kinematic analysis of diurnal growth rates in control and cold-treated corn leaves from germination until the completion of leaf 4 expansion showed that cold nights had an impact on both cell cycle time (+65%) and cell yield (−22%); meanwhile, the size of mature epidermal cells was unaffected [17]. This finding contrasts with the common belief that the reduction in growth caused by abiotic factors is typically attributed to a combination of decreased cell production and reduced mature cell size [17].
Plants undergo a series of physiological and morphological changes to cope with cold stress including increasing malondialdehyde (MDA) content, membrane permeability, proline accumulation, and altering antioxidant enzymes including superoxide dismutase (SOD), ascorbate peroxidase (APX), catalase (CAT), peroxidase (POD) and glutathione peroxidase (GPX) activities; such effects are generally achieved through significant transcriptional regulation [30]. However, an overreaction to any of these mechanisms may hinder seedling growth and biomass production, resulting in a trade-off between cold tolerance and overall yield. The influence of cold stress during early growth stages can significantly affect later growth phases. For example, prolonged cold stress through early planting has been found to induce cellular membrane damage and growth retardation in rice (Oryza sativa L.). via exposure to cold [31]. Plants promote flowering [19] and seed production [20] via the vernalization pathway, thus optimizing breeding [32]. Plants can enhance seed germination through cold stratification [33]. Experiences during one developmental stage can leave lasting impacts on later stages. Notably in arabidopsis (A. thaliana.), rosette vernalization has been shown to boost seed germination across various ecotypes [34] (Table 1). Cold stress at the reproductive stage leads to spikelet sterility [27], limits seed size [28], and ultimately leads to yield reduction (Table 1). For example, with each extra day of freezing at the critical temperature, plant mortality rates increased by 8.6%, 22.3%, 11.1%, and 9.4% for the wheat (Triticum aestivum L.) cultivars Jing411, Nongda211, Zhengmai366, and Yanzhan4110, respectively. Within the same cultivar, tillers demonstrated lower sensitivity to freezing duration (15–32% per day) when compared to younger leaves (25–35% per day) and older leaves (20–55% per day) [35]. Plant cold resistance can be reflected by measuring plant physiological indicators. This method is simple and easy to use. It is often used in cold resistance analysis and identification of plant resources, and is suitable for the screening and identification of various resources. It is also often used in combination with other biochemical markers. Traditional plant cold tolerance identification is still mainly performed by traditional evaluation methods using cold damage classification, and individual physiological and biochemical indicators for cold tolerance identification, which is simple and easy to implement and is usually used for plant resources, cold resistance analysis and identification. Therefore, it is necessary to identify and evaluate the cold tolerance of different plants, which can be reflected by measuring plant physiological indicators; however, some plant cold tolerance phenotypes are difficult to identify accurately, which directly impacts the effectiveness of the forward genetics approach in the discovery of cold tolerance genes.When temperatures drop below 0 °C and fall beneath the plant’s critical threshold, the formation of ice crystals within or on the surface of plant cells invariably leads to cellular damage and eventual plant death as ice crystals develop. Preventing ice crystal formation is regarded as a crucial mechanism of cold tolerance in plants. Cells may suffer mechanical damage or cellular dehydration resulting from cytoplasmic efflux due to an osmotic imbalance caused by the exclusion of solutes from recrystallized ice [36]. Downy birch (Betula pubescens Ehrh.) suffered no injury to any tissues under −70 °C cold stress [37]. Plants vary greatly in cold tolerance. It is of great significance to study the critical temperature of plants. Using the USDA Plant Hardiness Zone Map, the cold tolerance of plants can be preliminarily understood [38]. The impact of cold injury on plants depends on a variety of factors, including the intensity and duration of the cold injury, the stage of plant development at which cold injury occurs, and plant characteristics such as root structure, cuticle wax composition, and cell wall thickness [39]. Trichome coverage and 3D wax projections can be recognized as antifreezing strategies of plants, which increasing the thickness of the cuticle layer, stomatal density, and cuticular permeability [39]. Plants respond to cold stress through a series of physiological and morphological changes, such as ice-binding proteins and antifreezing proteins [36]. The ice-binding microalgal protein CmIBP1 effectively inhibits ice recrystallization both in vitro and in transgenic plants, as the intensity of IRI activity can be compared based on the size of the ice particles at a given temperature and duration [40]. CmIBP1 is secreted into apoplasts and improves freezing tolerance by inhibiting the growth of ice crystals in transgenic plants [40]. The reason for ice nucleation and freezing in plants is unknown. Studying cold perception in plants can provide the direction of ice diffusion from nucleation sites. Screening and cultivation of cold-tolerant germplasm resources based on strategies to prevent ice crystals could reveal ways to improve plant growth and defense.

2.1. Plant Cell Membrane

The normal functions and structures of cells are disrupted by the destruction of the plasma membrane and organelle membrane systems under cold stress. A study pointed out a direct relationship between the structure of plant cell membranes and plant cold tolerance [41]. At present, the mainstream view of plant perception of cold [42] is that cold damage occurs when the cell membrane transforms from a fluid state to a solid state, related to fatty acid changes (Figure 1). This hardening affects membrane proteins, leading to physiological and morphological changes with regulatory responses. Maintaining cell membrane integrity and fluidity is crucial for plants to cope with cold stress. Ice crystal formation can increase electrolyte leakage and causes lipid peroxidation [43]. Cold acclimation involves accumulation of unsaturated fatty acids and phospholipids [18,44], enhancing membrane integrity and cold tolerance (Table 1). Studies have shown that a specific haplotype of the OsSRO1c gene (OsSRO1cHap1) can significantly improve O. sativa cold tolerance at both seedling and heading stages. Further biochemical analysis shows that the OsSRO1c protein exhibits liquid–liquid separation characteristic properties in the nucleus and forms a condensate with OsDREB2B. These condensates can directly respond to cold stress and form small droplet structures through protein phase transformation [45]. The strong haplotype1 encoding the OsSRO1c protein d showed better mobility in condensed droplets formed by OsDREB2B and increased the transcriptional activity of OsDREB2B, thereby activating the first plant low-temperature sensors, including COLD1 [45] (Figure 1). This mechanism provides O. sativa with enhanced cold tolerance.

2.2. Chloroplast

Cold stress has a significant impact on plant growth and development, especially photosynthesis. Maintenance of normal chloroplast physiology is critical for plant growth and development. Under cold stress, the chloroplast structure is severely damaged due to excessive reactive oxygen species (ROS) accumulation (Figure 1). For example, cold stress induces overexpression of chloroplast Mpv17_PMP22 protein (MPD1) in A. thaliana which accelerates ROS generation and cold injury [46]. Chloroplasts are crucial targets for cold acclimation processes. Chlorophyll content is a key indicator of plant cold resistance. Chloroplast acclimation may be the limiting factor for cold adaptation [47]. Chloroplast damage affects photosynthesis and carbohydrate metabolism during cold acclimation [48]. Cold acclimatization in conifers involves seasonal changes in mesophyll cells. Chloroplast movement changes with season, suggesting that temperature drop dynamics depend on organelle movement [49]. To maintain PSII performance under cold stress, plants modify chloroplast structure. The seasonal movement of chloroplasts is mainly affected by low temperature stress. Under greenhouse conditions close to natural light and photoperiod, chloroplasts maintains their activity in the upper plate [49]. The timing of cytoplasmic cluster formation in the two Picea species studied coincides with the minimum seasonal level of the chlorophyll fluorescence parameter, characterizing the efficiency of the photosynthetic apparatus [49]. This period also corresponds to a decrease in grana development within the chloroplasts [49]. Triose phosphate and 3-phosphoglycerate export signals the chloroplast redox state, underlying cold photosynthesis [50]. Plant cold response depends on light quality, as cold exposure may induce PSII photoinhibition and oxidative damage [30]. Plants produce excess energy beyond their utilization capacity under stress [30]. This surplus energy leads to a decrease in the photosynthetic rate and electron transport capacity, resulting in photoinhibition. Although CA treatment can significantly mitigate the degree of PSII photoinhibition and oxidative damage in tobacco (Nicotiana tabacum L.) leaves under cold stress [30], seedlings can adapt to CA by regulating energy dissipation, thereby preventing excessive reduction in plastoquinone pools and subsequent PSII photoinhibition [30]. However, severe stress may progressively exacerbate PSII photoinhibition, if the excess excitation energy is not adequately dissipated through non-photochemical quenching pathways and other electron sinks in a timely manner. This further leads to increased ROS generation, thus damaging the photosynthetic mechanism [30]. The ability of many plant species to express robust phenotypes depends on light and photosynthetic activity during cold growth.

3. Mechanism of Cold Tolerance in Plants

Under cold stress, plants undergo various physiological and biochemical changes such as alterations in cell membrane permeability, osmotic regulators, peroxides, enzymes, and hormones. These changes are integrated into cold-induced signaling cascades to enhance cold tolerance. Plants perceive abiotic stress through key signal transduction pathways: abscisic acid (ABA) signal transduction pathway, Ca2+ signal transduction pathway and mitogen-activated protein kinase (MAPK/MPK) cascade pathway [51]. Cold stress modifies plant osmolytes, hormones, antioxidant systems, and transcription factors, enhancing ROS scavenging and upregulating cold-responsive genes [42]. ROS function as signaling molecules and regulate gene expression. This section explores plants’ cold stress tolerance mechanisms through phytohormones, plant growth regulators, ROS, protein kinases, and Ca2+.

3.1. Phytohormones and Plant Growth Regulators

Phytohormones play central regulatory roles in plant cold stress responses. Phytohormones are signal molecules that comprehensively respond to plant abiotic stress, and their signal transduction mechanism is related to the plant’s response to cold stress. Phytohormone signaling mediates abiotic stress responses through multiple mechanisms, including abscisic acid (ABA), gibberellin (GA), jasmonic acid (JA), salicylic acid (SA), brassinosteroids (BR), ethylene (ET), and melatonin (MET). As signaling molecules, they play a key role in activating the antioxidant enzyme system and regulating the expression of cold response genes, integrating external information into the cell interior, and influencing stress response pathways.

3.1.1. Abscisic Acid (ABA)

The plant hormone ABA plays an important role in plant development and cold response, although its complex network signaling pathways remain unclear. ABA is a tolerance inducer, and its role in cold-signaling pathways can be divided into ABA-dependent pathways and ABA-independent pathways. The current consensus is that both ABA-dependent and ABA-independent pathways are involved in the cold-stress response of plants through the inducer of CBF expression (ICE)—C-repeat binding factor (CBF)—cold regulated (COR) pathway [52] (Figure 2). The study provides evidence that RCAR5/PYL11 exerts two distinct functions in seeds and leaves through Aba-dependent and Aba-independent pathways, respectively [53]. ABA-mediated overexpression of the RCAR5 gene can inhibit both pre-germinative and post-germinative growth under cold stress conditions [53]. Regulation of cold stress in RCAR5 transgenic plants is controlled by OST1 and occurs independently of ABA [53]. Similarly, OST1-mediated phosphorylation of BTF3L positively regulates CBFs [54]. Phosphorylation of BTF3L Ser50 residue by OST1 is a necessary prerequisite for regulating antifreeze function [54]. ABA increases POD activity, SOD activity, and chlorophyll content, and reduces relative conductivity and MDA content to improve plant stability under cold stress [8]. Overexpression of VaPYL9 increases the ability of tomato (Solanum lycopersicum L.) to scavenge ROS, reduces membrane lipid peroxidation and cell desiccation, thereby protecting S. lycopersicum from cold stress [55]. ABA-dependent cold-resistance pathways have been identified in a variety of plants. ABA usually accumulates rapidly under cold stress and regulates the expression of downstream stress response genes by integrating multiple stress signals.

3.1.2. Jasmonic Acid (JA)

JA, as a key signal factor upstream of the ICE–CBF–COR transcriptional pathway, regulates plant cold resistance [56]. Jasmonate zim-domain 1(JAZ1) proteins, inhibitors of the JA signaling pathway (Figure 2), suppress cold stress responses in A. thaliana by inhibiting the transcriptional function of the transcription factor ICE1 [57]. Since JAZ protein contains a Jas domain that can interact with the receptor coronatine insensitive 1 (COI1) and a ZIM domain that can interact with Myelocytomatosis protein 2 (MYC2) of the bHLH family, it can physically interact with ICE1 of the bHLH family and repress its expression, thereby reducing cold tolerance [58]. When plants are subjected to cold stress, intracellular JA concentration increases, and JA-lle is transported to the nucleus through the JAZ1 protein, promoting the interaction of JAZ with the E3–ubiquitin ligase–SCF complex and the COI1 receptor [59,60]. This derepresses the ICE1 transcription factor through the JAZ protein, allowing the expression of downstream JA-responsive genes (Figure 2). Therefore, JA is a very important hormone in plants, which can improve plant resistance to cold stress and can interact with ABA metabolic pathways. Interestingly, in Cucumis melo, exogenous ABA and JA can improve cold tolerance, but their regulatory mechanisms are different. Exogenous ABA can improve cold tolerance of muskmelon by increasing the expression of CmMYC2, CmPYL1 and CmSOD1 genes and decreasing the expression of CmPP2C3 [61] (Figure 2). JA may enhance cold tolerance of melon by decreasing the expression of CmPP2C3, CmJAZ1 and CmMYC2 [61] (Figure 2). MYC2 is a core transcription factor for JA signaling and a hub between ABA, SA, GAs, and IAA signaling [62]. MYC2 positively and negatively regulates multiple functions in the JA signaling pathway [62]. Physical interactions with other key regulatory proteins, formation of heterodimers with other transcription factors, and the ability to activate or repress gene expression in response to multiple signals can contribute to the diverse regulatory roles of MYC2 [62]. However, the antagonism and synergies between JA and other plant hormones in cold stress are unknown, and identifying key components integrating JA and other hormonal pathways is important.

3.1.3. Gibberellin (GA)

GA is a vital plant growth regulator. GA, receptor GID1, and repressor DELLA shape the GA–GID1–DELLA module in the GA signaling cascade. DELLA, a key protein in GA signaling, is involved in CBF-regulated cold induction [63]. DELLA and JAZ interact through the Jas domain [64], and ICE1 is extracted from JAZ linked to DELLA (Figure 2). Also, the CBF3 gene inhibits GA biosynthesis and promotes DELLA accumulation by activating GA2ox7 expression in A. thaliana, participating in CBF-regulated cold induction [65]. In cold temperatures, ICE1 is modified to perform the function of activating CBF3 transcription. CBF3 activates GA2ox7 and reduces bioactive GA levels, thereby promoting DELLA accumulation. Increasing DELLA releases more ICE1 to enhance the next round of CBF3 cold induction [66]. GA3 alone enhances cold resistance of flower buds but decreases it in branches, showing organ-specific effects. ABA/GA3 content better represents plant cold resistance than ABA alone [67]. Exploring plant hormones’ role in cold response, analyzing their regulatory network, and achieving hormone ratios/balances to regulate cold tolerance are crucial research directions.

3.1.4. Auxin (IAA)

Under low temperature stress, exogenous IAA promotes root growth through ARR1/12, but cannot completely compensate for the inhibitory effect of low temperature on initial root growth in A. thaliana ecotype Columbia [68]. OsGH3-2 regulates the homeostasis of endogenous free IAA and ABA and has different effects on drought and cold tolerance of Oryza sativa subsp. japonica [69]. Transient cold stress at the tetrad stage of pollen development in Brassica rapa var. glabra Regel causes auxin-mediated starch-related energy metabolism imbalance that contributes to the decline in pollen germination rate and ultimately seed set [70]. SLR/IAA14 is a transcriptional repressor of IAA signaling and plays a crucial role in microRNA integration into IAA and cold response.

3.1.5. Salicylic Acid (SA)

SA combats abiotic stresses in plants and mitigates cold-induced changes. A new model [71] shows that SA activates ROS signal integration into Ca2+/CPK-dependent ABA signal transduction branches rather than stomatal opening factor 1 (OST1) signal transduction branches (Figure 2). Exogenous SA enhances cold response gene expression, PSII, Fv/Fm, and Pn, and reduces MDA. SA regulates cold-induced changes via ABA-dependent or independent pathways, Ca2+ signaling pathways, MAPKs pathways, and ROS pathways [72]. SA can improve freeze tolerance of Triticum aestivum leaves by affecting apoplastic proteins [73]. Exogenous SA can lead to increased ice nucleation activity at low temperatures (15/10 and 10/5 °C) [73]. SA may participate in cold resistance by regulating the activities of winter wheat (T. aestivum cv. Dogu-88) axoplast proteins and antioxidant enzymes, reducing the activities of CAT and POX, and increasing the activities of polyphenol oxidase [74]. SA-treated cells can maintain Ca2+ homeostasis under cold or heat stress and improve the tolerance of young grapes (Vitis vinifera L.) [75]. The alleviating effect of SA on chilling injury of peaches (Prunus persica (L.) Batch. cv. Beijing 24) during cold storage may be attributed to its ability to induce antioxidant systems and heat shock proteins. Cold stress activates the innate immune response through the SA-dependent pathway. Effects of 10 h cold-temperature treatment in A. thaliana were similar to those caused by pathogen infection, including increased expression of the SA pathway marker genes PR2 and PR5, as well as increased expression of genes that play an active role in defense against (hemi)-biotrophs [76]. After cold stress treatment, transcripts encodings some SA biosynthetic enzymes (but not ICS1/SID2) were more abundant, while transcripts encoding components involved in SA modification were less abundant [76]. However, it is unclear whether long-term cold stress-induced disease resistance is related to early short-term cold stress-induced responses.

3.1.6. Brassinosteroids (BRs)

BRs are widely used in agriculture, which impacts plant response to cold stress [77]. Brassinosteroid (BR) receptor brassinazole-resistant 1 (BZR1) regulates CBF1 and CBF2 gene expression [78], enhancing cold resistance without affecting growth (Figure 2). BRs actively control cold stress response by promoting expression and cold tolerance of CBF1 and CBF-regulated gene COR47 in A. thaliana [79]. PIF4 protein accumulation and transcriptional activity are regulated by light and temperature, linking the exogenous environment with endogenous BR levels [80]. An amount of 1 μM BRs can change the cold acclimation process by stimulating photosynthesis and carbohydrate metabolism, thereby improving cold tolerance of winter rye (Secale cereale L.)[81]. The crosstalk of H2O2 and NO is involved in the BR-induced cold tolerance of Medicago truncatula [82]. PSII efficiency induced by BRs under stressful conditions may depend on NO production rather than H2O2 production [82]. Endogenous BRs do not travel far, thus regulation of BR signal transduction at the specific tissue or organ level is crucial for plants to respond to cold stress.

3.1.7. Ethylene (ET)

ET is involved in regulating a series of biological processes. The ET signal pathway may inhibit the CBF transcriptional cascade in soybeans (Glycine max (L.) Merr.) through the action of ethylene-insensitive protein 3 (EIN3) [83] (Figure 2). Similarly, during cold storage of post-harvest loquat fruit, ethylene signal transducers were involved in the lignification process of cold injury through different regulatory sites to alleviate cold injury [84]. The expression of LeCBF1 is regulated by exogenous ethylene and 1-methylcyclopropene and is not expressed without cold induction in postharvest S. lycopersicum [85]. MtSKL1 is required for detecting changes in ET level in M. truncatula plants for the full development of the cold acclimation response by suppressing expression of MtEIN3 and MtEIN3/EIL1, which in turn downregulates expression of MtERFs, leading to the enhanced tolerance of M. truncatula to freezing by upregulating MtCBFs and MtCAS15 [86]. The MdERF1B–MdCIbHLH1 regulatory module plays a role in ET-mediated cold stress responses in Malus pumila [87] (Figure 2). Cold stress rapidly induces ethylene production and upregulates MdERF1B expression. Additionally, MdERF1B interacts with MdCIbHLH1 to activate MdERF1B-mediated cold tolerance and ethylene biosynthesis [87].

3.1.8. Cytokinin (CTK)

Under transition zone climate conditions, choosing to use seedlings with higher ABA values and CTK values can improve the winter survival and spring vegetation recovery of Cynodon spp. [88]. Verticillium dahliae Aspf2-like protein (VDAL) promoted potato (Solanum tuberosum L.) growth, particularly at low temperatures [89]. Time-course transcriptomic analysis and endogenous phytohormone detection revealed that CTK may play an important role in response to VDAL-promoted growth [89]. Phytohormone IAA and CTK signaling directly or indirectly regulate gravitropism response and root development under low-temperature stress [25]. Furthermore, the protective mechanisms against DNA damage in root stem cells induced by low-temperature stress involves the crosstalk of IAA and CTK. AHK2 and AHK3 and the cold-inducible ARRs play a negative regulatory role in cold stress signaling by inhibiting ABA responses, independent of the cold acclimation pathway in A. thaliana [90] (Figure 2). Moreover, transient expression of a subset of ARR genes, including ARR5, ARR6, ARR7, and ARR15, which are downstream targets of EIN3, may reveal an inherent connection between cold stress and ET [91] (Figure 2). ARRs are important nodes that integrate ET and CTK signals to control plant responses to environmental stress [91].

3.1.9. Melatonin (MET)

Exogenous MET can improve plant growth and tolerance to cold stress by regulating antioxidant enzymes. Pretreatment with 1 μM melatonin restores rhythmic accumulation of hydrogen peroxide, SOD, and CAT in hull-less barley seedlings under cold stress [92]. Melatonin scavenges reactive oxygen species (ROS) by accelerating the AsA-GSH cycle, balancing photosynthetic degradation, and inhibiting ROS production, enhancing cold tolerance in cucumber seedlings [93]. Exogenous MET application can increase the expression of CBFs, COR15a, ZAT10 and ZAT12, thereby improving the growth and cold tolerance of A. Thaliana [94] (Figure 2). MET alleviates cold stress by upregulating the expression of CsZat12 and regulating the metabolism of polyamines and ABA [95]. MET might activate MAPK through H2O2 or Ca2+ dependent pathways in A. thaliana [96] (Figure 2). Activated MAPK can phosphorylate and activate SOG1 and inhibitory MYB to alleviate DNA damage under abiotic stress [96]. Injecting MET into the mother plant during the grain filling period can promote the germination of offspring seeds by accelerating starch degradation, and improve the cold tolerance of seedlings by activating antioxidant enzymes and improving photosynthetic electron transfer efficiency [97]. MET is unstable, and further studies are needed to explore its interactions with other penetrants and hormones under cold stress, as well as MT signaling between organelles.

3.2. Signalling Compounds

3.2.1. Reactive Oxygen Species (ROS)

ROS signaling under cold stress generates excess ROS, leading to oxidative damage and cold stress response in plant cells. Boosting antioxidant enzymes to clear excess ROS is a key strategy for plants to handle cold stress. PCA analysis of mangrove species showed high antioxidant enzyme activity is crucial for cold tolerance [98]. tAPX triggers expression of COR15A, PAL1, and CHS proteins in response to cold stress [99]. Long-term exposure to cold stress showed that sAPX was not relevant and showed a strong dependence on tAPX [98]. Thylakoid protection mediated by tAPX acts as an initiation center that stores initiation information over time [99]. tAPX-mediated thylakoid protection serves as a priming hub, which stores information on priming over time [99]. Compared with plants induced by short-term cold stress, long-term cold stress induced stronger induction of non-chloroplast-specific ROS-regulated genes such as CHS and PAL1 (and COR15A), which support salicylic acid, lignin, flavonoids, and floral biosynthesis of various secondary stress protective mediators such as anthocyanins [99]. Heme-associated protein AtHAP5A enhances cold tolerance and suppresses ROS accumulation by binding AtXTH21 [100]. Genetic evidence indicated that AtHAP5A acts upstream of AtXTH21 in freezing stress response in A. thaliana [100]. These results revealed that AtHAP5A modulates freezing stress resistance through interaction with the CCAAT motif of AtXTH21 in A. thaliana [100]. Overexpressing AtHAP5A and AtXTH21 could alleviate 4 °C stress-induced ROS accumulation and related oxidative damage in A. thaliana, while AtHAP5A and AtXTH21 mutants had the opposite effect [100]. Overexpression of the ROS signal response gene AtZAT12 downregulates AtCBF1/2/3 genes under cold stress [101,102] (Figure 2). ZAT12 downregulated the expression of the CBF genes indicating a role for ZAT12 in a negative regulatory circuit that dampens expression of the CBF cold response pathway [101]. The role of ZAT12 regulation may help plants cope with oxidative stress. In this regard, it is of great interest that ZAT12 expression resulted in downregulation of transcripts encoding a putative l-ascorbate oxidase [101]. Cold triggers ROS accumulation in plant cells, reducing protein content, enzyme activities, and expression of cold resistance genes through biochemical reactions.

3.2.2. Protein Kinases

Histidine protein kinase (HPK) may act as a cold sensor. Upon receiving external stimuli, histidine residues are phosphorylated and passed to response regulatory proteins (RR), which receive phosphate groups and transmit signals. The second messenger regulates kinases like Ca2+-dependent protein kinases (CDPKs) and receptor-like protein kinases (RLKs) (Figure 1), while protein phosphatase (PP) regulates proteins and transmits signals to downstream cold-signaling pathways. OsCPK24, a cold-reactive kinase on the cell membrane, enhances cold tolerance by stimulating thiol transferase activity of OsGrx10 in response to proline and Ca2+, regulating glutathione levels to combat cold stress [103].
The mitogen-activated protein kinase (MAPK/MPK) cascade is an important way to sense cold signals and regulate gene expression [104]. H2O2 is a signal molecule to excite it [105]. The AtCRLK1-AtMEKK1/2-AtMPK4/6 cascade improves cold tolerance by antagonizing the AtMPK3/6 pathway [106,107]. MtCTLK1 or MfCTLK1 affect cold resistance through the CBF transcriptional cascade, antioxidant defense, and proline accumulation [108].

3.2.3. Ca2+

Cold stress increases intracellular second messenger Ca2+, which is transmitted to downstream cold-signaling pathways, leading to transcriptional remodeling in plants. Calmodulin (CaM) and CaM-like proteins (CML), Calcineurin B-like (CBL), and calcium-dependent protein kinases (CPKs), as three calcium signaling receptor proteins, are mainly involved in the perception and transmission of cold signals in plants [109].
Calmodulin-binding transcription activator (CAMTA), MYB, WRKY, NAC, basic leucine zipper (bZIP), and MADS-box are calmodulin-binding proteins (CaMBPs) (Figure 3). Cyclic nucleotide-gated channels (CNGCs), nonselective cation channels in plant cell membranes, are key components of plant temperature perception [110,111]. Previous analyses of plant CNGC family protein structures found that they all contain CaM-binding sites and that there is a physical link or functional interaction between CaM and CNGC [112,113].
CBL is a calcium-binding protein involved in plant signaling. AtCBL1-AtCIPK3 regulate cold stress tolerance in A. thaliana by transmitting signals to cold resistance-related transcription factors [114]. CBL2, CBL3, CIPK9, and CIPK27 negatively regulate the ABA signal pathway in A. thaliana [115]. OsCBL2 is the only CBL up-regulated in the O. sativa aleurone layer induced by GA [116]. AtCBL3 interacts with AtMTN1 to regulate ethylene biosynthesis and polyamines during plant growth [117]. CBL6/CBL8-CIPK14 is uniquely responsive to cold in citrus [118]. Interestingly, HvCBL4 maintains circadian rhythm at different temperatures, aiding in studying the effect of the biological clock on grain cold resistance [119].
In contrast to CaM and CBL, Ca2+ signals can be directly converted into a single CPK protein phosphorylation event [120]. In 1982, CPK was first identified in peas (Pisum sativum L.) [121]. Some CPKs are Ca2+-dependent, while others are Ca2+-independent (such as AtCPK7 and AtCPK32) [122]. CPK interacts with plant hormones during cold stress [123,124]. CPK11 is induced by H2O2 and regulates antioxidant enzymes through the ABA signal pathway [125,126]. Overexpression of VaCPK20, a calcium-dependent protein kinase gene, enhances drought and cold tolerance by regulating proteins [127]. In O. sativa, OsCPK24 is a cold-reactive kinase that enhances cold tolerance by regulating glutathione levels [103].

4. Downstream Mechanisms and Regulation for Cold Signaling in Plants

Cold tolerance in plants is a quantitative trait controlled by multiple genes, which are expressed synergistically under the regulation of transcription factors, and multiple signaling pathways jointly regulate plant responses to cold stress (Figure 2) [128]. The transcriptional regulation of plant cold-tolerance genes can be divided into two types: CBF-dependent and CBF-independent [129]. The CBF-regulated signaling pathway is currently the most studied transcription factor regulating cold-induced gene changes.

4.1. Transcription Factors in Plant Cold Signaling

Transcription factors have always been a hot spot in plant cold stress research, and they play an important regulatory role in signal transduction and gene expression under cold stress [130]. They bind to cis-acting elements upstream of stress-responsive genes to activate or repress gene expression [131]. For example, MYB, WRKY, NAC, and bZIP are the transcription factors which belong to calmodulin-binding proteins, and play an important role in plant cold induction (Figure 3) [132]. Hypothermia receptors mostly trigger calcium ions. It is important to create a suitable expression pattern for positive regulators of cold resistance without negatively affecting favorable agronomic traits. However, studies on these transcription factors have mainly focused on their functional mechanism, and little is known regarding the effects of phytohormones on the activity of these proteins.

4.1.1. AP2/ERF

Aptala2/ethylene response factor (AP2/ERF) is the most studied transcription factor involved in the mechanism of cold response [133]. C-repeat binding factor/dehydration response element binding factor 1 (CBF/DREB1) belongs to its subfamily, and there are many studies on its regulation under cold stress [51]. Most studies on cold signaling in plants still revolve around the ICE–CBF–COR transcriptional pathway [134]. Through this pathway, CBF transcription factors activate COR gene expression by binding to CRT/DRE elements in the COR promoter, enhancing cold tolerance [135]. ABA increases CBF transcriptional level, possibly through binding to CRT/DRE elements [136], which activates COR genes to improve cold tolerance. There are three main CBF genes: CBF1, CBF2, and CBF3. AtCBF1/3 transcription factors improve cold tolerance by binding to CRT/DRE elements of COR genes [137] (Table 2). However, AtCBF2 negatively regulates CBF1 and CBF3, reducing cold resistance [138]. The mechanism by which CBF mediates cold stress and its regulatory mechanism need to be explored. Overexpression of CBF enhances cold resistance but may affect plant growth. CBF transcription factors are also regulated by hormones (Figure 2).

4.1.2. MYB

MYB transcription factors play a two-way role in plant stress response (Table 2). On one hand, transcription factors A. thaliana AtMYB15 [139,140], Musa nana MpMYBS3 [141], and O. sativa OsMYBS3 [142] are inhibitors of cold signals by inhibiting the expression of CBF genes under cold stress. On the other hand, overexpression of MYB96 in plants can result in strong cold tolerance [143]. Interestingly, while the MYB15 transcription factor negatively regulates the CBF genes in A. thaliana [139] and pepper (Capsicum annuum L.) [144], SlMYB15 transcription factor positively regulates the CBF genes, which shows MYB15 in different plants has different responses to cold stress [145]. The OsMYB30 transcription factor regulates starch decomposition and cold tolerance by negatively regulating the expression of the α-amylase 1a (OsAMY1a) gene in O. sativa [146]. OsMYB30-OsTPP1 is a sugar signaling pathway that regulates seed germination in response to low temperature. Expression of OsMYB30 and OsTPP1 was induced by low temperatures during seed germination [146]. OsMYB30 binds to the promoter region of OsTPP1 to activate its expression [146]. Overaccumulation of trehalose was found in both OsMYB30- and OsTPP1-overexpressing lines, resulting in inhibition of OsAMY1a during seed germination [146] (Table 2). The CaMYB306 transcription factor inhibits the positive cold resistance regulator CaCIPK13 in C. annuum [147]. It also suppresses chlorophyll and anthocyanin contents and regulates ROS signaling.

4.1.3. WRKY

In 1994, Ishiguro and Nakamura found the first WRKY transcription factor SPF1 in sweet potato (Ipomoea batatas (L.) Poir.) [148]. WRKY mainly responds to a variety of induced stimuli by specifically recognizing W-Box and initiating the expression of regulatory genes, thus playing a regulatory role in plants [149]. The transcription factor AtWRKY34 (Table 2), specifically expressed in pollen, may participate in the cascade of CBF signals in mature pollen and negatively mediate the cold sensitivity of mature A. thaliana pollen [150]. Similarly, WRKY22 transcription factor is essential for cold stress adaptation in A. thaliana buds (Table 2). It alters SA-mediated wounding and osmotic stress responses in A. thaliana [151]. It suppresses the transcription of WRKY53 and WRKY70. WRKY53 gene expression induces SA expression [152,153]. In A. thaliana, AtWRKY70, downstream of SA receptor NPR1, is regulated by SA and JA [154]. Glutathione regulates MPK3 expression through WRKY40 [155]. Bermuda grass (Cynodon dactylon (L.) Pers.) CdWRKY2 and S. lycopersicum SlWRKY50 increase cold resistance via a CBF-dependent pathway [156,157]. In cucumber (Cucumis sativus Linn.), CsWRKY46 [158] enhances cold resistance via an ABA-dependent pathway (Table 2). Future research should focus on mining more target genes and exploring the interaction between transcription factors, WRKY, and other proteins to understand the regulatory mechanism.

4.1.4. bZIP

ACGT element-specific plant DNA-binding proteins belong to the bZip domain [159]. ABERs are bZIP transcription factors. AREB/ABF transcription factors regulate ABA-related gene expression and cold stress. Proteins like ABF1, ABF2/AREB1, and ABI5 are involved in ABA signal transduction under stress [160]. Chorispora bungeana (Chorispora bungeana Fisch. et Mey.) CbABF1 confers stress tolerance in N. tabacum by reducing ROS and enhancing antioxidant enzymes [161]. Actinidia eriantha Benth. AcePosF21 [162] reduces ROS damage by mediating AceGGP3 expression (Table 2). The transcription factor ZmbZIP68 [163] inhibits the CBF-signaling pathway in Z. mays (Table 2). In A. thaliana, tea (Camellia sinensis (L.) O. Kuntze) CsbZIP18 [164] negatively regulates cold tolerance through an ABA-dependent but not CBF-dependent pathway (Table 2). LIP19/OsOBF1 [165] heterodimer formation improves cold tolerance in O. sativa (Table 2). GmbZIP44, GmbZIP62, and GmbZIP78 play a negative role in ABA signaling by up-regulating ABI1 and ABI2 proteins, inducing genes like ERF5KIN1, COR15, and COR78 for salt and cold tolerance in transgenic plants [166].
Elongated hypocotyl5 (HY5) is a key bZIP transcription factor involved in crosstalk between light and cold response pathways, integrating ABA and ROS signaling to attenuate photo inhibition (Figure 2). The SlHY5 transcription factor binds to the promoter of ABI5, triggering respiratory burst oxidase homolog1 (RBOH1)-dependent H2O2 production in the apoplast [167] (Figure 2). The CRYPTOCHROME2 (CRY2)-COP1-HY5-BBX7/8 module regulates blue light-dependent cold acclimation in A. thaliana [168]. Additionally, BBX7 and BBX8 act as HY5 targets, positively regulating freezing tolerance by modulating cold-responsive gene expression, primarily in CBF-independent pathways [168].

4.1.5. NAC

NAC (NAM, ATAF1/2, CUC2) is one of the largest families of transcription factors endemic to plants [169]. A. thaliana NAC transcription factor JUNGBRUNNEN1 (AtJUB1) inhibits GA3ox1 and DWF4, and reduces PIF4 expression [170]. GA mutant cpk28 alters expression of NAC transcriptional regulators and GA3ox1 [171,172]. NAC transcription factors participate in plant growth, development, and cold stress response through hormone signal cross-pathways. MaNAC25 and MaNAC28 form a positive feedback loop, negatively regulating cold tolerance in banana (M. nana Lour.) fruit by up-regulating phospholipid degradation gene expression (Table 2) [172]. The CaNAC1 transcription factor can also negatively regulate cold stress in plants through phospholipid degradation (Table 2) [173]. C. annuum CaNAC064 and siberian crabapple (Malus baccata Borkh) MbNAC25 transcription factors regulate plant cold tolerance by enhancing ROS scavenging capacity in A. thaliana (Table 2) [174,175]. HuNAC20 and HuNAC25 transcription factors have cold tolerance in transgenic A. thaliana [176].

4.1.6. bHLH

bHLH is a conserved basic/helix–loop–helix domain with a basic amino acid region and a helix–loop–helix region [177]. ICE1 and ICE2 (e.g., Hevea brasiliensis (Willd. ex A. Juss.) Müll. Arg. HbICE2) bHLH-type transcription factors are positive regulatory factors of CBF [178]. Snow lotus (Saussurea involucrata (Kar. & Kir.) Sch. Bip.) SiICE1 (Table 2) binds to MYC in the CBF3 promoter, promoting CBF3 expression and cold tolerance in S. lycopersicum [179,180]. MeJA positively regulates HbICE2 overexpression in A. thaliana, enhancing cold tolerance [178]. Overexpression of ZjICE2 improves chlorophyll content, photosynthetic efficiency, and cold resistance injapanese lawngrass (Zoysia japonica Steud.) (Table 2) [181]. IbbHLH116 functions like ICE1, conferring cold tolerance to I. batatas (Table 2) [182]. The MdbHLH4 transcription factor represses MdCBF1/3 expression, promotes MdCAX3L-2 expression, and negatively regulates cold tolerance in M. pumila domestica (Table 2) [183].
Phytochrome-interacting factor (PIF) belongs to the 15th subfamily of the bHLH transcription factor family. PIF3, the first plant pigment interaction factor, was discovered in 1998 [184]. Under cold stress, red/far red light induces S. lycopersicum PIF4 accumulation via phytochrome A (phyA). SlPIF4 binds to the S. lycopersicum CBF promoter to promote expression and positively regulate cold tolerance [185]. Additionally, SlPIF4 activates SlDELLA4 expression in the GA signaling pathway, with SlGAI4 regulating tomato cold tolerance (Figure 2). When SlGAI4 accumulates, it inhibits SlPIF4, forming a negative feedback loop, and PIF4 transcription factor, forming a negative feedback regulation loop (Table 2) [185]. Further research is needed on PIF’s role in balancing stress and growth through hormonal pathways and adapting to changing environments.
MYC2 is also a bHLH transcription factor and the main regulator of the JA signal pathway. The PtrMYC2 (Table 2) transcription factor integrates JA signals by directly regulating the accumulation of PtrBADH-1 and GB in trifoliate orange (Poncirus trifoliata (L.) Raf.), thus regulating cold induction [186]. The SIMYC2 (Table 2) transcription factor may be involved in the cold tolerance induced by MeJA, possibly by improving the antioxidant enzyme system and increasing the levels of proline and lycopene in fruits [187].
Table 2. Identified gene resources related to cold response in plants.
Table 2. Identified gene resources related to cold response in plants.
FamiliesGenesSpeciesCold Stress RegulationReferences
AP2/ERFAtCBF1/DREB1BA. thaliana (L.) Heynh.Positive[137]
AtCBF2/DREB1CA. thaliana (L.) Heynh.Negative[138]
AtCBF3/DREB1AA. thaliana (L.) Heynh.Positive[137]
MYBAtMYB15A. thaliana (L.) Heynh.Negative[139,140]
SlMYB15S. lycopersicum L.Positive[145]
AtMYB96A. thaliana (L.) Heynh.Positive[188]
MpMYBS3M. nana Lour.Positive[141]
OsMYB30O. sativa L.Negative[189]
CaMYB306C. annuum L.Negative[147]
WRKYAtWRKY34A. thaliana (L.) Heynh.Negative[150]
AtWRKY22A. thaliana (L.) Heynh.Positive[151]
AtWRKY40A. thaliana (L.) Heynh.Negative[155]
SlWRKY50S. lycopersicum L.Positive[156]
CdWRKY2C. dactylon (L.) Pers.Positive[157]
CsWRKY46C. sativus L.Positive[158]
bZIPCbABF1C. bungeana Fisch. et Mey.Positive[161]
AcePosF21A. eriantha Benth.Positive[162]
ZmbZIP68Z. mays L.Negative[163]
CsbZIP18C. sinensis (L.) O. KuntzeNegative[164]
OsbZIP38/LIP19O. sativa L.Positive[165]
SlHY5S. lycopersicum L.Positive[145,167]
NACAtJUB1A. thaliana (L.) Heynh.Negative[170]
MaNAC25M. nana Lour.Negative[172]
MaNAC28M. nana Lour.Negative[172]
CaNAC1C. annuum L.Negative[173]
CaNAC064C. annuum L.Positive[174]
MbNAC25M. baccata Borkh.Positive[175]
bHLHHbICE2H. brasiliensis (Willd. ex A. Juss.) Müll. Arg.Positive[178]
SiICE1S. involucrata (Kar. & Kir.) Sch. Bip.Positive[180]
ZjICE2Z. japonica Steud.Positive[181]
IbbHLH116I. batatas (L.) Poir.Positive[182]
MdbHLH4M. pumila Mill.Negative[183]
SlPIF4S. lycopersicum L.Positive[185]
PtrMYC2P. trifoliata (L.) Raf.Positive[186]
SlMYC2S. lycopersicum L.Positive[187]

4.2. Post-Translational Modification in Plant Cold Signaling

4.2.1. Sumoylation (SUMO)

Under cold stress, the ubiquitin-like modifier E3 ubiquitin ligase SIZ1 positively regulates the ICE1 transcription factor via sumoylation (SUMO) [190,191] (Figure 2). High HOS1 expression negatively regulates ICE1 through ubiquitin modification, affecting CBF expression [192]. BIN2 phosphorylates ICE1, promoting its interaction with E3 ubiquitin ligase HOS1, leading to ICE1 degradation and reduced CBF expression [193].

4.2.2. Protein Phosphorylation

Protein phosphorylation is a post-translational modification in plant cold signaling. CBL9-CIPK3 phosphorylates ABI1 to form CBL9-CIPK3-ABR1 in A. thaliana [194]. Calcium-dependent protein kinase 21 phosphorylates 14-3-3 proteins in response to ABA signaling and salt stress in rice [195]. Cold stress activates OST1, which regulates CBF stability through the OST1-BTF3 complex [54] (Figure 2). OST1 phosphorylates BTF3 and BTF3L to promote CBF stability under cold stress (Figure 2). MaNAC1 is a target of MaICE1 and interacts with MaCBF1 to enhance cold stress response. MaICE1 phosphorylation and cold stress improve MaNAC1 binding, leading to cold tolerance in M. nana through the ICE-CBF-COR pathway [196].

4.2.3. Mitogen-Activated Protein Kinase (MAPK/MPK) Cascade

In A. thaliana, the MAPKKK protein AtANP1 triggers the phosphorylation cascade with AtMPK3 under cold stress [197]. The AtCRLK1–AtMEKK1/2–AtMPK4/6 cascade enhances cold tolerance by counteracting the AtMPK3/6 pathway [106]. AtMPK3/6 phosphorylates AtMYB15 and reduces the binding affinity of the AtCBF3 gene, resulting in enhanced transcriptional inhibition of CBF3 by MYB15, negatively regulating cold tolerance [139]. Glutathione regulates MPK3 expression through WRKY40 [155]. MPK3/6, negatively regulated by MKP1 [198], plays a key regulatory role during cold acclimation and may enhance freezing resistance and plant survival [199]. It phosphorylates CAMTA3, contributing to its degradation and inhibition of downstream target gene expression [200].

4.3. Multi-Omics Analysis Facilitates the Identification of Cold Resistance Genes

Advances in plant multi-omics research have promoted the discovery of plant cold tolerance molecular mechanisms. Using transcriptomes, S. lycopersicum SlWRKYs were induced by cold stress with CRT/DRE elements in SlWRKY2 and SlWRKY50 promoters, suggesting their role in CBF-mediated cold response [156]. One-month-old A. thaliana was exposed for 1 week to 4 °C in short-day conditions under white (100 and 20 μmol m−2s−1), blue, or red (20 μmol m−2s−1) light conditions [201]. Proteomic data showed distinct differences between red and blue light-treated plants under cold stress, with blue light activating cold-related proteins and red light upregulating chloroplast proteins [201]. Comparative proteomic analysis of chloroplast proteins in sugar beet (Beta vulgaris L.) provided insights into cold effects. It copes with cold stress by transporting photosynthetic proteins, forming starch granules, and scavenging ROS [202]. Advances in plant multi-omics, such as genomics, transcriptomics, proteomics, microbial functional genomics, epigenetics, and metabolomics, have accelerated the discovery of the molecular mechanisms behind plant cold tolerance [203].

5. Conclusions and Future Perspectives

In conclusion, this review discusses recent research into the ways some plants cope with cold stress, and regulatory mechanisms of cold tolerance in plants. Cold stress can quickly induce the expression of many transcription factors, thus activating a large number of downstream cold response gene expression transcription factors that play a key role in regulating gene expression (Figure 3). Plant hormones as signaling molecules are involved in the regulation of plant cold response, and the coordination of hormone and cold-signaling pathways can better deal with cold stress (Figure 4). At present, these studies focus on the function of a single gene, most of which are CBF-dependent pathways. However, the cold response transcriptional group regulated by CBFs accounts for only about 12% [204]. Therefore, further screening of various cold-related genes and the interaction mechanism between them is the key to exploring the mechanism of signal transduction and regulation in the future.
At present, the mechanisms of cold response regulation and control networks are not fully elaborated, and cannot be implemented in production practices. The phenotypes of different growth stages may be different. It is necessary to explore the evaluation system to use future genetic methods to mine excellent cold tolerance genes, especially cold response genes specific to a single species. Tissue-specific expression or induced expression is mainly used to reduce the effect of the constitutive expression on the growth of the plant sensors. How plants perceive cold is still unclear. Finding the upstream temperature signal sensor of the plant cold-signaling pathway is a fundamental problem in future research because these receptors are very likely to play a switching function for the whole cold-signaling pathway. Interestingly, there is a close relationship between drought stress and cold stress. MYB transcription factor gene ApMYB77 [205] and MbMYB4 [206] confer both freezing and drought tolerance. MdBES1 was a positive regulator for cold tolerance and disease resistance in M. baccata Borkh, but a negative regulator for drought tolerance [207]. Dong et al. (2019) discovered the regulation of leaf-derived jasmonic acid as a long-distance transport signal to regulate water uptake by cotton (Gossypium hirsutum L.) roots and invented a liquid fertilizer that promotes jasmonic acid synthesis under partial root-zone irrigation to cope with drought [208]. But there are still few corresponding hormone fertilizers for cold stress, so further research is necessary for the construction and mining of the cold stress network. This will be helpful to develop cultivation and management methods to enhance plant cold tolerance (such as the application of exogenous substances) to improve production.
An increase in extreme weather will amplify the impact of cold stress on agricultural production. A future challenge for improving crop cold tolerance is to combine knowledge gained from model systems with multi-omic and genetic data for new crop varieties and crop performance test systems. In addition, efforts should be made to identify natural cultivars of unknown stress-resistant resources and understand their underlying mechanisms. As more powerful resources are discovered and identified, these efforts must be effectively integrated into plant breeding to achieve sustainable global food security.

Author Contributions

Y.F. contributed to acquisition of information, and writing of the manuscript. Z.L., X.K., A.K., N.U. and X.Z. contributed to guidance in the design, critical revision, and correction of the manuscript. X.Z. contributed to the conception and design, and interpretation, draft and revision of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Programs for Science and Technology Development in Henan Province of China (242102111145), the Postgraduate Education Reform and Quality Improvement Project of Henan Province (YJS2024SZ20 and YJS2023JC16), and the Henan Center of Outstanding Overseas Scientists (GZS2024018).

Conflicts of Interest

The authors declare that this research was conducted in the absence of any commercial or financial relationships that could be construed as potential conflicts of interest.

References

  1. Karkute, S.G.; Krishna, R.; Ansari, W.A.; Singh, B.; Singh, P.M.; Singh, M.; Singh, A.K. Heterologous expression of the AtDREB1A gene in tomato confers tolerance to chilling stress. Biol. Plant. 2019, 63, 268–277. [Google Scholar] [CrossRef]
  2. Qin, X.H.; Li, W.K.; Duan, Y.S. Research progress of plant cold stress signal. Mol. Plant Breed. 2018, 16, 7187–7194. [Google Scholar] [CrossRef]
  3. Liu, L.L.; Ji, H.T.; An, J.P.; Shi, K.J.; Ma, J.F.; Liu, B.; Tang, L.; Cao, W.X.; Zhu, Y. Response of biomass accumulation in wheat to low-temperature stress at jointing and booting stages. Environ. Exp. Bot. 2019, 157, 46–57. [Google Scholar] [CrossRef]
  4. Liu, W.Q.; Zhang, R.Y.; Xiang, C.G.; Zhang, R.Y.; Wang, Q.; Wang, T.; Li, X.J.; Lu, X.H.; Gao, S.L.; Liu, Z.X. Transcriptomic and physiological analysis reveal that α-linolenic acid biosynthesis responds to early chilling tolerance in pumpkin rootstock varieties. Front. Plant Sci. 2021, 12, 669565. [Google Scholar] [CrossRef]
  5. Ding, Y.; Wang, X.T.; Wang, F.; Shao, Y.L.; Zhang, A.M.; Chang, W. The effects of chilling stress on antioxidant enzymes activities and proline, malondialdehyde, soluble sugar contents in three Paphiopedilum species. Russ. J. Plant Physiol. 2023, 70, 61. [Google Scholar] [CrossRef]
  6. Angmo, D.; Sharma, S.P.; Kalia, A.; Brar, N.S.; Bhardwaj, V. Effect of cold stress on field performance, chlorophyll fluorescence, electrolyte leakage and leaf gas exchange parameters of potato (Solanum tuberosum L.) genotypes. Potato Res. 2023, 66, 641–661. [Google Scholar] [CrossRef]
  7. Penfield, S.; Warner, S.; Wilkinson, L. Molecular responses to chilling in a warming climate and their impacts on plant reproductive development and yield. J. Exp. Bot. 2021, 72, 7374–7383. [Google Scholar] [CrossRef]
  8. Guo, J.H.; Beemster, G.T.S.; Liu, F.L.; Wang, Z.M.; Li, X.N. Abscisic Acid Regulates Carbohydrate Metabolism, Redox Homeostasis and Hormonal Regulation to Enhance Cold Tolerance in Spring Barley. Int. J. Mol. Sci. 2023, 24, 11348. [Google Scholar] [CrossRef]
  9. Liu, L.; Gai, Z.J.; Qiu, X.; Liu, T.H.; Li, S.X.; Ye, F.; Jian, S.L.; Shen, Y.H.; Li, X.N. Salt stress improves the low-temperature tolerance in sugar beet in which carbohydrate metabolism and signal transduction are involved. Environ. Exp. Bot. 2023, 208, 105239. [Google Scholar] [CrossRef]
  10. Awan, A. Calcium and boron effect on production and quality of autumn potato crop under chilling temperature. Commun. Soil Sci. Plant Anal. 2021, 52, 375–388. [Google Scholar] [CrossRef]
  11. Vaitkevičiūtė, G.; Aleliūnas, A.; Gibon, Y.; Armonienė, R. The effect of cold acclimation, deacclimation and reacclimation on metabolite profiles and freezing tolerance in winter wheat. Front. Plant Sci. 2022, 13, 959118. [Google Scholar] [CrossRef]
  12. Thomashow, M.F. Plant cold acclimation: Freezing tolerance genes and regulatory mechanisms. Annu. Rev. Plant Biol. 1999, 50, 571–599. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, Y.Y.; Zhou, Y.D.; Shen, L.; Zhao, E.P.; Yang, X. Effect of grafting on cold tolerance of eggplant. J. Anhui Agric. Sci. 2022, 50, 52–55. [Google Scholar] [CrossRef]
  14. Ratajczak, K.; Sulewska, H.; Panasiewicz, K.; Faligowska, A.; Szymańska, G. Phytostimulator application after cold stress for better maize (Zea mays L.) plant recovery. Agriculture 2023, 13, 569. [Google Scholar] [CrossRef]
  15. Min, X.Y.; Liu, Z.P.; Wang, Y.R.; Liu, W.X. Comparative transcriptomic analysis provides insights into the coordinated mechanisms of leaves and roots response to cold stress in common vetch. Ind. Crop. Prod. 2020, 158, 112949. [Google Scholar] [CrossRef]
  16. Wijewardana, C.; Hock, M.; Henry, B.; Reddy, K.R. Screening corn hybrids for cold tolerance using morphological traits for early-season seeding. Crop Sci. 2015, 55, 851–867. [Google Scholar] [CrossRef]
  17. Rymen, B.; Fiorani, F.; Kartal, F.; Vandepoele, K.; Inzé, D.; Beemster, G.T.S. Cold nights impair leaf growth and cell cycle progression in maize through transcriptional changes of cell cycle genes. Plant Physiol. 2007, 143, 1429–1438. [Google Scholar] [CrossRef]
  18. Juurakko, C.L.; Walker, V.K. Cold acclimation and prospects for cold-resilient crops. Plant Stress 2021, 2, 100028. [Google Scholar] [CrossRef]
  19. Fan, S.H.; Liu, H.F.; Liu, J.; Hua, W.; Li, J. BnGF14-2c positively regulates rlowering via the vernalization pathway in semi-winter rapeseed. Plants 2022, 11, 2312. [Google Scholar] [CrossRef]
  20. Khatun, L.; Karim, M.R.; Talukder, F.U.; Rahman, M.S. Combined effects of vernalization and gibberellic acid on quality seed production of summer onion (Allium cepa L.). Agric. Sci. 2020, 2, 148. [Google Scholar] [CrossRef]
  21. Lainé, C.M.S.; AbdElgawad, H.; Beemster, G.T.S. Cellular dynamics in the maize leaf growth zone during recovery from chilling depends on the leaf developmental stage. Plant Cell Rep. 2024, 43, 38. [Google Scholar] [CrossRef] [PubMed]
  22. Mazur, M.; Matoša Kočar, M.; Jambrović, A.; Sudarić, A.; Volenik, M.; Duvnjak, T.; Zdunić, Z. Crop-specific responses to cold stress and priming: Insights from chlorophyll fluorescence and spectral reflectance analysis in maize and soybean. Plants 2024, 13, 1204. [Google Scholar] [CrossRef] [PubMed]
  23. Driever, S.M.; Mossink, L.; Ocaña, D.N.; Kaiser, E. A simple system for phenotyping of plant transpiration and stomatal conductance response to drought. Plant Sci. 2023, 329, 111626. [Google Scholar] [CrossRef] [PubMed]
  24. Pacheco, J.M.; Gabarain, V.B.; Lopez, L.E.; Lehuedé, T.U.; Ocaranza, D.; Estevez, J.M. Understanding signaling pathways governing the polar development of root hairs in low-temperature, nutrient-deficient environments. Curr. Opin. Plant Biol. 2023, 75, 102386. [Google Scholar] [CrossRef]
  25. Tiwari, M.; Kumar, R.; Subramanian, S.; Doherty, C.J.; Jagadish, S.V.K. Auxin–cytokinin interplay shapes root functionality under low-temperature stress. Trends Plant Sci. 2023, 28, 447–459. [Google Scholar] [CrossRef]
  26. Zhou, Y.; Sommer, M.L.; Hochholdinger, F. Cold response and tolerance in cereal roots. J. Exp. Bot. 2021, 72, 7474–7481. [Google Scholar] [CrossRef]
  27. Lin, F.F.; Li, C.; Xu, B.; Chen, J.; Chen, A.H.; Hassan, M.A.; Liu, B.B.; Xu, H.; Chen, X.; Sun, J.Q. Late spring cold reduces grain number at various spike positions by regulating spike growth and assimilate distribution in winter wheat. Crop J. 2023, 11, 1272–1278. [Google Scholar] [CrossRef]
  28. Wang, R.J.; Bai, Y.G.; Low, N.H.; Tanino, K. Seed size variation in cold and freezing tolerance during seed germination of winterfat (Krascheninnikovia lanata) (Chenopodiaceae). Botany 2006, 84, 49–59. [Google Scholar] [CrossRef]
  29. Dong, Z.; Wang, H.; Li, X.; Ji, H. Enhancement of plant cold tolerance by soybean RCC1 family gene GmTCF1a. BMC Plant Biol. 2021, 21, 369. [Google Scholar] [CrossRef]
  30. Wei, Y.L.; Chen, H.Z.; Wang, L.; Zhao, Q.; Wang, D.; Zhang, T.E. Cold acclimation alleviates cold stress-induced PSII inhibition and oxidative damage in tobacco leaves. Plant Signal. Behav. 2022, 17, 2013638. [Google Scholar] [CrossRef]
  31. Martini, L.F.; Nilda, R.B.; Tseng, T.M.; Fipke, M.V.; Noldin, J.A.; Avila, L.A.D. Acclimation to cold stress reduces injury from low temperature and bispyribac-sodium on rice. Pest Manag. Sci. 2021, 77, 4016–4025. [Google Scholar] [CrossRef] [PubMed]
  32. Cha, J.K.; O’Connor, K.; Alahmad, S.; Lee, J.H.; Dinglasan, E.; Park, H.; Lee, S.M.; Hirsz, D.; Kwon, S.W.; Kwon, Y. Speed vernalization to accelerate generation advance in winter cereal crops. Mol. Plant 2022, 15, 1300–1309. [Google Scholar] [CrossRef] [PubMed]
  33. Song, Y.; Li, X.H.; Zhang, M.Y.; Xia, G.W.; Xiong, C. Effect of cold stratification on the temperature range for germination of Pinus koraiensis. J. For. Res. 2023, 34, 221–231. [Google Scholar] [CrossRef]
  34. Auge, G.A.; Blair, L.K.; Neville, H.; Donohue, K. Maternal vernalization and vernalization-pathway genes influence progeny seed germination. New Phytol. 2017, 216, 388–400. [Google Scholar] [CrossRef]
  35. Zheng, D.X.; Yang, X.G.; Mínguez, M.I.; Connor, D.J.; Mu, C.Y.; Guo, E.J.; Chen, X. Tolerance of different winter wheat cultivars to prolonged freezing injury at their critical temperatures. Crop Sci. 2018, 58, 1740–1750. [Google Scholar] [CrossRef]
  36. Bredow, M.; Walker, V.K. Ice-binding proteins in plants. Front. Plant Sci. 2017, 8, 325466. [Google Scholar] [CrossRef]
  37. Welling, A.; Rinne, P.; Aarnio, A.V.; Soppela, S.K.; Heino, P.; Palva, E.T. Photoperiod and temperature differentially regulate the expression of two dehydrin genes during overwintering of birch (Betula pubescens Ehrh.). J. Exp. Bot. 2004, 55, 507–516. [Google Scholar] [CrossRef]
  38. Widrlechner, M.P.; Daly, C.; Keller, M.; Kaplan, K. Horticultural applications of a newly revised USDA Plant Hardiness Zone Map. HortTechnology 2012, 22, 6–19. [Google Scholar] [CrossRef]
  39. Gorb, S.N.; Gorb, E.V. Anti-icing strategies of plant surfaces: The ice formation on leaves visualized by Cryo-SEM experiments. Sci. Nat. 2022, 109, 24. [Google Scholar] [CrossRef]
  40. Cho, S.M.; Kim, S.H.; Cho, H.J.; Lee, H.S.; Lee, J.H.; Lee, H.; Park, H.; Kang, S.H.; Choi, H.G.; Lee, J.G. Type II ice-binding proteins isolated from an arctic microalga are similar to adhesin-like proteins and increase freezing tolerance in transgenic plants. Plant Cell Physiol. 2019, 60, 2744–2757. [Google Scholar] [CrossRef]
  41. Lyons, J.M. Chilling injury in plants. Annu. Rev. Plant Physiol. 1973, 24, 445–466. [Google Scholar] [CrossRef]
  42. Jahed, K.R.; Saini, A.K.; Sherif, S.M. Coping with the cold: Unveiling cryoprotectants, molecular signaling pathways, and strategies for cold stress resilience. Front. Plant Sci. 2023, 14, 1246093. [Google Scholar] [CrossRef] [PubMed]
  43. Aaliya, K.; Nasir, I.A.; Khan, A.; Toufiq, N.; Yousaf, I.; Adeyinka, O.S.; Iftikhar, S.; Farooq, A.M.; Tabassum, B. Expression of ice recrystallization inhibition protein in transgenic potato lines associated with reduced electrolyte leakage and efficient recovery post freezing injury. J. Biotechnol. 2021, 327, 97–105. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, M.J.; Liu, R.M.; Huang, X.H.; Du, Z.H.; Heng, S.P.; Zeng, W. Characterization of low temperature-induced plasma membrane lipidome remodeling combined with gene expression analysis reveals mechanisms that regulate membrane lipid desaturation in Carica papaya. Sci. Hortic. 2020, 272, 109505. [Google Scholar] [CrossRef]
  45. Hu, D.; Yao, Y.l.; Lv, Y.; You, J.; Zhang, Y.; Lv, Q.Y.; Li, J.W.; Hutin, S.; Xiong, H.Y.; Zubieta, C. The OsSRO1c-OsDREB2B complex undergoes protein phase transition to enhance cold tolerance in rice. Mol. Plant 2024, 17, 1520–1538. [Google Scholar] [CrossRef]
  46. Lunn, D.; Smith, G.A.; Wallis, J.G.; Browse, J. Overexpression mutants reveal a role for a chloroplast MPD protein in regulation of reactive oxygen species during chilling in Arabidopsis. J. Exp. Bot. 2022, 73, 2666–2681. [Google Scholar] [CrossRef]
  47. Gao, Y.; Thiele, W.; Saleh, O.; Scossa, F.; Arabi, F.; Zhang, H.M.; Sampathkumar, A.; Kühn, K.; Fernie, A.; Bock, R. Chloroplast translational regulation uncovers nonessential photosynthesis genes as key players in plant cold acclimation. Plant Cell 2022, 34, 2056–2079. [Google Scholar] [CrossRef]
  48. Xu, C.; Wang, Y.T.; Yang, H.D.; Tang, Y.Q.; Liu, B.C.; Hu, X.L.; Hu, Z.D. Cold acclimation alleviates photosynthetic inhibition and oxidative damage induced by cold stress in citrus seedlings. Plant Signal. Behav. 2023, 18, 2285169. [Google Scholar] [CrossRef]
  49. Ovsyannikov, A.Y.; Koteyeva, N.K. Seasonal movement of chloroplasts in mesophyll cells of two Picea species. Protoplasma 2020, 257, 183–195. [Google Scholar] [CrossRef]
  50. Herrmann, H.A.; Dyson, B.C.; Miller, M.A.E.; Schwartz, J.M.; Johnson, G.N. Metabolic flux from the chloroplast provides signals controlling photosynthetic acclimation to cold in Arabidopsis thaliana. Plant Cell Environ. 2021, 44, 171–185. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Xia, P.G. The DREB transcription factor, a biomacromolecule, responds to abiotic stress by regulating the expression of stress-related genes. Int. J. Biol. Macromol. 2023, 243, 125231. [Google Scholar] [CrossRef] [PubMed]
  52. Ahmad, M.; Li, J.Z.; Yang, Q.o.; Jamil, W.; Teng, Y.W.; Bai, S.L. Phylogenetic, molecular, and functional characterization of PpyCBF proteins in Asian pears (Pyrus pyrifolia). Int. J. Mol. Sci. 2019, 20, 2074. [Google Scholar] [CrossRef] [PubMed]
  53. Lim, C.W.; Lee, S.C. ABA-dependent and ABA-independent functions of RCAR5/PYL11 in response to cold stress. Front. Plant Sci. 2020, 11, 587620. [Google Scholar] [CrossRef]
  54. Ding, Y.L.; Jia, Y.X.; Shi, Y.T.; Zhang, X.Y.; Song, C.P.; Gong, Z.Z.; Yang, S.H. OST 1-mediated BTF 3L phosphorylation positively regulates CBF s during plant cold responses. EMBO J. 2018, 37, e98228. [Google Scholar] [CrossRef]
  55. Nai, G.J.; Liang, G.P.; Ma, W.F.; Lu, S.X.; Li, Y.M.; Gou, H.M.; Guo, L.L.; Chen, B.H.; Mao, J. Overexpression VaPYL9 improves cold tolerance in tomato by regulating key genes in hormone signaling and antioxidant enzyme. BMC Plant Biol. 2022, 22, 344. [Google Scholar] [CrossRef]
  56. Yu, M.M.; Wang, R.; Xia, J.Q.; Li, C.; Xu, Q.H.; Cang, J.; Wang, Y.Y.; Zhang, D. JA-induced TaMPK6 enhanced the freeze tolerance of Arabidopsis thaliana through regulation of ICE-CBF-COR module and antioxidant enzyme system. Plant Sci. 2023, 329, 111621. [Google Scholar] [CrossRef]
  57. Hu, Y.R.; Jiang, L.Q.; Wang, F.; Yu, D.Q. Jasmonate regulates the inducer of CBF expression–c-repeat binding factor/DRE binding factor1 cascade and freezing tolerance in Arabidopsis. Plant Cell 2013, 25, 2907–2924. [Google Scholar] [CrossRef]
  58. Fu, A.Z.; Zheng, Y.Y.; Lv, Y.H.; Watkins, C.B.; Bai, C.M.; Ma, L.; Yuan, S.Z.; Zheng, S.F.; Jia, L.E.; Gao, L.P. Multi-omics analysis reveals specific modifications associated with reduced chilling injury in bell pepper fruit by methyl jamonate. Postharvest Biol. Technol. 2022, 185, 111799. [Google Scholar] [CrossRef]
  59. Singh, A.; Singh, S.; Singh, I.K. Recent insights into the molecular mechanism of jasmonate signaling during insect-plant interaction. Australas. Plant Pathol. 2016, 45, 123–133. [Google Scholar] [CrossRef]
  60. Che, L.L.; Lu, S.X.; Gou, H.M.; Li, M.; Guo, L.L.; Yang, J.B.; Mao, J. VvJAZ13 positively regulates cold tolerance in Arabidopsis and grape. Int. J. Mol. Sci. 2024, 25, 4458. [Google Scholar] [CrossRef]
  61. Luo, R.; Wang, R.; Cao, L.; Li, L.L.; Li, X.; Yuan, Y.; Yan, J.R.; Hou, J.; Hu, J.B. Effects of plant growth regulators on physiological characteristics and related gene expression in melon seedlings under cold stress. J. Henan Agric. Univ. 2022, 56, 411–419+428. [Google Scholar] [CrossRef]
  62. Kazan, K.; Manners, J.M. MYC2: The master in action. Mol. Plant 2013, 6, 686–703. [Google Scholar] [CrossRef] [PubMed]
  63. Caccialupi, G.; Milc, J.; Caradonia, F.; Nasar, M.F.; Francia, E. The Triticeae CBF gene cluster—To frost resistance and beyond. Cells 2023, 12, 2606. [Google Scholar] [CrossRef] [PubMed]
  64. Yang, J.; Duan, G.H.; Li, C.Q.; Liu, L.; Han, G.Y.; Zhang, Y.L.; Wang, C.M. The crosstalks between jasmonic acid and other plant hormone signaling highlight the involvement of jasmonic acid as a core component in plant response to biotic and abiotic stresses. Front. Plant Sci. 2019, 10, 1349. [Google Scholar] [CrossRef]
  65. Albertos, P.; Wlk, T.; Griffiths, J.; Pimenta Lange, M.J.; Unterholzner, S.J.; Rozhon, W.; Lange, T.; Jones, A.M.; Poppenberger, B. Brassinosteroid-regulated bHLH transcription factor CESTA induces the gibberellin 2-oxidase GA2ox7. Plant Physiol. 2022, 188, 2012–2025. [Google Scholar] [CrossRef]
  66. Zhou, M.Q.; Chen, H.; Wei, D.H.; Ma, H.; Lin, J. Arabidopsis CBF3 and DELLAs positively regulate each other in response to low temperature. Sci. Rep. 2017, 7, 39819. [Google Scholar] [CrossRef]
  67. Rapacz, M.; Waligórski, P.; Janowiak, F. ABA and gibberellin-like substances during prehardening, cold acclimation, de-and reacclimation of oilseed rape. Acta Physiol. Plant. 2003, 25, 151–161. [Google Scholar] [CrossRef]
  68. Zhu, J.; Zhang, K.X.; Wang, W.S.; Gong, W.; Liu, W.C.; Chen, H.G.; Xu, H.H.; Lu, Y.T. Low temperature inhibits root growth by reducing auxin accumulation via ARR1/12. Plant Cell Physiol. 2015, 56, 727–736. [Google Scholar] [CrossRef]
  69. Du, H.; Wu, N.; Fu, J.; Wang, S.P.; Li, X.H.; Xiao, J.H.; Xiong, L.Z. A GH3 family member, OsGH3-2, modulates auxin and abscisic acid levels and differentially affects drought and cold tolerance in rice. J. Exp. Bot. 2012, 63, 6467–6480. [Google Scholar] [CrossRef]
  70. Liu, D.D.; He, Y.R.; Wang, Y.J.; Chen, W.W.; Yang, J.L.; Zhang, Y.Z.; Feng, Y.Y.; Zhao, Y.X.; Lin, S.; Huang, L. Tetrad stage transient cold stress skews auxin-mediated energy metabolism balance in Chinese cabbage pollen. Plant Physiol. 2024, 195, 1312–1332. [Google Scholar] [CrossRef]
  71. Cui, X.; Zhao, P.Y.; Liang, W.W.; Cheng, Q.; Mu, B.B.; Niu, F.F.; Yan, J.L.; Liu, C.L.; Xie, H.; Kav, N.N.V. A rapeseed WRKY transcription factor phosphorylated by CPK modulates cell death and leaf senescence by regulating the expression of ROS and SA-synthesis-related genes. J. Agric. Food Chem. 2020, 68, 7348–7359. [Google Scholar] [CrossRef] [PubMed]
  72. Saleem, M.; Fariduddin, Q.; Janda, T. Multifaceted role of salicylic acid in combating cold stress in plants: A review. J. Plant Growth Regul. 2021, 40, 464–485. [Google Scholar] [CrossRef]
  73. Taşgín, E.; Atící, Ö.; Nalbantoğlu, B. Effects of salicylic acid and cold on freezing tolerance in winter wheat leaves. Plant Growth Regul. 2003, 41, 231–236. [Google Scholar] [CrossRef]
  74. Taşgın, E.; Atıcı, Ö.; Nalbantoğlu, B.; Popova, L.P. Effects of salicylic acid and cold treatments on protein levels and on the activities of antioxidant enzymes in the apoplast of winter wheat leaves. Phytochemistry 2006, 67, 710–715. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, L.J.; Li, S.H. Salicylic acid-induced heat or cold tolerance in relation to Ca2+ homeostasis and antioxidant systems in young grape plants. Plant Sci. 2006, 170, 685–694. [Google Scholar] [CrossRef]
  76. Wu, Z.J.; Han, S.M.; Zhou, H.D.; Tuang, Z.K.; Wang, Y.Z.; Jin, Y.; Shi, H.Z.; Yang, W.N. Cold stress activates disease resistance in Arabidopsis thaliana through a salicylic acid dependent pathway. Plant Cell Environ. 2019, 42, 2645–2663. [Google Scholar] [CrossRef]
  77. Wang, Y.T.; Jiang, Q.Q.; Wang, X.F.; Xi, Z.M. Brassinosteroid stimulates hydrogen peroxide biosynthesis and reduces the effect of cold stress. J. Plant Growth Regul. 2023, 42, 3757–3769. [Google Scholar] [CrossRef]
  78. Li, H.; Ye, K.Y.; Shi, Y.T.; Cheng, J.K.; Zhang, X.Y.; Yang, S.H. BZR1 positively regulates freezing tolerance via CBF-dependent and CBF-independent pathways in Arabidopsis. Mol. Plant 2017, 10, 545–559. [Google Scholar] [CrossRef]
  79. Kagale, S.; Divi, U.K.; Krochko, J.E.; Keller, W.A.; Krishna, P. Brassinosteroid confers tolerance in Arabidopsis thaliana and Brassica napus to a range of abiotic stresses. Planta 2007, 225, 353–364. [Google Scholar] [CrossRef]
  80. Wang, L.M.; Yang, R.Z.; Sun, J.Q. Regulation of crop agronomic traits and abiotic stress responses by brassinosteroids: A review. Chin. J. Biotechnol. 2022, 38, 34–49. [Google Scholar] [CrossRef]
  81. Pociecha, E.; Dziurka, M.; Oklestkova, J.; Janeczko, A. Brassinosteroids increase winter survival of winter rye (Secale cereale L.) by affecting photosynthetic capacity and carbohydrate metabolism during the cold acclimation process. Plant Growth Regul. 2016, 80, 127–135. [Google Scholar] [CrossRef]
  82. Arfan, M.; Zhang, D.W.; Zou, L.J.; Luo, S.S.; Tan, W.R.; Zhu, T.; Lin, H.H. Hydrogen peroxide and nitric oxide crosstalk mediates brassinosteroids induced cold stress tolerance in Medicago truncatula. Int. J. Mol. Sci. 2019, 20, 144. [Google Scholar] [CrossRef] [PubMed]
  83. Robison, J.D.; Yamasaki, Y.; Randall, S.K. The ethylene signaling pathway negatively impacts CBF/DREB-regulated cold response in soybean (Glycine max). Front. Plant Sci. 2019, 10, 121. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, J.; Yin, X.R.; Li, H.; Xu, M.; Zhang, M.X.; Li, S.J.; Liu, X.F.; Shi, Y.N.; Grierson, D.; Chen, K.S. ETHYLENE RESPONSE FACTOR39–MYB8 complex regulates low-temperature-induced lignification of loquat fruit. J. Exp. Bot. 2020, 71, 3172–3184. [Google Scholar] [CrossRef]
  85. Zhao, D.Y.; Shen, L.; Fan, B.; Yu, M.M.; Zheng, Y.; Lv, S.N.; Sheng, J.P. Ethylene and cold participate in the regulation of LeCBF1 gene expression in postharvest tomato fruits. FEBS Lett. 2009, 583, 3329–3334. [Google Scholar] [CrossRef]
  86. Zhao, M.G.; Liu, W.J.; Xia, X.Z.; Wang, T.Z.; Zhang, W.H. Cold acclimation-induced freezing tolerance of Medicago truncatula seedlings is negatively regulated by ethylene. Physiol. Plant. 2014, 152, 115–129. [Google Scholar] [CrossRef]
  87. Wang, Y.C.; Jiang, H.Y.; Mao, Z.L.; Liu, W.J.; Jiang, S.H.; Xu, H.F.; Su, M.Y.; Zhang, J.; Wang, N.; Zhang, Z.Y. Ethylene increases the cold tolerance of apple via the MdERF1B–MdCIbHLH1 regulatory module. Plant J. 2021, 106, 379–393. [Google Scholar] [CrossRef]
  88. Zhang, X.Z.; Ervin, E.H.; Waltz, C.; Murphy, T. Metabolic changes during cold acclimation and deacclimation in five bermudagrass varieties: II. cytokinin and abscisic acid metabolism. Crop Sci. 2011, 51, 847–853. [Google Scholar] [CrossRef]
  89. Chun, J.; Wan, M.; Guo, H.W.; Zhang, Q.P.; Feng, Y.; Tang, Y.C.; Zhu, B.; Sang, Y.S.; Jing, S.L.; Chen, T. Cytokinin-mediated enhancement of potato growth and yield by Verticillium Dahliae effector VDAL under low temperature stress. BMC Plant Biol. 2024, 24, 1115. [Google Scholar] [CrossRef]
  90. Jeon, J.; Kim, N.Y.; Kim, S.; Kang, N.Y.; Novák, O.; Ku, S.J.; Cho, C.; Lee, D.J.; Lee, E.J.; Strnad, M. A subset of cytokinin two-component signaling system plays a role in cold temperature stress response in Arabidopsis. J. Biol. Chem. 2010, 285, 23371–23386. [Google Scholar] [CrossRef]
  91. Shi, Y.T.; Tian, S.W.; Hou, L.Y.; Huang, X.Z.; Zhang, X.Y.; Guo, H.W.; Yang, S.H. Ethylene signaling negatively regulates freezing tolerance by repressing expression of CBF and type-A ARR genes in Arabidopsis. Plant Cell 2012, 24, 2578–2595. [Google Scholar] [CrossRef] [PubMed]
  92. Chang, T.L.; Xi, Q.Q.; Wei, X.Y.; Xu, L.; Wang, Q.Q.; Fu, J.Y.; Ling, C.; Zuo, Y.; Zhao, Y.; He, H.Y. Rhythmical redox homeostasis can be restored by exogenous melatonin in hulless barley (Hordeum vulgare L. var. nudum) under cold stress. Environ. Exp. Bot. 2022, 194, 104756. [Google Scholar] [CrossRef]
  93. Zhao, H.L.; Ye, L.; Wang, Y.P.; Zhou, X.T.; Yang, J.W.; Wang, J.W.; Cao, K.; Zou, Z.R. Melatonin increases the chilling tolerance of chloroplast in cucumber seedlings by regulating photosynthetic electron flux and the ascorbate-glutathione cycle. Front. Plant Sci. 2016, 7, 1814. [Google Scholar] [CrossRef] [PubMed]
  94. Bajwa, V.S.; Shukla, M.R.; Sherif, S.M.; Murch, S.J.; Saxena, P.K. Role of melatonin in alleviating cold stress in Arabidopsis thaliana. J. Pineal Res. 2014, 56, 238–245. [Google Scholar] [CrossRef]
  95. Zhao, H.L.; Zhang, K.; Zhou, X.T.; Xi, L.J.; Wang, Y.P.; Xu, H.J.; Pan, T.H.; Zou, Z.R. Melatonin alleviates chilling stress in cucumber seedlings by up-regulation of CsZat12 and modulation of polyamine and abscisic acid metabolism. Sci. Rep. 2017, 7, 4998. [Google Scholar] [CrossRef]
  96. Maity, S.; Guchhait, R.; Pramanick, K. Melatonin mediated activation of MAP kinase pathway may reduce DNA damage stress in plants: A review. BioFactors 2022, 48, 965–971. [Google Scholar] [CrossRef]
  97. Li, X.N.; Brestic, M.; Tan, D.X.; Zivcak, M.; Zhu, X.C.; Liu, S.Q.; Song, F.B.; Reiter, R.J.; Liu, F.L. Melatonin alleviates low PS I-limited carbon assimilation under elevated CO 2 and enhances the cold tolerance of offspring in chlorophyll b-deficient mutant wheat. J. Pineal Res. 2018, 64, e12453. [Google Scholar] [CrossRef]
  98. Wang, S.M.; Wang, Y.S.; Su, B.Y.; Zhou, Y.Y.; Chang, L.F.; Ma, X.Y.; Li, X.M. Ecophysiological responses of five mangrove species (Bruguiera gymnorrhiza, Rhizophora stylosa, Aegiceras corniculatum, Avicennia marina, and Kandelia obovata) to chilling stress. Front. Mar. Sci. 2022, 9, 846566. [Google Scholar] [CrossRef]
  99. Buer, J.V.; Cvetkovic, J.; Baier, M. Cold regulation of plastid ascorbate peroxidases serves as a priming hub controlling ROS signaling in Arabidopsis thaliana. BMC Plant Biol. 2016, 16, 1–20. [Google Scholar] [CrossRef]
  100. Shi, H.T.; Ye, T.T.; Zhong, B.; Liu, X.; Jin, R.; Chan, Z.L. At HAP 5A modulates freezing stress resistance in Arabidopsis through binding to CCAAT motif of AtXTH21. New Phytol. 2014, 203, 554–567. [Google Scholar] [CrossRef]
  101. Vogel, J.T.; Zarka, D.G.; The Plant JournalBuskirk, H.A.V.; Fowler, S.G.; Thomashow, M.F. Roles of the CBF2 and ZAT12 transcription factors in configuring the low temperature transcriptome of Arabidopsis. Plant J. 2005, 41, 195–211. [Google Scholar] [CrossRef] [PubMed]
  102. Han, Y.C.; Fu, C.C. Cold-inducible MaC2H2s are associated with cold stress response of banana fruit via regulating MaICE1. Plant Cell Rep. 2019, 38, 673–680. [Google Scholar] [CrossRef] [PubMed]
  103. Liu, Y.; Xu, C.J.; Zhu, Y.F.; Zhang, L.N.; Chen, T.Y.; Zhou, F.; Chen, H.; Lin, Y.J. The calcium-dependent kinase OsCPK24 functions in cold stress responses in rice. J. Integr. Plant Biol. 2018, 60, 173–188. [Google Scholar] [CrossRef]
  104. Chen, Y.; Chen, L.; Sun, X.M.; Kou, S.; Liu, T.T.; Dong, J.K.; Tu, W.; Zhang, Y.L.; Song, B.T. The mitogen-activated protein kinase kinase MKK2 positively regulates constitutive cold resistance in the potato. Environ. Exp. Bot. 2022, 194, 104702. [Google Scholar] [CrossRef]
  105. Zhang, M.; Shi, H.T.; Li, N.N.; Wei, N.N.; Tian, Y.; Peng, J.F.; Chen, X.C.; Zhang, L.Y.; Zhang, M.X.; Dong, H.S. Aquaporin OsPIP2; 2 links the H2O2 signal and a membrane-anchored transcription factor to promote plant defense. Plant Physiol. 2022, 188, 2325–2341. [Google Scholar] [CrossRef]
  106. Teige, M.; Scheikl, E.; Eulgem, T.; Doczi, R.; Ichimura, K.; Shinozaki, K.; Dangl, J.L.; Hirt, H. The MKK2 pathway mediates cold and salt stress signaling in Arabidopsis. Mol. Cell 2004, 15, 141–152. [Google Scholar] [CrossRef]
  107. Ding, Y.L.; Yang, S.H. Surviving and thriving: How plants perceive and respond to temperature stress. Dev. Cell 2022, 57, 947–958. [Google Scholar] [CrossRef]
  108. Geng, B.H.; Wang, Q.; Huang, R.S.; Liu, Y.J.; Guo, Z.F.; Lu, S.Y. A novel LRR-RLK (CTLK) confers cold tolerance through regulation on the C-repeat-binding factor pathway, antioxidants, and proline accumulation. Plant J. 2021, 108, 1679–1689. [Google Scholar] [CrossRef]
  109. Constable, G.A. Temperature effects on the early field development of cotton. Aust. J. Exp. Agric. 1976, 16, 905–910. [Google Scholar] [CrossRef]
  110. Cui, Y.M.; Lu, S.; Li, Z.; Cheng, J.W.; Hu, P.; Zhu, T.Q.; Wang, X.; Jin, M.; Wang, X.X.; Li, L.Q. 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]
  111. Wang, J.C.; Ren, Y.L.; Liu, X.; Luo, S.; Zhang, X.; Liu, X.; Lin, Q.B.; Zhu, S.S.; Wan, H.; Yang, Y. Transcriptional activation and phosphorylation of OsCNGC9 confer enhanced chilling tolerance in rice. Mol. Plant 2021, 14, 315–329. [Google Scholar] [CrossRef] [PubMed]
  112. Barnham, E.J.; Wang, L.M.; Ning, Y.Z.; Davies, J.M. The complex story of plant cyclic nucleotide-gated channels. Int. J. Mol. Sci. 2021, 22, 874. [Google Scholar] [CrossRef]
  113. Zhang, Z.L.; Hou, C.C.; Tian, W.; Li, L.G.; Zhu, H.F. Electrophysiological studies revealed CaM1-mediated regulation of the Arabidopsis calcium channel CNGC12. Front. Plant Sci. 2019, 10, 1090. [Google Scholar] [CrossRef] [PubMed]
  114. Huang, C.L.; Ding, S.; Zhang, H.; Du, H.; An, L.Z. CIPK7 is involved in cold response by interacting with CBL1 in Arabidopsis thaliana. Plant Sci. 2011, 181, 57–64. [Google Scholar] [CrossRef] [PubMed]
  115. Song, S.J.; Feng, Q.N.; Li, C.L.; Li, E.; Liu, Q.; Kang, H.; Zhang, W.; Zhang, Y.; Li, S. A tonoplast-associated calcium-signaling module dampens ABA signaling during stomatal movement. Plant Physiol. 2018, 177, 1666–1678. [Google Scholar] [CrossRef]
  116. Hwang, Y.S.; Bethke, P.C.; Cheong, Y.H.; Chang, H.S.; Zhu, T.; Jones, R.L. A gibberellin-regulated calcineurin B in rice localizes to the tonoplast and is implicated in vacuole function. Plant Physiol. 2005, 138, 1347–1358. [Google Scholar] [CrossRef]
  117. Washington, E.J.; Mukhtar, M.S.; Finkel, O.M.; Wan, L.; Banfield, M.J.; Kieber, J.J.; Dangl, J.L. Pseudomonas syringae type III effector HopAF1 suppresses plant immunity by targeting methionine recycling to block ethylene induction. Proc. Natl. Acad. Sci. USA 2016, 113, E3577–E3586. [Google Scholar] [CrossRef]
  118. Xiao, C.; Zhang, H.; Xie, F.; Pan, Z.Y.; Qiu, W.M.; Tong, Z.; Wang, Z.Q.; He, X.J.; Xu, Y.H.; Sun, Z.H. Evolution, gene expression, and protein–protein interaction analyses identify candidate CBL-CIPK signalling networks implicated in stress responses to cold and bacterial infection in citrus. BMC Plant Biol. 2022, 22, 420. [Google Scholar] [CrossRef]
  119. Cai, J.; Chang, T.L.; Zhao, Y.; He, H.Y.; Li, J.X.; Xi, Q.Q.; Fu, J.Y.; Zhao, Y.W. Effects of varying temperature on rhythmic expression of abiotic stress-responding genes in Tibetan hulless barley. Acta Physiol. Plant. 2022, 44, 38. [Google Scholar] [CrossRef]
  120. Mohanta, T.K.; Mohanta, N.; Mohanta, Y.K.; Bae, H.H. Genome-wide identification of calcium dependent protein kinase gene family in plant lineage shows presence of novel DxD and DEL motifs in EF-hand domain. Front. Plant Sci. 2015, 6, 1146. [Google Scholar] [CrossRef]
  121. Hetherington, A.; Trewavas, A. Calcium-dependent protein kinase in pea shoot membranes. FEBS Lett. 1982, 145, 67–71. [Google Scholar] [CrossRef]
  122. Boudsocq, M.; Droillard, M.J.; Regad, L.; Laurière, C. Characterization of Arabidopsis calcium-dependent protein kinases: Activated or not by calcium? Biochem. J. 2012, 447, 291–299. [Google Scholar] [CrossRef] [PubMed]
  123. Monroy, A.F.; Dhindsa, R.S. Low-temperature signal transduction: Induction of cold acclimation-specific genes of alfalfa by calcium at 25 degrees C. Plant Cell 1995, 7, 321–331. [Google Scholar] [CrossRef] [PubMed]
  124. Weckwerth, P.; Ehlert, B.; Romeis, T. ZmCPK1, a calcium-independent kinase member of the Zea mays CDPK gene family, functions as a negative regulator in cold stress signalling. Plant Cell Environ. 2015, 38, 544–558. [Google Scholar] [CrossRef]
  125. Szczegielniak, J.; Borkiewicz, L.; Szurmak, B.; Gnatowska, E.L.; Statkiewicz, M.; Klimecka, M.; Cieśla, J.; Muszyńska, G. Maize calcium-dependent protein kinase (ZmCPK11): Local and systemic response to wounding, regulation by touch and components of jasmonate signaling. Physiol. Plant. 2012, 146, 1–14. [Google Scholar] [CrossRef]
  126. Ding, Y.F.; Cao, J.M.; Ni, L.; Zhu, Y.; Zhang, A.Y.; Tan, M.P.; Jiang, M.Y. ZmCPK11 is involved in abscisic acid-induced antioxidant defence and functions upstream of ZmMPK5 in abscisic acid signalling in maize. J. Exp. Bot. 2013, 64, 871–884. [Google Scholar] [CrossRef]
  127. Dubrovina, A.S.; Kiselev, K.V.; Khristenko, V.S.; Aleynova, O.A. VaCPK21, a calcium-dependent protein kinase gene of wild grapevine Vitis amurensis Rupr., is involved in grape response to salt stress. J. Plant Physiol. 2016, 124, 137–150. [Google Scholar] [CrossRef]
  128. Xu, X.J.; Liang, C.Z.; Liu, H.Y. Joint multi-generations genetic analysis on chilling tolerance of melon (Cucumis melo) seedlings. J. Agric. Biotechnol. 2020, 28, 420–428. [Google Scholar] [CrossRef]
  129. Xie, Y.P.; Chen, P.X.; Yan, Y.; Bao, C.N.; Li, X.W.; Wang, L.P.; Shen, X.X.; Li, H.Y.; Liu, X.F.; Niu, C.D. An atypical R2R3 MYB transcription factor increases cold hardiness by CBF-dependent and CBF-independent pathways in apple. New Phytol. 2018, 218, 201–218. [Google Scholar] [CrossRef]
  130. Abdullah, S.N.A.; Azzeme, A.M.; Yousefi, K. Fine-tuning cold stress response through regulated cellular abundance and mechanistic actions of transcription factors. Front. Plant Sci. 2022, 13, 850216. [Google Scholar] [CrossRef]
  131. Hayashi, Y.; Takahashi, Y.; Fukatsu, K.; Tada, Y.; Takahashi, K.; Kuwata, K.; Suzuki, T.; Kinoshita, T. Identification of abscisic acid-dependent phosphorylated basic helix-Loop-Helix transcription factors in guard cells of Vicia faba by mass spectrometry. Front. Plant Sci. 2021, 12, 735271. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, W.D.; Gao, T.; Chen, J.F.; Yang, J.K.; Huang, H.Y.; Yu, Y.B. The late embryogenesis abundant gene family in tea plant (Camellia sinensis): Genome-wide characterization and expression analysis in response to cold and dehydration stress. Plant Physiol. Biochem. 2019, 135, 277–286. [Google Scholar] [CrossRef] [PubMed]
  133. Lv, K.W.; Li, J.; Zhao, K.; Chen, S.; Nie, J.; Zhang, W.L.; Liu, G.F.; Wei, H.R. Overexpression of an AP2/ERF family gene, BpERF13, in birch enhances cold tolerance through upregulating CBF genes and mitigating reactive oxygen species. Plant Sci. 2020, 292, 110375. [Google Scholar] [CrossRef]
  134. Hwarari, D.; Guan, Y.L.; Ahmad, B.; Movahedi, A.; Min, T.; Hao, Z.D.; Lu, Y.; Chen, J.H.; Yang, L.M. ICE-CBF-COR signaling cascade and its regulation in plants responding to cold stress. Int. J. Mol. Sci. 2022, 23, 1549. [Google Scholar] [CrossRef]
  135. Song, Y.; Zhang, X.Y.; Li, M.Z.; Yang, H.; Fu, D.Y.; Lv, J.; Ding, Y.L.; Gong, Z.Z.; Shi, Y.T.; Yang, S.H. The direct targets of CBFs: In cold stress response and beyond. J. Integr. Plant Biol. 2021, 63, 1874–1887. [Google Scholar] [CrossRef]
  136. Shan, X.Z.; Yang, Y.; Wei, S.Q.; Wang, C.; Shen, W.S.; Chen, H.B.; Shen, J.Y. Involvement of CBF in the fine-tuning of litchi flowering time and cold and drought stresses. Front. Plant Sci. 2023, 14, 1167458. [Google Scholar] [CrossRef]
  137. Novillo, F.; Medina, J.; Salinas, J. Arabidopsis CBF1 and CBF3 have a different function than CBF2 in cold acclimation and define different gene classes in the CBF regulon. Proc. Natl. Acad. Sci. USA 2007, 104, 21002–21007. [Google Scholar] [CrossRef]
  138. Novillo, F.; Alonso, J.M.; Ecker, J.R.; Salinas, J. CBF2/DREB1C is a negative regulator of CBF1/DREB1B and CBF3/DREB1A expression and plays a central role in stress tolerance in Arabidopsis. Proc. Natl. Acad. Sci. USA 2004, 101, 3985–3990. [Google Scholar] [CrossRef]
  139. Kim, S.H.; Kim, H.S.; Bahk, S.; An, J.; Yoo, Y.; Kim, J.Y.; Chung, W.S. Phosphorylation of the transcriptional repressor MYB15 by mitogen-activated protein kinase 6 is required for freezing tolerance in Arabidopsis. Nucleic Acids Res. 2017, 45, 6613–6627. [Google Scholar] [CrossRef]
  140. Wang, X.; Ding, Y.L.; Li, Z.Y.; Shi, Y.T.; Wang, J.L.; Hua, J.; Gong, Z.Z.; Zhou, J.M.; Yang, S.H. PUB25 and PUB26 promote plant freezing tolerance by degrading the cold signaling negative regulator MYB15. Dev. Cell 2019, 51, 222–235. [Google Scholar] [CrossRef]
  141. Dou, T.X.; Hu, C.H.; Sun, X.X.; Shao, X.H.; Wu, J.H.; Ding, L.J.; Gao, J.; He, W.D.; Biswas, M.K.; Yang, Q.s. MpMYBS3 as a crucial transcription factor of cold signaling confers the cold tolerance of banana. Plant Cell Tissue Organ. Cult. 2016, 125, 93–106. [Google Scholar] [CrossRef]
  142. Su, C.F.; Wang, Y.C.; Hsieh, T.H.; Lu, C.A.; Tseng, T.H.; Yu, S.M. A novel MYBS3-dependent pathway confers cold tolerance in rice. Plant Physiol. 2010, 153, 145–158. [Google Scholar] [CrossRef] [PubMed]
  143. Hu, M.Y.; Yang, Y.; Guo, Y.D.; Zhang, X.C. Functional analysis of SlMYB96 gene in tomato under cold stress. Biotechnol. Bull. 2023, 39, 236. [Google Scholar] [CrossRef]
  144. Sanghera, G.S.; Wani, S.H.; Hussain, W.; Singh, N.B. Engineering cold stress tolerance in crop plants. Curr. Genom. 2011, 12, 30–43. [Google Scholar] [CrossRef] [PubMed]
  145. Zhang, L.Y.; Jiang, X.C.; Liu, Q.Y.; Ahammed, G.J.; Lin, R.; Wang, L.Y.; Shao, S.J.; Yu, J.Q.; Zhou, Y.H. The HY5 and MYB15 transcription factors positively regulate cold tolerance in tomato via the CBF pathway. Plant Cell Environ. 2020, 43, 2712–2726. [Google Scholar] [CrossRef] [PubMed]
  146. Yu, H.H.; Teng, Z.N.; Liu, B.H.; Lv, J.H.; Chen, Y.K.; Qin, Z.E.; Peng, Y.; Meng, S.; He, Y.C.; Duan, M.J. Transcription factor OsMYB30 increases trehalose content to inhibit α-amylase and seed germination at low temperature. Plant Physiol. 2024, 194, 1815–1833. [Google Scholar] [CrossRef]
  147. Ma, X.; Yu, Y.N.; Jia, J.H.; Li, Q.H.; Gong, Z.H. The pepper MYB transcription factor CaMYB306 accelerates fruit coloration and negatively regulates cold resistance. Sci. Hortic. 2022, 295, 110892. [Google Scholar] [CrossRef]
  148. Ishiguro, S.; Nakamura, K. Characterization of a cDNA encoding a novel DNA-binding protein, SPF1, that recognizes SP8 sequences in the 5′ upstream regions of genes coding for sporamin and β-amylase from sweet potato. Mol. Gen. Genet. 1994, 244, 563–571. [Google Scholar] [CrossRef]
  149. Zhang, J.N.; Zhao, H.Q.; Chen, L.; Lin, J.C.; Wang, Z.L.; Pan, J.Q.; Yang, F.; Ni, X.L.; Wang, Y.A.; Wang, Y.H. Multifaceted roles of WRKY transcription factors in abiotic stress and flavonoid biosynthesis. Front. Plant Sci. 2023, 14, 1303667. [Google Scholar] [CrossRef]
  150. Zou, C.S.; Jiang, W.B.; Yu, D.Q. Male gametophyte-specific WRKY34 transcription factor mediates cold sensitivity of mature pollen in Arabidopsis. J. Exp. Bot. 2010, 61, 3901–3914. [Google Scholar] [CrossRef]
  151. Stock, J.; Bräutigam, A.; Melzer, M.; Bienert, G.P.; Bunk, B.; Nagel, M.; Overmann, J.; Keller, E.R.J.; Mock, H.P. The transcription factor WRKY22 is required during cryo-stress acclimation in Arabidopsis shoot tips. J. Exp. Bot. 2020, 71, 4993–5009. [Google Scholar] [CrossRef] [PubMed]
  152. Sarkar, S.; Das, A.; Khandagale, P.; Maiti, I.B.; Chattopadhyay, S.; Dey, N. Interaction of Arabidopsis TGA3 and WRKY53 transcription factors on Cestrum yellow leaf curling virus (CmYLCV) promoter mediates salicylic acid-dependent gene expression in planta. Planta 2018, 247, 181–199. [Google Scholar] [CrossRef] [PubMed]
  153. Negi, N.; Khurana, P. A salicylic acid inducible mulberry WRKY transcription factor, MiWRKY53 is involved in plant defence response. Plant Cell Rep. 2021, 40, 2151–2171. [Google Scholar] [CrossRef] [PubMed]
  154. Li, J.; Brader, G.; Palva, E.T. The WRKY70 transcription factor: A node of convergence for jasmonate-mediated and salicylate-mediated signals in plant defense. Plant Cell 2004, 16, 319–331. [Google Scholar] [CrossRef]
  155. Boro, P.; Sultana, A.; Mandal, K.; Chattopadhyay, S. Interplay between glutathione and mitogen-activated protein kinase 3 via transcription factor WRKY40 under combined osmotic and cold stress in Arabidopsis. J. Plant Physiol. 2022, 271, 153664. [Google Scholar] [CrossRef]
  156. Wang, L.H.; Chen, H.; Chen, G.Y.; Luo, G.B.; Shen, X.Y.; Ouyang, B.; Bie, Z.L. Transcription factor SlWRKY50 enhances cold tolerance in tomato by activating the jasmonic acid signaling. Plant Physiol. 2024, 194, 1075–1090. [Google Scholar] [CrossRef]
  157. Huang, X.B.; Cao, L.W.; Fan, J.B.; Ma, G.J.; Chen, L. CdWRKY2-mediated sucrose biosynthesis and CBF-signalling pathways coordinately contribute to cold tolerance in bermudagrass. Plant Biotechnol. J. 2022, 20, 660–675. [Google Scholar] [CrossRef]
  158. Zhang, Y.; Yu, H.J.; Yang, X.Y.; Li, Q.; Ling, J.; Wang, H.; Gu, X.F.; Huang, S.W.; Jiang, W.J. CsWRKY46, a WRKY transcription factor from cucumber, confers cold resistance in transgenic-plant by regulating a set of cold-stress responsive genes in an ABA-dependent manner. Plant Physiol. Biochem. 2016, 108, 478–487. [Google Scholar] [CrossRef]
  159. Foster, R.; Izawa, T.; Chua, N.H. Plant bZIP proteins gather at ACGT elements. FASEB J. 1994, 8, 192–200. [Google Scholar] [CrossRef]
  160. Wang, X.J.; Guo, C.; Peng, J.; Li, C.; Wan, F.F.; Zhang, S.M.; Zhou, Y.Y.; Yan, Y.; Qi, L.J.; Sun, K.W. ABRE-BINDING FACTORS play a role in the feedback regulation of ABA signaling by mediating rapid ABA induction of ABA co-receptor genes. New Phytol. 2019, 221, 341–355. [Google Scholar] [CrossRef]
  161. Yue, X.L.; Zhang, G.Y.; Zhao, Z.; Yue, J.L.; Pu, X.H.; Sui, M.J.; Zhan, Y.; Shi, Y.L.; Wang, Z.Y.; Meng, G.H. A cryophyte transcription factor, CbABF1, confers freezing, and drought tolerance in tobacco. Front. Plant Sci. 2019, 10, 699. [Google Scholar] [CrossRef] [PubMed]
  162. Liu, X.Y.; Bulley, S.M.; Gasic, E.V.; Zhong, C.H.; Li, D.W. Kiwifruit bZIP transcription factor AcePosF21 elicits ascorbic acid biosynthesis during cold stress. Plant Physiol. 2023, 192, 982–999. [Google Scholar] [CrossRef] [PubMed]
  163. Li, Z.Y.; Fu, D.Y.; Wang, X.; Zeng, R.; Zhang, X.; Tian, J.; Zhang, S.S.; Yang, X.H.; Tian, F.; Lai, J.S. The transcription factor bZIP68 negatively regulates cold tolerance in maize. Plant Cell 2022, 34, 2833–2851. [Google Scholar] [CrossRef] [PubMed]
  164. Yao, L.N.; Hao, X.Y.; Cao, H.L.; Ding, C.Q.; Yang, Y.J.; Wang, L.; Wang, X.C. ABA-dependent bZIP transcription factor, CsbZIP18, from Camellia sinensis negatively regulates freezing tolerance in Arabidopsis. Plant Cell Rep. 2020, 39, 553–565. [Google Scholar] [CrossRef]
  165. Shimizu, H.; Sato, K.; Berberich, T.; Miyazaki, A.; Ozaki, R.; Imai, R.; Kusano, T. LIP19, a basic region leucine zipper protein, is a Fos-like molecular switch in the cold signaling of rice plants. Plant Cell Physiol. 2005, 46, 1623–1634. [Google Scholar] [CrossRef]
  166. Liao, Y.; Zou, H.F.; Wei, W.; Hao, Y.J.; Tian, A.G.; Huang, J.; Liu, Y.F.; Zhang, J.S.; Chen, S.Y. Soybean GmbZIP44, GmbZIP62 and GmbZIP78 genes function as negative regulator of ABA signaling and confer salt and freezing tolerance in transgenic Arabidopsis. Planta 2008, 228, 225–240. [Google Scholar] [CrossRef]
  167. Wang, F.; Wu, N.; Zhang, L.Y.; Ahammed, G.J.; Chen, X.X.; Xiang, X.; Zhou, J.; Xia, X.J.; Shi, K.; Yu, J.Q. Light signaling-dependent regulation of photoinhibition and photoprotection in tomato. Plant Physiol. 2018, 176, 1311–1326. [Google Scholar] [CrossRef]
  168. Li, Y.P.; Shi, Y.T.; Li, M.Z.; Fu, D.Y.; Wu, S.F.; Li, J.G.; Gong, Z.Z.; Liu, H.T.; Yang, S.H. The CRY2–COP1–HY5–BBX7/8 module regulates blue light-dependent cold acclimation in Arabidopsis. Plant Cell 2021, 33, 3555–3573. [Google Scholar] [CrossRef]
  169. Su, H.Y.; Zhang, S.Z.; Yuan, X.W.; Chen, C.T.; Wang, X.F.; Hao, Y.J. Genome-wide analysis and identification of stress-responsive genes of the NAM–ATAF1, 2–CUC2 transcription factor family in apple. Plant Physiol. Biochem. 2013, 71, 11–21. [Google Scholar] [CrossRef]
  170. Pooam, M.; El-Ballat, E.M.; Jourdan, N.; Ali, H.M.; Hano, C.; Ahmad, M.; El-Esawi, M.A. SNAC3 transcription factor enhances arsenic stress tolerance and grain yield in rice (Oryza sativa L.) through regulating physio-biochemical mechanisms, stress-responsive genes, and cryptochrome 1b. Plants 2023, 12, 2731. [Google Scholar] [CrossRef]
  171. Matschi, S.; Werner, S.; Schulze, W.X.; Legen, J.; Hilger, H.H.; Romeis, T. Function of calcium-dependent protein kinase CPK 28 of Arabidopsis thaliana in plant stem elongation and vascular development. Plant J. 2013, 73, 883–896. [Google Scholar] [CrossRef] [PubMed]
  172. Song, C.B.; Wu, M.B.; Zhou, Y.; Gong, Z.H.; Yu, W.W.; Zhang, Y.; Yang, Z.F. NAC-mediated membrane lipid remodeling negatively regulates fruit cold tolerance. Hortic. Res. 2022, 9, uhac039. [Google Scholar] [CrossRef] [PubMed]
  173. Kong, X.M.; Zhou, Q.; Zhou, X.; Wei, B.D.; Ji, S.J. Transcription factor CaNAC1 regulates low-temperature-induced phospholipid degradation in green bell pepper. J. Exp. Bot. 2020, 71, 1078–1091. [Google Scholar] [CrossRef] [PubMed]
  174. Hou, X.M.; Zhang, H.F.; Liu, S.Y.; Wang, X.K.; Zhang, Y.M.; Meng, Y.C.; Luo, D.; Chen, R.G. The NAC transcription factor CaNAC064 is a regulator of cold stress tolerance in peppers. Plant Sci. 2020, 291, 110346. [Google Scholar] [CrossRef]
  175. Han, D.G.; Du, M.; Zhou, Z.Y.; Wang, S.; Li, T.M.; Han, J.X.; Xu, T.L.; Yang, G.H. Overexpression of a Malus baccata NAC transcription factor gene MbNAC25 increases cold and salinity tolerance in Arabidopsis. Int. J. Mol. Sci. 2020, 21, 1198. [Google Scholar] [CrossRef]
  176. Hu, X.L.; Xie, F.F.; Liang, W.W.; Liang, Y.H.; Zhang, Z.K.; Zhao, J.T.; Hu, G.B.; Qin, Y.H. HuNAC20 and HuNAC25, two novel NAC genes from pitaya, confer cold tolerance in transgenic Arabidopsis. Int. J. Mol. Sci. 2022, 23, 2189. [Google Scholar] [CrossRef]
  177. Sun, X.; Wang, Y.; Sui, N. Transcriptional regulation of bHLH during plant response to stress. Biochem. Biophys. Res. Commun. 2018, 503, 397–401. [Google Scholar] [CrossRef]
  178. Chen, W.J.; Wang, X.; Yan, S.; Huang, X.; Yuan, H.M. The ICE-like transcription factor HbICE2 is involved in jasmonate-regulated cold tolerance in the rubber tree (Hevea brasiliensis). Plant Cell Rep. 2019, 38, 699–714. [Google Scholar] [CrossRef]
  179. Ritonga, F.N.; Chen, S. Physiological and molecular mechanism involved in cold stress tolerance in plants. Plants 2020, 9, 560. [Google Scholar] [CrossRef]
  180. Wu, C.L.; Lin, L.F.; Hsu, H.C.; Huang, L.F.; Hsiao, C.D.; Chou, M.L. Saussurea involucrata (Snow Lotus) ICE1 and ICE2 orthologues involved in regulating cold stress tolerance in transgenic Arabidopsis. Int. J. Mol. Sci. 2021, 22, 10850. [Google Scholar] [CrossRef]
  181. Zuo, Z.F.; Kang, H.G.; Hong, Q.C.; Park, M.Y.; Sun, H.J.; Kim, J.; Song, P.S.; Lee, H.Y. A novel basic helix-loop-helix transcription factor, ZjICE2 from Zoysia japonica confers abiotic stress tolerance to transgenic plants via activating the DREB/CBF regulon and enhancing ROS scavenging. Plant Mol. Biol. 2020, 102, 447–462. [Google Scholar] [CrossRef] [PubMed]
  182. Pan, R.; Buitrago, S.; Peng, Y.; Elwafa, S.F.A.; Wan, K.; Liu, Y.; Wang, R.; Yang, X.S.; Zhang, W.Y. Genome-wide identification of cold-tolerance genes and functional analysis of IbbHLH116 gene in sweet potato. Gene 2022, 837, 146690. [Google Scholar] [CrossRef] [PubMed]
  183. Yang, J.; Guo, X.; Mei, Q.L.; Qiu, L.; Chen, P.H.; Li, W.H.; Mao, K.; Ma, F.W. MdbHLH4 negatively regulates apple cold tolerance by inhibiting MdCBF1/3 expression and promoting MdCAX3L-2 expression. Plant Physiol. 2023, 191, 789–806. [Google Scholar] [CrossRef] [PubMed]
  184. Ni, M.; Tepperman, J.M.; Quail, P.H. PIF3, a phytochrome-interacting factor necessary for normal photoinduced signal transduction, is a novel basic helix-loop-helix protein. Cell 1998, 95, 657–667. [Google Scholar] [CrossRef]
  185. Wang, F.; Chen, X.X.; Dong, S.J.; Jiang, X.C.; Wang, L.Y.; Yu, J.Q.; Zhou, Y.H. Crosstalk of PIF4 and DELLA modulates CBF transcript and hormone homeostasis in cold response in tomato. Plant Biotechnol. J. 2020, 18, 1041–1055. [Google Scholar] [CrossRef]
  186. Ming, R.H.; Zhang, Y.; Wang, Y.; Khan, M.; Dahro, B.; Liu, J.H. The JA-responsive MYC2-BADH-like transcriptional regulatory module in Poncirus trifoliata contributes to cold tolerance by modulation of glycine betaine biosynthesis. New Phytol. 2021, 229, 2730–2750. [Google Scholar] [CrossRef]
  187. Min, D.D.; Li, F.J.; Zhang, X.H.; Cui, X.X.; Shu, P.; Dong, L.L.; Ren, C.T. SlMYC2 involved in methyl jasmonate-induced tomato fruit chilling tolerance. J. Agric. Food Chem. 2018, 66, 3110–3117. [Google Scholar] [CrossRef]
  188. Lee, H.G.; Seo, P.J. The MYB 96–HHP module integrates cold and abscisic acid signaling to activate the CBF–COR pathway in Arabidopsis. Plant J. 2015, 82, 962–977. [Google Scholar] [CrossRef]
  189. Lv, Y.; Yang, M.; Hu, D.; Yang, Z.Y.; Ma, S.Q.; Li, X.H.; Xiong, L.Z. The OsMYB30 transcription factor suppresses cold tolerance by interacting with a JAZ protein and suppressing β-amylase expression. Plant Physiol. 2017, 173, 1475–1491. [Google Scholar] [CrossRef]
  190. Zhang, Z.Y.; Cheng, L.; Zhang, W.H.; Hu, J.L.; Liu, Y.; Lin, Y.Z. Characterization of transcription activation domain of EcaICE1 and its interaction with EcaSIZ1 in Eucalyptus camaldulensis. Trees 2020, 34, 1243–1253. [Google Scholar] [CrossRef]
  191. Lv, J.; Liu, J.Y.; Ming, Y.H.; Shi, Y.T.; Song, C.P.; Gong, Z.Z.; Yang, S.H.; Ding, Y.L. Reciprocal regulation between the negative regulator PP2CG1 phosphatase and the positive regulator OST1 kinase confers cold response in Arabidopsis. J. Integr. Plant Biol. 2021, 63, 1568–1587. [Google Scholar] [CrossRef] [PubMed]
  192. Cheng, L.; Zhang, W.H.; Hu, J.L.; Zhang, Z.Y.; Liu, Y.; Lin, Y.Z. Characterization of the key region and putative phosphorylation sites of EcaICE1 in its molecular interaction with the EcaHOS1 protein in Eucalyptus camaldulensis. Plant Biol. 2021, 23, 400–406. [Google Scholar] [CrossRef] [PubMed]
  193. Ye, K.Y.; Li, H.; Ding, Y.L.; Shi, Y.T.; Song, C.P.; Gong, Z.Z.; Yang, S.H. BRASSINOSTEROID-INSENSITIVE2 negatively regulates the stability of transcription factor ICE1 in response to cold stress in Arabidopsis. Plant Cell 2019, 31, 2682–2696. [Google Scholar] [CrossRef] [PubMed]
  194. Bange, M.P.; Milroy, S.P. Impact of short-term exposure to cold night temperatures on early development of cotton (Gossypium hirsutum L.). Aust. J. Agric. Res. 2004, 55, 655–664. [Google Scholar] [CrossRef]
  195. Chen, Y.X.; Zhou, X.J.; Chang, S.; Chu, Z.L.; Wang, H.M.; Han, S.C.; Wang, Y.D. 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]
  196. Shan, W.; Kuang, J.F.; Lu, W.J.; Chen, J.Y. Banana fruit NAC transcription factor MaNAC 1 is a direct target of MaICE 1 and involved in cold stress through interacting with MaCBF 1. Plant Cell Environ. 2014, 37, 2116–2127. [Google Scholar] [CrossRef]
  197. Kovtun, Y.; Chiu, W.L.; Tena, G.; Sheen, J. Functional analysis of oxidative stress-activated mitogen-activated protein kinase cascade in plants. Proc. Natl. Acad. Sci. USA 2000, 97, 2940–2945. [Google Scholar] [CrossRef]
  198. Uchida, K.; Yamaguchi, M.; Kanamori, K.; Ariga, H.; Isono, K.; Kajino, T.; Tanaka, K.; Saijo, Y.; Yotsui, I.; Sakata, Y. MAP KINASE PHOSPHATASE1 promotes osmotolerance by suppressing PHYTOALEXIN DEFICIENT4-independent immunity. Plant Physiol. 2022, 189, 1128–1138. [Google Scholar] [CrossRef]
  199. Zhao, C.Z.; Wang, P.C.; Si, T.; Hsu, C.C.; Wang, L.; Zayed, O.; Yu, Z.P.; Zhu, Y.F.; Dong, J.; Tao, W.A. MAP kinase cascades regulate the cold response by modulating ICE1 protein stability. Dev. Cell 2017, 43, 618–629. [Google Scholar] [CrossRef]
  200. Jiang, X.Y.; Hoehenwarter, W.; Scheel, D.; Lee, J. Phosphorylation of the CAMTA3 transcription factor triggers its destabilization and nuclear export. Plant Physiol. 2020, 184, 1056–1071. [Google Scholar] [CrossRef]
  201. Kameniarová, M.; Černý, M.; Novák, J.; Ondrisková, V.; Hrušková, L.; Berka, M.; Vankova, R.; Brzobohatý, B. Light quality modulates plant cold response and freezing tolerance. Front. Plant Sci. 2022, 13, 887103. [Google Scholar] [CrossRef]
  202. Long, J.L.; Xing, W.; Wang, Y.G.; Wu, Z.D.; Li, W.J.; Zou, Y.; Sun, J.P.; Zhang, F.S.; Pi, Z. Comparative proteomic analysis on chloroplast proteins provides new insights into the effects of low temperature in sugar beet. Bot. Stud. 2022, 63, 18. [Google Scholar] [CrossRef] [PubMed]
  203. Jian, H.J.; Xie, L.; Wang, Y.H.; Cao, Y.R.; Wan, M.Y.; Lv, D.Q.; Li, J.N.; Lu, K.; Xu, X.F.; Liu, L.Z. Characterization of cold stress responses in different rapeseed ecotypes based on metabolomics and transcriptomics analyses. PeerJ 2020, 8, e8704. [Google Scholar] [CrossRef]
  204. Fowler, S.; Thomashow, M.F. Arabidopsis transcriptome profiling indicates that multiple regulatory pathways are activated during cold acclimation in addition to the CBF cold response pathway. Plant Cell 2002, 14, 1675–1690. [Google Scholar] [CrossRef] [PubMed]
  205. Zhu, L.; Li, S.S.; Ma, Q.Y.; Yan, K.Y.; Ren, J.; Chen, Z.; Wen, J.; Li, Q.Z. An Acer palmatum R2R3-MYB gene, ApMYB77, confers freezing and drought tolerance in Arabidopsis thaliana. J. Plant Growth Regul. 2023, 42, 1017–1030. [Google Scholar] [CrossRef]
  206. Yao, C.Y.; Li, X.G.; Li, Y.M.; Yang, G.H.; Liu, W.D.; Shao, B.T.; Zhong, J.L.; Huang, P.F.; Han, D.G. Overexpression of a Malus baccata MYB transcription factor gene MbMYB4 increases cold and drought tolerance in Arabidopsis thaliana. Int. J. Mol. Sci. 2022, 23, 1794. [Google Scholar] [CrossRef]
  207. Liu, X.F.; Zhao, C.D.; Gao, Y.Q.; Xu, Y.; Wang, S.J.; Li, C.S.; Xie, Y.P.; Chen, P.X.; Yang, P.Z.; Yuan, L. A multifaceted module of BRI1 ETHYLMETHANE SULFONATE SUPRESSOR1 (BES1)-MYB88 in growth and stress tolerance of apple. Plant Physiol. 2021, 185, 1903–1923. [Google Scholar] [CrossRef]
  208. Luo, Z.; Kong, X.Q.; Zhang, Y.J.; Li, W.J.; Zhang, D.M.; Dai, J.L.; Fang, S.; Chu, J.F.; Dong, H.Z. Leaf-derived jasmonate mediates water uptake from hydrated cotton roots under partial root-zone irrigation. Plant Physiol. 2019, 180, 1660–1676. [Google Scholar] [CrossRef]
Figure 1. Cold stress impact on plants. When plants are subjected to cold stress, their cell membrane system and chloroplast antioxidant system are disrupted, and the cell membrane changes from a fluid-liquid-crystal state to a solidified gel state, with accelerated synthesis of ROS. Ion channels, calcium channels, RLK, CRLK, and COLD1 act as receptors. Abbreviations: RLK, receptor-like protein kinases; CRLK, calcium receptor-like protein kinases; COLD1, chilling tolerance divergence 1; ROS, reactive oxygen species; SOD, superoxide dismutase; CAT, catalase; POD, peroxidase; GPX, glutathione peroxidase; APX, ascorbate peroxidase; RBOH, respiratory burst oxidase homologue; H2O2, hydrogen peroxide; O2−, superoxide anion; ·OH, hydroxyl radical; OH, hydroxyl; 14-3-3, 14-3-3 protein.
Figure 1. Cold stress impact on plants. When plants are subjected to cold stress, their cell membrane system and chloroplast antioxidant system are disrupted, and the cell membrane changes from a fluid-liquid-crystal state to a solidified gel state, with accelerated synthesis of ROS. Ion channels, calcium channels, RLK, CRLK, and COLD1 act as receptors. Abbreviations: RLK, receptor-like protein kinases; CRLK, calcium receptor-like protein kinases; COLD1, chilling tolerance divergence 1; ROS, reactive oxygen species; SOD, superoxide dismutase; CAT, catalase; POD, peroxidase; GPX, glutathione peroxidase; APX, ascorbate peroxidase; RBOH, respiratory burst oxidase homologue; H2O2, hydrogen peroxide; O2−, superoxide anion; ·OH, hydroxyl radical; OH, hydroxyl; 14-3-3, 14-3-3 protein.
Cells 14 00110 g001
Figure 2. The mechanism of cold stress tolerance in plants. Arrows and lines ending in bars indicate activation and suppression processes, respectively. Solid lines represent direct interactions and dotted lines indicate indirect interactions. See the text for additional details. Abbreviations: ET, ethylene; CK, cytokinins; JA, jasmonate; GA, gibberellins; IAA, indole acetic acid; BRs, brassinosteroids; ABA, abscisic acid; SA, salicylic acid; MET, melatonin; ETR1, ethylene response 1; CTR1, constitutive triple response 1; EIN, ethylene-insensitive; AHK, A. thaliana histidine kinase; AHP, histidine phosphotransfer protein; ARR, A. thaliana. response regulator; JAR1, jasmonate resistant 1; COI1, coronatine insensitive 1; JAZ, jasmonate ZIM-domain protein; MYC2, myelocytomatosis proteins 2; PIF, phytochrome-interacting factor; GID, gibberellin-insensitive dwarf 1; GAI, gibberellic acid insensitive; ARF, auxin response factor; BRI1, brassinazole-resistant 1; BAK1, BRI1-associated receptor kinase 1; BIN, brassinosteroid-insensitive; BZR1, brassinazole-resistant 1; AREB/ABF, ABA-responsive element binding protein/ABRE-binding factor; PYR/RYL/RCAR, pyrabactin resistance 1/PYR-like/regulatory components of ABA receptors; PP2C2, protein phosphatase 2C; OST1, open stomata 1; PIP2, phosphatidylinositol 4,5-bisphosphate; IP3, inositol trisphosphate; CaM, calmodulin; CAMTA, calmodulin-binding transcription activator; CBL, calcineurin B-like protein; CIPK, calcineurin B-like protein; CPK, calcium-dependent protein kinases; MYB, v-myb avian myeloblastosis viral; MAPK, mitogen-activated protein kinases; SIZ1, small ubiquitin-like modifier E3 ubiquitin ligase; HOS1, high expression of osmotically responsive gene 1; BTF3, basic transcription factor 3; NO, nitric oxide; HY5, elongated hypocotyl 5; DELLA, DELLA protein; RBOH, respiratory burst oxidase homologue; ROS, reactive oxygen species; ZAT12, Zinc finger of Arabidopsis thaliana 12; ICE, inducer of CBF expression; CBF, C-repeat binding factor; COR, cold regulated gene.
Figure 2. The mechanism of cold stress tolerance in plants. Arrows and lines ending in bars indicate activation and suppression processes, respectively. Solid lines represent direct interactions and dotted lines indicate indirect interactions. See the text for additional details. Abbreviations: ET, ethylene; CK, cytokinins; JA, jasmonate; GA, gibberellins; IAA, indole acetic acid; BRs, brassinosteroids; ABA, abscisic acid; SA, salicylic acid; MET, melatonin; ETR1, ethylene response 1; CTR1, constitutive triple response 1; EIN, ethylene-insensitive; AHK, A. thaliana histidine kinase; AHP, histidine phosphotransfer protein; ARR, A. thaliana. response regulator; JAR1, jasmonate resistant 1; COI1, coronatine insensitive 1; JAZ, jasmonate ZIM-domain protein; MYC2, myelocytomatosis proteins 2; PIF, phytochrome-interacting factor; GID, gibberellin-insensitive dwarf 1; GAI, gibberellic acid insensitive; ARF, auxin response factor; BRI1, brassinazole-resistant 1; BAK1, BRI1-associated receptor kinase 1; BIN, brassinosteroid-insensitive; BZR1, brassinazole-resistant 1; AREB/ABF, ABA-responsive element binding protein/ABRE-binding factor; PYR/RYL/RCAR, pyrabactin resistance 1/PYR-like/regulatory components of ABA receptors; PP2C2, protein phosphatase 2C; OST1, open stomata 1; PIP2, phosphatidylinositol 4,5-bisphosphate; IP3, inositol trisphosphate; CaM, calmodulin; CAMTA, calmodulin-binding transcription activator; CBL, calcineurin B-like protein; CIPK, calcineurin B-like protein; CPK, calcium-dependent protein kinases; MYB, v-myb avian myeloblastosis viral; MAPK, mitogen-activated protein kinases; SIZ1, small ubiquitin-like modifier E3 ubiquitin ligase; HOS1, high expression of osmotically responsive gene 1; BTF3, basic transcription factor 3; NO, nitric oxide; HY5, elongated hypocotyl 5; DELLA, DELLA protein; RBOH, respiratory burst oxidase homologue; ROS, reactive oxygen species; ZAT12, Zinc finger of Arabidopsis thaliana 12; ICE, inducer of CBF expression; CBF, C-repeat binding factor; COR, cold regulated gene.
Cells 14 00110 g002
Figure 3. A model to explain transcription factors of cold tolerance in plants. By regulating a large number of transcription factors, the mechanisms of triggering a series of responses to plant cold tolerance can be redesigned. Abbreviations: CaMBP, CaM-binding proteins; bZIP, basic leucine zipper; AREB/ABF, ABA-responsive element binding protein/ABRE-binding factor; HY5, hypocotyl 5; MYB, v-myb avian myeloblastosis viral; WRKY, WRKY transcription factors; NAC, NAC transcription factors; CAMTA, calmodulin-binding transcription activator; AP2/ERF, aptala2/ethylene response factor; bHLH, basic helix–Loop–helix; MYC, myelocytomatosis proteins; PIF, phytochrome-interacting factor; ICE, an inducer of CBF expression; CBF, C-repeat binding factor; COR, cold regulated gene.
Figure 3. A model to explain transcription factors of cold tolerance in plants. By regulating a large number of transcription factors, the mechanisms of triggering a series of responses to plant cold tolerance can be redesigned. Abbreviations: CaMBP, CaM-binding proteins; bZIP, basic leucine zipper; AREB/ABF, ABA-responsive element binding protein/ABRE-binding factor; HY5, hypocotyl 5; MYB, v-myb avian myeloblastosis viral; WRKY, WRKY transcription factors; NAC, NAC transcription factors; CAMTA, calmodulin-binding transcription activator; AP2/ERF, aptala2/ethylene response factor; bHLH, basic helix–Loop–helix; MYC, myelocytomatosis proteins; PIF, phytochrome-interacting factor; ICE, an inducer of CBF expression; CBF, C-repeat binding factor; COR, cold regulated gene.
Cells 14 00110 g003
Figure 4. The mechanism of cold tolerance in some plants. The intercellular signal (e.g., phytohormone, osmoregulatory and inorganic ions) binds to the receptor under cold stress and is converted into an intracellular signal through the signal transduction system on the cell membrane (e.g., ion channels, RLK, and protein kinases). Cold can also be directly translated into intracellular signals. Abbreviations as above.
Figure 4. The mechanism of cold tolerance in some plants. The intercellular signal (e.g., phytohormone, osmoregulatory and inorganic ions) binds to the receptor under cold stress and is converted into an intracellular signal through the signal transduction system on the cell membrane (e.g., ion channels, RLK, and protein kinases). Cold can also be directly translated into intracellular signals. Abbreviations as above.
Cells 14 00110 g004
Table 1. Positive and negative impacts of cold on plants.
Table 1. Positive and negative impacts of cold on plants.
ImpactsClassContentReferences
PositiveCold acclimationImproving plant cold tolerance[18]
VernalizationImprovement of flowering[19]
Improvement of seed yield[20]
NegativeInsufficient leaf developmentReducing leaf elongation[21]
Leaf chlorosis (wilting, even necrosis)[22]
Reducing stomatal conductance[23]
Insufficient root developmentSwelling root tips[24]
Thicker root axis[16]
Less lateral and more seminal roots[25]
Reducing root length[26]
Growth retardationSpikelet sterility[27]
Limiting seed size[28]
Lower survival rate[29]
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

Feng, Y.; Li, Z.; Kong, X.; Khan, A.; Ullah, N.; Zhang, X. Plant Coping with Cold Stress: Molecular and Physiological Adaptive Mechanisms with Future Perspectives. Cells 2025, 14, 110. https://doi.org/10.3390/cells14020110

AMA Style

Feng Y, Li Z, Kong X, Khan A, Ullah N, Zhang X. Plant Coping with Cold Stress: Molecular and Physiological Adaptive Mechanisms with Future Perspectives. Cells. 2025; 14(2):110. https://doi.org/10.3390/cells14020110

Chicago/Turabian Style

Feng, Yan, Zengqiang Li, Xiangjun Kong, Aziz Khan, Najeeb Ullah, and Xin Zhang. 2025. "Plant Coping with Cold Stress: Molecular and Physiological Adaptive Mechanisms with Future Perspectives" Cells 14, no. 2: 110. https://doi.org/10.3390/cells14020110

APA Style

Feng, Y., Li, Z., Kong, X., Khan, A., Ullah, N., & Zhang, X. (2025). Plant Coping with Cold Stress: Molecular and Physiological Adaptive Mechanisms with Future Perspectives. Cells, 14(2), 110. https://doi.org/10.3390/cells14020110

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