Next Article in Journal
Natural Control of Weed Invasions in Hyper-Arid Arable Farms: Allelopathic Potential Effect of Conocarpus erectus against Common Weeds and Vegetables
Previous Article in Journal
The Potential of Pig Sludge Fertilizer for Some Pasture Agricultural Lands’ Improvement: Case Study in Timiș County, Romania
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plant Low-Temperature Stress: Signaling and Response

1
Center for Genomics and Biotechnology, College of Agriculture, Fujian Agriculture and Forestry University, Fuzhou 350002, China
2
Department of Biology, University of Massachusetts, Amherst, MA 01003, USA
3
College of Plant Science & Technology, Huazhong Agricultural University, Wuhan 430070, China
4
Guangxi Key Lab of Sugarcane Biology, State Key Laboratory for Conservation and Utilization of Subtropical Agro-Bioresources, College of Agriculture, Guangxi University, Nanning 530004, China
*
Authors to whom correspondence should be addressed.
Agronomy 2022, 12(3), 702; https://doi.org/10.3390/agronomy12030702
Submission received: 12 December 2021 / Revised: 4 March 2022 / Accepted: 9 March 2022 / Published: 14 March 2022

Abstract

:
Cold stress has always been a significant limitation for plant development and causes substantial decreases in crop yield. Some temperate plants, such as Arabidopsis, have the ability to carry out internal adjustment, which maintains and checks the metabolic machinery during cold temperatures. This cold acclimation process requires prior exposure to low, chilling temperatures to prevent damage during subsequent freezing stress and maintain the overall wellbeing of the plant despite the low-temperature conditions. In comparison, plants of tropical and subtropical origins, such as rice, are sensitive to chilling stress and respond differently to low-temperature stress. Plants have evolved various physiological, biochemical, and molecular mechanisms to sense and respond to low-temperature stress, including membrane modifications and cytoskeletal rearrangement. Moreover, the transient increase in cytosolic calcium level leads to the activation of many calcium-binding proteins and calcium-dependent protein kinases during low-temperature stress. Recently, mitogen-activated protein kinases have been found to regulate low-temperature signaling through ICE1. Besides, epigenetic control plays a crucial role during the cold stress response. This review primarily focuses on low-temperature stress experienced by plants and their strategies to overcome it. We have also reviewed recent progress and previous knowledge for a better understanding of plant cold stress response.

1. Introduction

Cold stress is among important abiotic stresses that substantially controls the crop production and geographical distribution of many plant species [1,2]. Low-temperature stress affects many aspects of plants, including germination, growth, development, and reproduction (Figure 1) [3,4,5,6]. Generally, plants encounter two forms of low-temperature stress i.e., chilling and freezing. For plants, chilling temperatures are low but positive temperatures (0–15 °C) that could vary with the plant’s tolerance level and variety. Chilling temperatures also depend upon air temperature and wind speed during the exposure. In contrast, at freezing temperature (below 0 °C), the plant starts to struggle against freezing injury [7].
Rice (Oryza sativa L.), is susceptible to chilling stress, especially in high elevation and high latitude temperate zones [8,9]. Chilling temperature for more than 4 days results in poor germination, retarded seeding growth, sometimes death of rice seedlings. Besides, some researchers also reported a large-scale epidemic of rice blast during chilling stress, resulting in a significant loss to rice production [10]. At chilling temperatures, sensitive plants are affected at all stages (from vegetative to reproductive stages) of their life. In chillling sensitive plants, chilling injury can be observed in the form of chlorosis, stiffness of seedlings, withering and often the death of seedlings. These phenotypes often work as indicators to differentiate chilling sensitive and tolerant varieties [9].
However, many temperate plants, such as Arabidopsis, rapeseed, wheat, and rye, have developed abilities that detect and respond to freezing stress via cold acclimation. The prerequisite for cold-acclimation is exposure to a low, non-freezing temperature that brings about coping ability during subsequent cold stress. The cold acclimation mechanism maintains the overall wellbeing of the plant despite the non-conducive low-temperature conditions [11].
Perennial plants, such as trees, detect the impending season by sensing changes in photoperiod and temperature throughout the year. Seasonal day length and temperature variation signals trigger transitions from active growth to dormant phases and frost sensitivity to cold hardiness in these plants. This shift is obligatory for woody plants to survive the winter in temperate climates. Compared to annual herbaceous plants, the morphological and physiological characteristics of overwintering plants help them synchronize with climatic rhythms resulting in their survival in harsh cold habitats [12]. Endogenous regulating systems such as photoreceptors and the circadian clock assess light changes and measure photoperiods. Even under optimal conditions, if day length goes below a certain range, growth cessation and development of apical buds are stimulated in trees. Simultaneous with the growth cessation, overwintering tissues increase freezing tolerance. Previous studies have suggested that either short-day or low temperatures can substantially enhance the freezing tolerance of trees. However, seasonal cold acclimation of woody perennials is predominantly induced by short-day during autumn, which is further assisted by low temperature. For example, under short-day conditions, hardwood trees such as Silver Birch and Populus spp. develop initial provisional freezing tolerance, which increases further upon subsequent exposure to low nonfreezing and freezing temperatures [13,14,15].
During low-temperature stress, multiple pathways such as protein and nucleic acid biosynthesis, respiratory, defence, and secondary metabolism are affected in these plants. Consistently, the expression level of different genes involved in carbohydrate metabolism, hormone biosynthesis, transcripts linked to membranes lipids and polysaccharides change due to low temperature stress [16]. Guy et al. (1985) first reported a group of genes contributing to freezing tolerance and named them cold-regulated (COR) genes. In Arabidopsis, COR genes include early dehydration-inducible (ERD) genes, low-temperature induced (LTI), responsive to desiccation (RD), etc. [17]. Since then, considerable progress has been made in our understanding of molecular mechanisms of cold stress response in plants.
In this review article, we have summed up primary responses exhibited by the plant during low-temperature stress and included recent findings of plant cold stress research.

2. Perception of Cold Signal

2.1. Membrane Modification as a Signal of Low-Temperature Stress

The plasma membrane of an organism is similar to a safety net that protects and maintains the cell shape and discriminately exchanges proteins, signal molecules, gases, and other substances required for various physiological and developmental processes [18]. This transport is facilitated either by open spaces in the plasma membrane or different transport proteins, pumps and gated channels [19,20,21]. The major components of the plasma membrane include lipids, proteins, and carbohydrates. Cold stress initially causes injury to the plasma membrane by altering its physical and chemical properties [8]. The plasma membrane adopts unconventional ways to produce stress-inducible lipids through the endoplasmic reticulum in response to cold stress. Remarkably during cold stress, glycerolipids function as signal molecules besides protecting organellar membranes [22].
The lipid composition of lipid bilayers and their associated lipid head groups (esterified fatty acids form) vary at different growth temperatures [23]. Cold stress is not an exception in this case. For instance, during 1980, several groups reported that the high melting point fatty acids (hexadecanoic, trans-3-hexadecenoic and octadecanoic acid) and saturated fatty acids (hexadecanoic and octadecanoic acids) differ among higher plants during their response to chilling stress [24,25,26,27,28]. An excellent and comprehensive study by Norio Murata further sheds light on the role of phosphatidylglycerols during the chilling stress tolerance. It was demonstrated that there is a clear difference between chilling stress tolerant and sensitive plants for phosphatidylglycerol, dipalmitoyl plus the l-palmitoyl-2-(trans-3-hexadecenoyl) contents, which is around 3–19% in chilling resistant plants compared to 26–65% in the susceptible plants [29]. Additionally, the chilling sensitivity of plants depends on the degree of unsaturated fatty acids in the phosphotidylglycerol residues in the chloroplast membranes [30]. Chilling-resistant plants, such as spinach and Arabidopsis thaliana, contain more cis-unsaturated fatty acids, mediated by chloroplast-localized enzyme glycerol-3-phosphate acyltransferase, compared to chilling-sensitive plants like Nicotiana tabacum. Consistently, transformation with cDNA of glycerol-3-phosphate acyltransferase alters fatty acid unsaturation and chilling stress tolerance in several plants [30].
Moreover, lipid flippases of P-type ATPases can exchange lipid molecules between membrane leaflets [31]. In cold-stressed plants, following functional deformities in the plasma membrane, cytosolic pH declines to an extent where tonoplast starts crumpling [32,33]. Mechanosensitive Ca2+ channels are suggested to be involved in this breakdown of the vacuolar membrane [34]. Appealingly, cold-acclimated plants express tonoplast monosaccharide transporters (TMT), which increase the level of sugars and provide cellular homeostasis [35]. The flexibility of the chloroplast membrane is also compromised after a period of cold [36]. The process involves modification in acyl-lipid/CoA desaturase 1 gene, ensuring a decrease in fatty acid composition [37]. Additionally, a thylakoid localized lipoxygenases (LOX) reversibly communicates with thylakoid membrane lipids under dark, chilling conditions. LOXs are involved in the first step of the biosynthesis of oxylipin, one of the many stress inhibitors which increase plant tolerance to stresses [38]. Besides, a nonessential plastid-encoded ribosomal protein in tobacco, Rpl33, is required for cold stress tolerance and in the Rpl33 knockout plants the recovery from chilling stress is severely compromised [39]. These reports collectively indicate that most plants modify their membrane as a necessary physiological process to combat low-temperature stress.
Apart from the membrane modification, changes in the plant cytoskeleton are also implicated in cold stress sensing. The cold exposure for only a few minutes can affect the cytoplasmic architecture and organization of the cytoskeleton [40]. The cytoskeleton maintains the structural integrity by reorganizing or disassembling during cold stress exposure. Besides, the cold-induced depolymerization of the cytoskeleton is believed to be necessary to induce low temperature-responsive genes [40,41,42,43].

2.2. Ca2+ Signaling in Response to Cold

Calcium (Ca2+) is a versatile signal transduction element, which acts as a secondary messenger and regulates several physiological processes [44,45]. Cytosolic Ca2+ concentration increases transiently due to different abiotic and biotic stimuli in the living organism, including plants [46]. These Ca2+ signatures are specific for each cue in terms of frequency, duration, amplitude, and location (Figure 2) [47]. The prolonged period of high cytosolic Ca2+ could cause reactive oxygen species (ROS) accumulation and metabolic dysfunctions leading to physiological damage to the cell [48]. The physical alterations in the plasma membrane during cold stress also change the Ca2+ concentration [1,49]. Cold-induced signals that curb intracellular Ca2+ sometimes lead to the initiation of a protein phosphorylation cascade, which starts several other signaling molecules apart from activating second messengers.
Calcium channels, including cyclic nucleotide-gated channels (CNGCs), nonspecific cation channels, contribute to Ca2+ fluxes in various stress responses [50]. In plants like Arabidopsis and rice, CNGCs form a large family of non-selective cation-conducting channels that are mostly localized at the plasma membrane. In response to cold stress, the expression levels of rice CNGCs have been shown to be differentially regulated, suggesting their role in cold stress response [51]. Consistently, the G-protein regulator Chilling Tolerance Divergence 1 (COLD1), in combination with Rice G-protein a Subunit 1 (RGA1), regulates the cold-induced influx of Ca2+ to confer cold sensing in rice [52]. However, it is unclear whether COLD1 acts as a Ca2+ channel or only modulates its activity.
The calcium channels also play an important role in the cold acclimation of Arabidopsis [53]. For example, Mid1-Complementing Activity 1 (MCA1) and MCA2 are mechanosensitive Ca2+ permeable channels contributing to cold-induced Ca2+ influx; although they are not the only ones transporting Ca2+ across the plasma membrane [54]. Recently Liu et al. (2021) reported that AtANN1, a Ca2+ permeable transporter and a member of the subfamily of phospholipid and Ca2+ binding proteins, participates in freezing tolerance by mediating the generation of Ca2+ signals [55].
The cytosolic Ca2+ signal is primarily transmitted through Ca2+ sensors. Upon sensing the elevation in Ca2+ concentration, these sensors change their phosphorylation status by binding to calcium in their specific motifs (EF-hands). To date, several calcium sensors such as calmodulin (CaM), calcineurin B-like proteins (CBLs), CBL-interacting protein kinases (CIPKs), calcium-dependent protein kinases (CDPKs) and phosphatases have been reported [56]. Two important cold-induced calcium sensors are discussed below.

2.2.1. Calmodulin (CaM) Involvement in Cold Stress Signaling

Calmodulins (CaM) are Ca2+ binding proteins that can sense the changes in the concentration of Ca2+ ions and function as a calcium buffer. CaMs transduce signals and maintain a homeostatic balance of Ca2+ to minimize its cytotoxic effects. Single plant species may have several isoforms of CaM, displaying a wide array of responses [57,58]. CaM has four functional EF-hands that cause conformational changes of CaM upon binding with Ca2+. After attaching to Ca2+, CaM activates several other protein targets involved in numerous cellular processes [58]. During environmental cues such as low temperature, rapid transcription of CaM proteins are reported in several plants, including Arabidopsis [59,60,61].
Calmodulin-binding transcription activators (CAMTA) proteins belong to a class of transcription factors conserved in eukaryotes and respond to calcium signals by binding to calmodulin. Arabidopsis contains six CAMTA genes [62] that act as a transcriptional regulator after binding to a specific (G/A/C)CGCG(C/G/T) cis-element. CAMTA proteins perceive an increase in Ca2+ level as a response to cold stress by interacting with various calmodulin proteins [59]. Doherty, et al. [63] first suggested the role of CaM signaling during cold stress. They showed that CAMTA3 functions as a positive regulator of the CBF2 transcription factor by binding to the cis-element (CM2 motif) of CBF2. Consistently in the camta1camta3 double mutant, the CBF gene family expression was reduced by 50% approximately [63]. Since then, several reports have suggested the involvement of the CAMTA during cold stress [64,65,66,67].

2.2.2. CBL-CIPK Module in Cold Stress Response

Calcineurin like (CBL) proteins (also SOS3-like Ca2+ binding proteins, SCaBLs) are a family of Ca2+ sensors identified in different plant species, including Arabidopsis [68]. These CBLs/SCaBPs do not possess enzymatic activity by themselves, however, CBLs are triggered by the change in Ca2+ signature and precisely interact with CIPK (CBL-interacting protein kinases) or SOS2-like protein kinase [69]. Increasing evidence suggests that the CBL-CIPK network plays a key role during cold stress response. In Arabidopsis, CBL1 participates in cold stress signaling along with CIPK7 [70]. Similarly, in Cassava (Manihot esculenta) MeCBL2 is induced by cold stress with MeCIPK7 suggesting their role in cold signal transduction [71]. In 6-week-old grapevine leaves, VvCBLs and VvCIPKs showed differential expression against various stress conditions. Heat stress down-regulated the expression level of VvCBL10, VvCBL11 and VvCBL12; however, their expression increased by salt, PEG and cold stress. Besides, VvCIPK34 transcript was increased during cold and heat stress, but decreased by salt and PEG treatments [72]. Similarly, in pea plants, exposure of NaCl, wounding and cold up-regulated PsCBL and PsCIPK, whereas drought and ABA did not display changes in transcript level of PsCBL and PsCIPK [73]. Abiotic treatments in canola also increased the transcript levels of BnaCBLs and BnaCIKPs. Six hours of cold stress up-regulated the expression of BnaCBL1, but BnaCBL10 took 24 h, whereas BnaCBL2, −3, −4 were down-regulated by cold exposure. Besides, a significant up-regulation of BnaCIPK3, −6, −12, −15, −23, −26 were found during cold, suggesting their participation in cold stress signaling [74]. Consistently, rice plants over-expressing CIPK genes displayed enhanced resistance to drought, salt, and cold stresses. OsCIPK03, OsCIPK12, and OsCIPK15 over-expressing plants showed an increase in resistance against abiotic stresses. Moreover, compared to wild-type plants, they accumulated a higher level of compatible solutes and proline [75]. Similarly, ectopic expression of Lepidium CIPK gene in N. tabacum resulted in an increased level of proline accumulation, cell membrane stability and cold stress tolerance [76,77]. These findings indicate that the CBL-CIPK module is a crucial signaling module engaged in cold stress signaling. However, the functions of several CIPK proteins remain elusive in Arabidopsis and other important crop plants. Further studies will help us to understand the specificity of CBL-CIPK signaling pathway, which could be utilized as a tool in engineering stress hardy plants.

3. Transcriptional Regulation: CBF and Regulation of Cold

Transcription factors (TFs) display a pivotal role in gene regulation by binding to cis-elements in the promoter region of the target gene. The concerned gene is either turned off/on via TF mediated regulation and brings about a particular outcome. There are approximately 58 families of different TFs that have been identified in plants [78]. The majority of the TFs belong to multigene families [79]. Structurally, TFs are multidomain proteins with DNA-binding domain and activation domain [79]. Individual members of the same family may respond differently to different stress conditions. However, a significant overlap of the gene-expression profiles suggests that some genes get induced by the same transcription factor [80]. Several transcription factors are reported to be involved in low-temperature stress responses such as C-repeat binding factors (CBFs), basic helix-loop-helix (bHLH), NAC, MYB, etc. [81,82,83,84]. In Arabidopsis, many transcription factor genes are induced during low-temperature stress [85].
The DREB/CBF-mediated cold-responsive pathway is well studied because of the upregulation of CBF1-3/DREB1A-C due to cold stress [7,86,87]. Overexpression of CBF/DREB induces cold tolerance in Arabidopsis and other plants as well [88,89,90]. Interestingly, CBF/DREB transcription factors lack the DRE element in their promoter region and are unable to regulate their own expression. But, some of the upstream regulators, such as ICE (inducer of CBF expression) [91,92], CAMTA3 (calmodulin-binding transcription activator 3) [63], ZAT12 [93], and EIN3 (Ethylene Insensitive 3) [94] regulate the expression of CBF/DREB. Among ICEs, ICE1 and ICE2 are the most upstream cold-inducible gene and consequently, they regulate downstream CBFs and CORs. Both ICE1 and ICE2 encode MYC-like basic helix-loop-helix (bHLH) transcription activator and binds to the promoter regions of CBF1-3 to regulate their expression positively [92,95,96]. CBF transcription factors have been characterized extensively by using transgenic plants that have the ectopic expression of CBF/DREB1 genes, RNAi or CRISPR/Cas9 editing [97,98,99,100]. Overexpression of CBF increases the expression of COR genes resulting in cold acclimation even at optimum temperatures [87,97,101]. The transcript level of ICE1 itself is not induced by cold, suggesting that ICE1 function is post-translationally regulated [91]. In recent years, several factors identified to regulate ICE1 expression, including HOS1 (high expression of osmotically responsive gene 1, SIZ1 (SAP and Miz), and OST1 (open stomata 1) [102,103,104]. HOS1 acts as a RING-type ubiquitin E3 ligase and is also reported to function as a chromatin remodeling factor. HOS1 negatively regulates the CBF expression by ubiquitinating and degrading ICE1 during low-temperature stress [102,105].
Moreover, SIZ1 mediated sumoylation of ICE1 controls the CBF expression by stabilizing ICE1 [106]. A transcription factor, MYB15 (a member of the R2R3-MYB family), also negatively regulates the expression of CBF genes in Arabidopsis [107]. MYB15 is expressed in the absence of cold stress and represses the expression of CBFs by binding to MYB recognition sites of their promoters [84,108]. Additionally, ZAT12, a C2H2 zinc finger transcription factor, also functions as a negative regulator of CBF. Arabidopsis plants overexpressing ZAT12 decrease the expression level of CBF under cold stress [109]. As discussed above, during cold stress several negative regulatory pathways work on the finetuning of CBF-dependent signaling. However, the physiological significance of repressive pathways still remains elusive. Probably, high levels of CBF could be harmful to plant growth and a balance is maintained in the form of these repressors. Indeed, CBF regulon plays an indispensable role in cold acclimation, though it is not the only player. These studies have significantly contributed to our understanding of CBF pathway and its significance during low-temperature stress. However, recently, the established ICE-CBF pathway has been challenged. Kidokoro et al. demonstrated that neither ice1-1 mutant nor ICE1 overexpression regulates the expression of DREB1A/ICE3 [110,111]. At first hand, the finding sends a shock wave among the cold stress community, but their results also showed that the promoter region DREB1A is hypermethylated due to another transgene, not ICE1 [110,111]. Although the ICE-CBF pathway is well-studied and established for cold stress signaling, it is eye-opening that suggests revalidating the pathway and findings associated with this signaling cascade.

4. Mitogen-Activated Protein Kinase: The Attenuator of ICE1

MAPKs (mitogen-activated protein kinases) play an indispensable role in integrating multiple intracellular signals transmitted by various stress-induced secondary messengers, thereby activating or deactivating proteins via post-translational phosphorylation of target [112]. Generally, the MAPK cascade consists of three kinases acting in a row and inactive MAPKKKs are activated by a stimulus or signal messenger. Once activated, MAPKKKs activate MAPKKs by phosphorylating it at conserved serine/threonine. Activated MAPKKs further activate MAPKs by phosphorylating MAPK. Recently, the involvement of MAPK signaling has been found to regulate the cold stress response via the ICE1 pathway in A. thaliana [113,114]. MPK3 and MPK6 phosphorylate ICE1, which inhibits CBF expression, whereas MEKK1-MKK1/2-MPK4 cascade enhances cold tolerance by acting antagonistically to MPK3/MPK6 (Figure 3) [114].
MPK3, MPK4, and MPK6 are induced within 30 min of the cold exposure indicating their involvement in cold stress response. Consistently, mpk3/mpk6 mutant display enhanced freezing tolerance; however, mpk3 and mpk6 inducible MKK5DD displayed opposite results [113,114]. When subjected to cold MPK mediated phosphorylation leads to degradation of ICE1, resulting in the fine-tuning of ICE1 mediated CBF regulon [113,114]. Interestingly, earlier reports showed that OST1 mediated phosphorylation increased ICE1 stability [104]. Additionally, OST1 increases the stability of ICE1 by antagonizing HOS1 E3 ubiquitin ligase [102]. In comparison, MPK3/6 mediated degradation of ICE1 does not affect the HOS-ICE interaction and could be working differently [113]. These findings suggest a ying-yang mechanism in ICE1 stability by two different types of kinases. Deciphering the puzzle of MPK3/MPK6 mediated ICE1 degradation mechanism will further enhance our knowledge of ICE1 regulation and turnover. In contrast to MPK3/MPK6, MEKK1-MEKK1/2-MPK4 cascade works under calcium/calmodulin-regulated receptor-like kinase1 (CLRK1) during cold stress. CRLK1 and CRLK1 induced MPK4 inhibits MPK3/6 activation, thereby positively regulating freezing tolerance [114]. CRLKs induce downstream MEKK cascade, which further amplifies the downstream signal (Figure 3). The MPK4 mediated inhibition of MPK3/6 promotes the stability of ICE1, thereby modulating the cold-responsive transcription factor-like CAMTA and MYB [64,108]. These studies demonstrate that fine-tuning of cold stress signaling is an essential step, and sometimes a signaling component could work in both directions as an activator/suppressor. It also corroborates that calcium sensors such as CRLKs play an indispensable role in cold signal transduction.

5. The Role of MicroRNAs during the Cold Stress Response

MicroRNAs (miRNAs) are small non-coding RNAs that play a vital role in post-transcriptional gene regulation. In general, microRNAs size range from 20 to 22 nucleotides for animals and 20 to 24 nucleotides for plants. miRNAs act as negative regulators of translation by binding to complementary mRNA molecules and guiding degradation and/or translational repression of the target [115]. It is now well established that the miRNAs play a crucial role in almost every aspect of growth, development, response to biotic and abiotic factors, response to various environmental challenges by post-transcriptionally regulating the gene expression [116,117,118]. MicroRNA’s role during low-temperature response and other environmental cues has been well documented [119,120,121,122]. Several miRNAs have been identified for their involvement in low-temperature stress in Arabidopsis (see Megha et al., 2018 for more details, Supplementary Table S1). In various reports, several cold stress-responsive microRNAs have been reported in several plant species [117,122,123]. The up-regulation of several microRNAs viz. miR168, miR169, miR172, miR393, miR319, and miR408 during cold stress has been observed in different plant species [122,124]. Additionally, the heterologous expression of rice miR393a in switchgrass significantly increased cold tolerance through the regulation of auxin signaling [125]. Recently, 353 cold-regulated miRNAs were identified by comparing cold-tolerant and cold-sensitive varieties of Brassica. This study further suggested that miR166e, miR319, and Bra-novel-miR3936-5p may play essential roles during cold stress response [126]. Yang, et al. [127] also obtained 412 sugarcane miRNAs (261 known and 151 novel miRNAs) by deep sequencing. They reported 62 miRNAs were differentially regulated by cold and selected miR156 for functional analysis in tobacco. Transient overexpression of miR156 displays improved parameters of cold stress tolerance [127]. In the young spikes of common wheat, 192 conserved miRNAs and 9 novel miRNAs were identified [117]. A total of 34 conserved miRNAs and five novel miRNAs were differentially regulated between the cold-stressed and control samples. In another study, miRNA396b (ptr-miR396b) enhanced cold tolerance in transgenic lemon (Citrus limon) plants overexpressing ptr-miR396b as compared to wild-type. Ptr-miR396b guided the cleavage of the 1-aminocyclopropane-1-carboxylic acid oxidase (ACO, a rate-limiting enzyme in ethylene biosynthesis) and overexpressing lines displayed a reduction in ACO transcript levels and ethylene content compared with the WT. The low level of ethylene under cold stress could be because of ACO transcript reduction by Ptr-miR396b [128]. Moreover, ethylene targets the CBF pathway and acts as a negative regulatory signal in cold stress response [94]. Interestingly, a recent paper describes that a transcriptional repressor of auxin signaling, SLR/IAA14 protein, plays a crucial role in integrating miRNAs in auxin and cold stress response [129].
Collectively, these reports indicate that miRNAs play a crucial role during cold stress response by attenuating the gene expression and degrading the target. mRNA targets degraded by miRNAs could be deleterious and energy-consuming if translated during cold stress.

6. Antioxidants during the Cold Stress Response

Reactive Oxygen Species (ROS) are many chemically reactive compounds produced during oxygen metabolism. Examples of ROS include peroxide, superoxide, hydroxyl radical & singlet oxygen [130]. Adverse conditions like various environmental stresses (high light intensity, metals, high toxic levels of oxygen and freezing temperature) and biotic stresses like infections bring about oxidative injuries due to the overproduction of ROS [131,132]. Their overabundance oxidizes several types of molecules like proteins, lipids, etc. [133]. Additionally, ROS disrupts the double-layer membrane integrity by disabling its receptors and enzymes. ROS also interferes with DNA structure and its repair mechanism [134]. To restore redox homeostasis, plant activates several genes encoding defensive enzymes such as glutathione peroxidase, glutathione and ascorbate. In addition, enzymes of the ascorbate-glutathione cycle (viz. ascorbate peroxidase, dehydroascorbate reductase and glutathione reductase), fat-soluble vitamins like tocopherol and carotenoid predominantly assist in hunting ROS [135,136,137,138,139]. Since cold stress accompanies low to freezing temperatures that vary with geographical distribution, hence plants too exhibit the variety of antioxidant mechanisms depending upon the severity of chilling stress. Chloroplasts are the worst affected by peroxides, which is a crucial feature during chilling temperature with high light stress [140,141,142]. Chloroplast ascorbate peroxidase is a foreground enzyme during cold and high light stress [143,144,145]. Its reduced activity demonstrates an overall fall in photosynthesis and an increase of free radicals (peroxides and non-functional proteins) due to oxidation [146]. This causes lipid peroxidation of organelle membranes and other lipid inclusions besides disrupting metabolic pathways (the reductive carbon pathway and electron transport) and stomatal conductance [132,147]. Meanwhile, recent advances in research provide insight on the interaction of chloroplast and mitochondria in maintaining hydrogen peroxide equilibrium through malate shuttle, in addition to inducing programmed cell death (Figure 4) [148]. Hence, we can state that oxidative stress is an unavoidable side effect that follows low-temperature stress [149,150,151].

7. Contribution of Phytohormones during the Cold Stress Response

Phytohormones are well known for the regulation of growth and development in plants. Being small, they are capable of moving to a distant location from the place of their biosynthesis and carry out the appropriate function in exceptionally low concentrations [152]. Abscisic acid (ABA) is a bonafide stress hormone that plays key functions during stress [153]. Several reports suggest that ABA concentration elevates against stress in conjunction with the regulation of several genes [154,155,156]. During stress conditions, carotenoids are converted into bioactive ABA by de novo biosynthesis pathway. ABA biosynthesis is increased during cold stress, which improves plant ability to resist during unfavorable cold conditions [157,158]. Consistently, mutants of ABA biosynthetic genes display impaired cold acclimation response, whereas exogenous application of ABA improves the cold tolerance [159,160,161]. ABA biosynthesis and catabolic genes display their regulation in organ dependent manner during cold stress [162]. Plants subjected to cold stress induce a number of ABA-responsive genes [163,164]. Several COR genes that are induced during cold stress are also induced by ABA, indicating the involvement of ABA signal transduction in cold stress responses [165]. These observations suggest that ABA is an essential phytohormone participating in cold stress response.
Several findings suggest that ABA does not affect the CBF expression and believed to mediate the cold response through the CBF-independent pathway [166,167]. In support to this hypothesis, a set of ABA-independent, cold- and drought-induced gene contains Dehydration-Responsive/C-Repeat Element/Low-temperature Responsive Element (DRE/ CTR/LTRE) cis-acting element in their promoter regions [168,169,170]. However, some recent findings indicate that ABA-mediated cold response could be mediated by CBF-dependent pathway as well [171]. For instance, both ABA and low temperature induce the expression of transcription factor MYB96 [96]. MYB96 interacts with HHP1-3 (HEPTAHELICAL PROTEIN) and these HHP proteins interact with ICE1, ICE2, and CAMTA [96,172]. Altogether, it shows the possibility of ABA-mediated low-temperature stress response in CBF-dependent pathway [171].
Another phytohormone, auxin (indole-3-acetic acid, IAA) has been established to control virtually all aspects of plant development and response. Emerging evidences indicate the involvement of auxin in regulation of plant growth and development under high as well as low temperature stresses [173,174,175,176,177,178,179]. GH3 genes encode auxin-conjugating enzyme and modulate endogenous levels of active auxin through negative feedback regulation. In rice, overexpression of OsGH3-2 increased cold tolerance and reduced free IAA content, alleviated oxidative damage, and decreased membrane permeability [180]. Low temperature-induced root growth inhibition of Arabidopsis is also reported as a response of reduced auxin accumulation [181]. Recently, the involvement of auxin has been shown to preserve the root stem cells at a quiescent status under chilling stress [182]. Phytohormones cytokinin signal transduction is mediated by the two-component system (TCS), which is perceived by histidine kinase receptors and regulate the phosphorylation of response regulators [183]. Involvement of the Arabidopsis cytokinin TCS during cold stress signaling and response has been shown in various reports [181,184]. Moreover, Cytokinin Response Factors (CRF2) and CRF3 are involved in lateral root formation and initiation during cold stress whereas, CRF4 gets induced during cold and contributes to freezing tolerance [185,186].
In addition to ABA, auxin and cytokinin other phytohormones are also involved in cold stress response. For example, after low-temperature stress, the accumulation of ethylene has been reported in several plant species including Arabidopsis, tomato, winter rye, etc. [187,188,189]. Ethylene mediates physiological, developmental and stress responses through the activation of Ethylene Response Factors (ERFs) transcription factors [190]. Recently, ERF105 has been shown to operate in conjunction with the CBF pathway and play a key role during cold acclimation and freezing tolerance [191]. Additionally, high ethylene content increases freezing tolerance in non-acclimated Arabidopsis plants. Consistently, the exogenous application of ACC also enhances the survival rate during freezing stress [192]. Moreover, increased freezing response by ethylene overproducing mutant eto1-3 further supports the role of ethylene during freezing stress [192]. In contrast, Shi et al. (2012) showed that ethylene negatively regulates the freezing tolerance in Arabidopsis, accounting for significant sensitivity in ethylene overproducer mutants. They further showed that the ethylene insensitive mutant’s etr1-1, ein4-1, ein2-5, ein3-1, and ein3eil1 display enhanced endurance during freezing conditions [94]. These contrast results indicate that the role of ethylene further needs to be investigated for a better understanding of its role during low-temperature stress.
Brassinosteroids (BRs) are another class of phytohormones that play key role in plant development and defense [193,194,195,196]. Several reports suggest that the BRs are involved during cold stress response [197,198]. Consistently, exogenous application of BRs enhance the chilling tolerance in several plant species [197,198,199,200]. BRs regulate the expression of several COR genes and CBF regulon, thereby control the freezing tolerance [200]. These results clearly indicate that apart from growth regulation, BRs also enhance plant tolerance against freezing stress. Besides, other phytohormones such as Gibberellic acid (GA), Salicylic acid (SA), and Jasmonic acid (JA) also contribute significantly to cold stress response [171]. Additionally, the hormonal crosstalk among phytohormones has created hindrance to decipher the effect of individual phytohormone during the cold stress response. Together, these observations indicate the involvement of phytohormones during cold stress response. However, additional studies are required to identify the molecular components that integrate phytohormones and cold stress response.

8. Epigenetic Regulation

Plants show changes in chromatin modifications such as DNA and histone methylation when exposed to various stresses. Consistently, cold stress alters the expression of several genes due to changes in the chromatin structure [201,202]. Recent reports suggest that epigenetic regulators, such as histone deacetylases, contribute to the transcriptional regulation of COR genes. In particular histone deacetylase (HDACs), the enzyme that catalyzes the removal of an acetyl group from histones, regulate the cold stress response. For example, histone deacetylase HDA6 plays essential role during cold acclimation and freezing tolerance of Arabidopsis [203]. Another deacetylase of Arabidopsis, a WD40-repeat protein high expression of osmotically responsive genes 15 (HOS15) repress cold stress tolerance via histone deacetylation [204,205]. In response to cold stress, HOS15 acts as a positive regulator of cold stress by mediating the degradation of histone deacetylase 2C (HD2C) and changing the chromatin structure from repressive to active state thereby facilitating the expression of CBF mediated COR genes [206]. In maize, global deacetylation of histones H3 and H4 is observed due to cold induced up-regulation of HDACs. Maize HDACs regulate the cold induced DREB1 expression by altering the histone modification and changing chromatin structure at both for gene and specific site [207]. At the same time, active chromatin mark, H3K9ac, is reported be enriched at selective tandem repeats in heterochromatin of maize following cold stress [208]. Recently, Kindren et al. (2018) discovered a lncRNA, SVALKA, in Arabidopsis genome at cold-sensitive region that regulates the freezing tolerance by affecting the CBF1 expression [209]. Although much details are required to fully understand the epigenetic mechanism during cold stress response these finding clearly indicate that epigenetic regulation is indispensable for plant in response to cold stress.

9. Conclusions and Future Perspective

Plants display a plethora of responses to low-temperature stress, such as transcriptomic reprogramming, membrane modification, etc. Low-temperature responsive genes produce protective proteins, such as COR-polypeptides (Cold regulated proteins), antifreeze proteins (AFPs) and some of them even shield lipids. The major determinant for cells to survive during cold stress is their ability to tolerate dehydration and withstand repeated dehydrative/rehydrative cycles. In the plasma membrane, the extent of injury depends on lipid composition and the manifestation of specific cryoprotectants. The production of cryoprotectants is determined by the degree of cold shock and the kind of protection needed for the plant’s survival. Besides, microRNA regulation is also an indispensable response during cold stress. It would be interesting to further elucidate the regulation of miRNA itself by studying their promoter and identifying the overlap in key scis-elements compared to protein-coding genes or discovering different factors that control their expression. With a globally changing climate, the immediate challenge in crop production is to identify and integrate the missing links of the molecular basis of tolerance. The integration of knowledge in the breeding program of susceptible crops is expected to enhance their tolerance to stress and increase crop production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/agronomy12030702/s1, Table S1: Plant microRNAs reported to be regulated by low temperature stress. References [210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, M.A.; data curation, M.A.; writing—original draft preparation, M.A., B.F., M.A.A. and Y.C.; writing—review and editing, M.A. and B.W.; funding acquisition, B.W. and Y.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Major Project of Guangxi grant number Gui Ke 2018-266-Z01 and Guangxi Distinguished Experts Fellowship to YQ. The APC was funded by National Natural Science Foundation of China (31970333).

Acknowledgments

We thank all members of the Qin lab for their assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhu, J.K. Abiotic Stress Signaling and Responses in Plants. Cell 2016, 167, 313–324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Boyer, J.S. Plant productivity and environment. Science 1982, 218, 443–448. [Google Scholar] [CrossRef] [PubMed]
  3. Sanchez, B.; Rasmussen, A.; Porter, J.R. Temperatures and the growth and development of maize and rice: A review. Glob. Chang. Biol. 2014, 20, 408–417. [Google Scholar] [CrossRef] [PubMed]
  4. Zeng, Y.; Zhang, Y.; Xiang, J.; Uphoff, N.T.; Pan, X.; Zhu, D. Effects of Low Temperature Stress on Spikelet-Related Parameters during Anthesis in Indica-Japonica Hybrid Rice. Front. Plant Sci. 2017, 8, 1350. [Google Scholar] [CrossRef] [Green Version]
  5. Moraes de Freitas, G.P.; Basu, S.; Ramegowda, V.; Thomas, J.; Benitez, L.C.; Braga, E.B.; Pereira, A. Physiological and transcriptional responses to low-temperature stress in rice genotypes at the reproductive stage. Plant Signal. Behav. 2019, 14, e1581557. [Google Scholar] [CrossRef] [Green Version]
  6. Albertos, P.; Wagner, K.; Poppenberger, B. Cold stress signalling in female reproductive tissues. Plant Cell Environ. 2019, 42, 846–853. [Google Scholar] [CrossRef]
  7. Thomashow, M.F. Molecular basis of plant cold acclimation: Insights gained from studying the CBF cold response pathway. Plant Physiol. 2010, 154, 571–577. [Google Scholar] [CrossRef] [Green Version]
  8. Uemura, M.; Joseph, R.A.; Steponkus, P.L. Cold Acclimation of Arabidopsis thaliana (Effect on Plasma Membrane Lipid Composition and Freeze-Induced Lesions). Plant Physiol. 1995, 109, 15–30. [Google Scholar] [CrossRef] [Green Version]
  9. Wei, X.; Liu, S.; Sun, C.; Xie, G.; Wang, L. Convergence and Divergence: Signal Perception and Transduction Mechanisms of Cold Stress in Arabidopsis and Rice. Plants 2021, 10, 1864. [Google Scholar] [CrossRef]
  10. Najeeb, S.; Mahender, A.; Anandan, A.; Hussain, W.; Li, Z.; Ali, J. Genetics and Breeding of Low-Temperature Stress Tolerance in Rice. In Rice Improvement: Physiological, Molecular Breeding and Genetic Perspectives; Ali, J., Wani, S.H., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 221–280. [Google Scholar]
  11. Hincha, D.K.; Zuther, E. Introduction: Plant Cold Acclimation and Winter Survival. Methods Mol. Biol. 2020, 2156, 1–7. [Google Scholar] [CrossRef]
  12. Sakai, A.; Larcher, W. Cold Acclimation in Plants. In Frost Survival of Plants: Responses and Adaptation to Freezing Stress; Sakai, A., Larcher, W., Eds.; Springer Berlin Heidelberg: Berlin/Heidelberg, Germany, 1987; pp. 97–137. [Google Scholar]
  13. Arakawa, K.; Kasuga, J.; Takata, N. Mechanism of Overwintering in Trees. In Survival Strategies in Extreme Cold and Desiccation: Adaptation Mechanisms and Their Applications; Iwaya-Inoue, M., Sakurai, M., Uemura, M., Eds.; Springer: Singapore, 2018; pp. 129–147. [Google Scholar]
  14. Kalcsits, L.A.; Silim, S.; Tanino, K. Warm temperature accelerates short photoperiod-induced growth cessation and dormancy induction in hybrid poplar (Populus × spp.). Trees 2009, 23, 971–979. [Google Scholar] [CrossRef]
  15. Li, C.; Junttila, O.; Ernstsen, A.; Heino, P.; Palva, E.T. Photoperiodic control of growth, cold acclimation and dormancy development in silver birch (Betula pendula) ecotypes. Physiol. Plant 2003, 117, 206–212. [Google Scholar] [CrossRef]
  16. Maul, P.; McCollum, G.T.; Popp, M.; Guy, C.L.; Porat, R. Transcriptome profiling of grapefruit flavedo following exposure to low temperature and conditioning treatments uncovers principal molecular components involved in chilling tolerance and susceptibility. Plant Cell Environ. 2008, 31, 752–768. [Google Scholar] [CrossRef] [PubMed]
  17. Guy, C.L.; Niemi, K.J.; Brambl, R. Altered gene expression during cold acclimation of spinach. Proc. Natl. Acad. Sci. USA 1985, 82, 3673–3677. [Google Scholar] [CrossRef] [Green Version]
  18. Chrispeels, M.J.; Crawford, N.M.; Schroeder, J.I. Proteins for transport of water and mineral nutrients across the membranes of plant cells. Plant Cell 1999, 11, 661–676. [Google Scholar] [CrossRef] [Green Version]
  19. Lalonde, S.; Frommer, W.B. SUT Sucrose and MST Monosaccharide Transporter Inventory of the Selaginella Genome. Front. Plant Sci. 2012, 3, 24. [Google Scholar] [CrossRef] [Green Version]
  20. Seidel, T.; Siek, M.; Marg, B.; Dietz, K.J. Energization of vacuolar transport in plant cells and its significance under stress. Int. Rev. Cell Mol. Biol. 2013, 304, 57–131. [Google Scholar] [CrossRef]
  21. Sze, H.; Li, X.; Palmgren, M.G. Energization of plant cell membranes by H+-pumping ATPases. Regulation and biosynthesis. Plant Cell 1999, 11, 677–690. [Google Scholar]
  22. Zheng, G.; Li, L.; Li, W. Glycerolipidome responses to freezing- and chilling-induced injuries: Examples in Arabidopsis and rice. BMC Plant Biol. 2016, 16, 70. [Google Scholar] [CrossRef] [Green Version]
  23. Iba, K. Acclimative response to temperature stress in higher plants: Approaches of gene engineering for temperature tolerance. Annu. Rev. Plant Biol. 2002, 53, 225–245. [Google Scholar] [CrossRef] [Green Version]
  24. Kenrick, J.R.; Bishop, D.G. The fatty acid composition of phosphatidylglycerol and sulfoquinovosyldiacylglycerol of higher plants in relation to chilling sensitivity. Plant Physiol. 1986, 81, 946–949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Kaniuga, Z.; Sączyńska, V.; Miśkiewicz, E.; Garstka, M. The fatty acid composition of phosphatidylglycerol and sulfoquinovosyldiacylglycerol of Zea mays genotypes differing in chilling susceptibility. J. Plant Physiol. 1999, 154, 256–263. [Google Scholar] [CrossRef]
  26. Murata, N.; Sato, N.; Takahashi, N.; Hamazaki, Y. Compositions and positional distributions of fatty acids in phospholipids from leaves of chilling-sensitive and chilling-resistant plants. Plant Cell Physiol. 1982, 23, 1071–1079. [Google Scholar]
  27. Norman, H.A.; McMillan, C.; Thompson, G.A., Jr. Phosphatidylglycerol molecular species in chilling-sensitive and chilling-resistant populations of Avicennia germinans (L.) L. Plant Cell Physiol. 1984, 25, 1437–1444. [Google Scholar] [CrossRef]
  28. Roughan, P.G. Phosphatidylglycerol and chilling sensitivity in plants. Plant Physiol. 1985, 77, 740–746. [Google Scholar] [CrossRef] [PubMed]
  29. Murata, N. Molecular species composition of phosphatidylglycerols from chilling-sensitive and chilling-resistant plants. Plant Cell Physiol. 1983, 24, 81–86. [Google Scholar] [CrossRef]
  30. Murata, N.; Ishizaki-Nishizawa, O.; Higashi, S.; Hayashi, H.; Tasaka, Y.; Nishida, I. Genetically engineered alteration in the chilling sensitivity of plants. Nature 1992, 356, 710–713. [Google Scholar] [CrossRef]
  31. Lopez-Marques, R.L.; Theorin, L.; Palmgren, M.G.; Pomorski, T.G. P4-ATPases: Lipid flippases in cell membranes. Pflug. Archiv. 2014, 466, 1227–1240. [Google Scholar] [CrossRef] [Green Version]
  32. Murai, M.; Yoshida, S. Vacuolar membrane lesions induced by a freeze-thaw cycle in protoplasts isolated from deacclimated tubers of Jerusalem artichoke (Helianthus tuberosus L.). Plant Cell Physiol. 1998, 39, 87–96. [Google Scholar] [CrossRef] [Green Version]
  33. Zheng, G.; Tian, B.; Li, W. Membrane lipid remodelling of Meconopsis racemosa after its introduction into lowlands from an alpine environment. PLoS ONE 2014, 9, e106614. [Google Scholar] [CrossRef]
  34. Ohnishi, M.; Kadohama, N.; Suzuki, Y.; Kajiyama, T.; Shichijo, C.; Ishizaki, K.; Fukaki, H.; Iida, H.; Kambara, H.; Mimura, T. Involvement of Ca2+ in Vacuole Degradation Caused by a Rapid Temperature Decrease in Saintpaulia Palisade Cells: A Case of Gene Expression Analysis in a Specialized Small Tissue. Plant Cell Physiol. 2015, 56, 1297–1305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Neuhaus, H.E. Transport of primary metabolites across the plant vacuolar membrane. FEBS Lett. 2007, 581, 2223–2226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Lucinski, R.; Jackowski, G. AtFtsH heterocomplex-mediated degradation of apoproteins of the major light harvesting complex of photosystem II (LHCII) in response to stresses. J. Plant Physiol. 2013, 170, 1082–1089. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, M.; Thelen, J.J. Acyl-lipid desaturase 1 primes cold acclimation response in Arabidopsis. Physiol. Plant 2016, 158, 11–22. [Google Scholar] [CrossRef]
  38. Mazur, R.; Trzcinska-Danielewicz, J.; Kozlowski, P.; Kowalewska, L.; Rumak, I.; Shiell, B.J.; Mostowska, A.; Michalski, W.P.; Garstka, M. Dark-chilling and subsequent photo-activation modulate expression and induce reversible association of chloroplast lipoxygenase with thylakoid membrane in runner bean (Phaseolus coccineus L.). Plant Physiol Biochem. 2018, 122, 102–112. [Google Scholar] [CrossRef]
  39. Rogalski, M.; Schottler, M.A.; Thiele, W.; Schulze, W.X.; Bock, R. Rpl33, a nonessential plastid-encoded ribosomal protein in tobacco, is required under cold stress conditions. Plant Cell 2008, 20, 2221–2237. [Google Scholar] [CrossRef] [Green Version]
  40. PokornÁ, J.; SchwarzerovÁ, K.; ZelenkovÁ, S.; PetrÁŠEk, J.; JanotovÁ, I.; ČApkovÁ, V.; OpatrnÝ, Z. Sites of actin filament initiation and reorganization in cold-treated tobacco cells. Plant Cell Environ. 2004, 27, 641–653. [Google Scholar] [CrossRef]
  41. Lazareva, E.M.; Chentsov Iu, S.; Smirnova, E.A. The effect of low temperature on the microtubules in root meristem cells of spring and winter cultivars of wheat Triticum aestivum L. Tsitologiia 2008, 50, 597–612. [Google Scholar] [CrossRef]
  42. Guo, X.; Liu, D.; Chong, K. Cold signaling in plants: Insights into mechanisms and regulation. J. Integr Plant Biol. 2018, 60, 745–756. [Google Scholar] [CrossRef] [Green Version]
  43. Wang, L.; Sadeghnezhad, E.; Nick, P. Upstream of gene expression: What is the role of microtubules in cold signalling? J. Exp. Bot. 2020, 71, 36–48. [Google Scholar] [CrossRef]
  44. Dodd, A.N.; Kudla, J.; Sanders, D. The language of calcium signaling. Annu Rev. Plant Biol. 2010, 61, 593–620. [Google Scholar] [CrossRef] [PubMed]
  45. Tang, R.J.; Luan, S. Regulation of calcium and magnesium homeostasis in plants: From transporters to signaling network. Curr Opin Plant Biol. 2017, 39, 97–105. [Google Scholar] [CrossRef] [PubMed]
  46. Tuteja, N.; Mahajan, S. Calcium signaling network in plants: An overview. Plant Signal. Behav. 2007, 2, 79–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Monshausen, G.B. Visualizing Ca(2+) signatures in plants. Curr Opin Plant Biol. 2012, 15, 677–682. [Google Scholar] [CrossRef] [PubMed]
  48. Sanders, D.; Pelloux, J.; Brownlee, C.; Harper, J.F. Calcium at the crossroads of signaling. Plant Cell 2002, 14, S401–S417. [Google Scholar] [CrossRef] [Green Version]
  49. Kawashima, T.; Lorkovic, Z.J.; Nishihama, R.; Ishizaki, K.; Axelsson, E.; Yelagandula, R.; Kohchi, T.; Berger, F. Diversification of histone H2A variants during plant evolution. Trends Plant Sci. 2015, 20, 419–425. [Google Scholar] [CrossRef]
  50. Finka, A.; Cuendet, A.F.; Maathuis, F.J.; Saidi, Y.; Goloubinoff, P. Plasma membrane cyclic nucleotide gated calcium channels control land plant thermal sensing and acquired thermotolerance. Plant Cell 2012, 24, 3333–3348. [Google Scholar] [CrossRef] [Green Version]
  51. Nawaz, Z.; Kakar, K.U.; Saand, M.A.; Shu, Q.Y. Cyclic nucleotide-gated ion channel gene family in rice, identification, characterization and experimental analysis of expression response to plant hormones, biotic and abiotic stresses. BMC Genom. 2014, 15, 853. [Google Scholar] [CrossRef] [Green Version]
  52. Ma, Y.; Dai, X.; Xu, Y.; Luo, W.; Zheng, X.; Zeng, D.; Pan, Y.; Lin, X.; Liu, H.; Zhang, D.; et al. COLD1 confers chilling tolerance in rice. Cell 2015, 160, 1209–1221. [Google Scholar] [CrossRef] [Green Version]
  53. Hong-Bo, S.; Li-Ye, C.; Ming-An, S.; Shi-Qing, L.; Ji-Cheng, Y. Bioengineering plant resistance to abiotic stresses by the global calcium signal system. Biotechnol Adv. 2008, 26, 503–510. [Google Scholar] [CrossRef]
  54. Mori, K.; Renhu, N.; Naito, M.; Nakamura, A.; Shiba, H.; Yamamoto, T.; Suzaki, T.; Iida, H.; Miura, K. Ca(2+)-permeable mechanosensitive channels MCA1 and MCA2 mediate cold-induced cytosolic Ca(2+) increase and cold tolerance in Arabidopsis. Sci Rep. 2018, 8, 550. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, Q.; Ding, Y.; Shi, Y.; Ma, L.; Wang, Y.; Song, C.; Wilkins, K.A.; Davies, J.M.; Knight, H.; Knight, M.R.; et al. The calcium transporter ANNEXIN1 mediates cold-induced calcium signaling and freezing tolerance in plants. EMBO J. 2021, 40, e104559. [Google Scholar] [CrossRef] [PubMed]
  56. Ranty, B.; Aldon, D.; Cotelle, V.; Galaud, J.P.; Thuleau, P.; Mazars, C. Calcium Sensors as Key Hubs in Plant Responses to Biotic and Abiotic Stresses. Front. Plant Sci. 2016, 7, 327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Gifford, J.L.; Jamshidiha, M.; Mo, J.; Ishida, H.; Vogel, H.J. Comparing the calcium binding abilities of two soybean calmodulins: Towards understanding the divergent nature of plant calmodulins. Plant Cell 2013, 25, 4512–4524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Abbas, N.; Maurya, J.P.; Senapati, D.; Gangappa, S.N.; Chattopadhyay, S. Arabidopsis CAM7 and HY5 physically interact and directly bind to the HY5 promoter to regulate its expression and thereby promote photomorphogenesis. Plant Cell 2014, 26, 1036–1052. [Google Scholar] [CrossRef] [Green Version]
  59. McCormack, E.; Tsai, Y.C.; Braam, J. Handling calcium signaling: Arabidopsis CaMs and CMLs. Trends Plant Sci. 2005, 10, 383–389. [Google Scholar] [CrossRef]
  60. Zhu, X.; Perez, M.; Aldon, D.; Galaud, J.P. Respective contribution of CML8 and CML9, two arabidopsis calmodulin-like proteins, to plant stress responses. Plant Signal. Behav. 2017, 12, e1322246. [Google Scholar] [CrossRef]
  61. Zhu, X.; Robe, E.; Jomat, L.; Aldon, D.; Mazars, C.; Galaud, J.P. CML8, an Arabidopsis Calmodulin-Like Protein, Plays a Role in Pseudomonas syringae Plant Immunity. Plant Cell Physiol. 2017, 58, 307–319. [Google Scholar] [CrossRef] [Green Version]
  62. Bouche, N.; Scharlat, A.; Snedden, W.; Bouchez, D.; Fromm, H. A novel family of calmodulin-binding transcription activators in multicellular organisms. J. Biol. Chem. 2002, 277, 21851–21861. [Google Scholar] [CrossRef] [Green Version]
  63. Doherty, C.J.; Van Buskirk, H.A.; Myers, S.J.; Thomashow, M.F. Roles for Arabidopsis CAMTA transcription factors in cold-regulated gene expression and freezing tolerance. Plant Cell 2009, 21, 972–984. [Google Scholar] [CrossRef] [Green Version]
  64. Kim, Y.; Park, S.; Gilmour, S.J.; Thomashow, M.F. Roles of CAMTA transcription factors and salicylic acid in configuring the low-temperature transcriptome and freezing tolerance of Arabidopsis. Plant J. 2013, 75, 364–376. [Google Scholar] [CrossRef] [PubMed]
  65. Kim, S.M.; Suh, J.P.; Lee, C.K.; Lee, J.H.; Kim, Y.G.; Jena, K.K. QTL mapping and development of candidate gene-derived DNA markers associated with seedling cold tolerance in rice (Oryza sativa L.). Mol. Genet. Genom. 2014, 289, 333–343. [Google Scholar] [CrossRef] [PubMed]
  66. Shen, C.; Yang, Y.; Du, L.; Wang, H. Calmodulin-binding transcription activators and perspectives for applications in biotechnology. Appl. Microbiol Biotechnol. 2015, 99, 10379–10385. [Google Scholar] [CrossRef] [PubMed]
  67. Kim, Y.S.; An, C.; Park, S.; Gilmour, S.J.; Wang, L.; Renna, L.; Brandizzi, F.; Grumet, R.; Thomashow, M.F. CAMTA-Mediated Regulation of Salicylic Acid Immunity Pathway Genes in Arabidopsis Exposed to Low Temperature and Pathogen Infection. Plant Cell 2017, 29, 2465–2477. [Google Scholar] [CrossRef] [Green Version]
  68. Batistic, O.; Kudla, J. Integration and channeling of calcium signaling through the CBL calcium sensor/CIPK protein kinase network. Planta 2004, 219, 915–924. [Google Scholar] [CrossRef]
  69. Kolukisaoglu, U.; Weinl, S.; Blazevic, D.; Batistic, O.; Kudla, J. Calcium sensors and their interacting protein kinases: Genomics of the Arabidopsis and rice CBL-CIPK signaling networks. Plant Physiol. 2004, 134, 43–58. [Google Scholar] [CrossRef] [Green Version]
  70. Huang, C.; Ding, S.; Zhang, H.; Du, H.; An, L. CIPK7 is involved in cold response by interacting with CBL1 in Arabidopsis thaliana. Plant Sci. 2011, 181, 57–64. [Google Scholar] [CrossRef]
  71. Mo, C.; Wan, S.; Xia, Y.; Ren, N.; Zhou, Y.; Jiang, X. Expression Patterns and Identified Protein-Protein Interactions Suggest That Cassava CBL-CIPK Signal Networks Function in Responses to Abiotic Stresses. Front. Plant Sci. 2018, 9, 269. [Google Scholar] [CrossRef] [Green Version]
  72. Xi, Y.; Liu, J.; Dong, C.; Cheng, Z.M. The CBL and CIPK Gene Family in Grapevine (Vitis vinifera): Genome-Wide Analysis and Expression Profiles in Response to Various Abiotic Stresses. Front. Plant Sci. 2017, 8, 978. [Google Scholar] [CrossRef] [Green Version]
  73. Mahajan, S.; Sopory, S.K.; Tuteja, N. Cloning and characterization of CBL-CIPK signalling components from a legume (Pisum sativum). FEBS J. 2006, 273, 907–925. [Google Scholar] [CrossRef] [Green Version]
  74. Zhang, H.; Yang, B.; Liu, W.Z.; Li, H.; Wang, L.; Wang, B.; Deng, M.; Liang, W.; Deyholos, M.K.; Jiang, Y.Q. Identification and characterization of CBL and CIPK gene families in canola (Brassica napus L.). BMC Plant Biol. 2014, 14, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Xiang, Y.; Huang, Y.; Xiong, L. Characterization of stress-responsive CIPK genes in rice for stress tolerance improvement. Plant Physiol. 2007, 144, 1416–1428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Aslam, M.; Fakher, B.; Anandhan, S.; Pande, V.; Ahmed, Z.; Qin, Y. Ectopic Expression of Cold Responsive LlaCIPK Gene Enhances Cold Stress Tolerance in Nicotiana tabacum. Genes 2019, 10, 446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Aslam, M.; Sinha, V.B.; Singh, R.K.; Anandhan, S.; Pande, V.; Ahmed, Z. Isolation of cold stress-responsive genes from Lepidium latifolium by suppressive subtraction hybridization. Acta Physiol. Plant 2009, 32, 205–210. [Google Scholar] [CrossRef]
  78. Jin, J.; Tian, F.; Yang, D.C.; Meng, Y.Q.; Kong, L.; Luo, J.; Gao, G. PlantTFDB 4.0: Toward a central hub for transcription factors and regulatory interactions in plants. Nucleic Acids Res. 2017, 45, D1040–D1045. [Google Scholar] [CrossRef] [Green Version]
  79. Salih, H.; Gong, W.; He, S.; Sun, G.; Sun, J.; Du, X. Genome-wide characterization and expression analysis of MYB transcription factors in Gossypium hirsutum. BMC Genet. 2016, 17, 129. [Google Scholar] [CrossRef] [Green Version]
  80. Seki, M.; Narusaka, M.; Abe, H.; Kasuga, M.; Yamaguchi-Shinozaki, K.; Carninci, P.; Hayashizaki, Y.; Shinozaki, K. Monitoring the expression pattern of 1300 Arabidopsis genes under drought and cold stresses by using a full-length cDNA microarray. Plant Cell 2001, 13, 61–72. [Google Scholar] [CrossRef] [Green Version]
  81. Carretero-Paulet, L.; Galstyan, A.; Roig-Villanova, I.; Martinez-Garcia, J.F.; Bilbao-Castro, J.R.; Robertson, D.L. Genome-wide classification and evolutionary analysis of the bHLH family of transcription factors in Arabidopsis, poplar, rice, moss, and algae. Plant Physiol. 2010, 153, 1398–1412. [Google Scholar] [CrossRef] [Green Version]
  82. Aslam, M.; Grover, A.; Sinha, V.B.; Fakher, B.; Pande, V.; Yadav, P.V.; Gupta, S.M.; Anandhan, S.; Ahmed, Z. Isolation and characterization of cold responsive NAC gene from Lepidium latifolium. Mol. Biol. Rep. 2012, 39, 9629–9638. [Google Scholar] [CrossRef]
  83. An, J.P.; Li, R.; Qu, F.J.; You, C.X.; Wang, X.F.; Hao, Y.J. An apple NAC transcription factor negatively regulates cold tolerance via CBF-dependent pathway. J. Plant Physiol. 2018, 221, 74–80. [Google Scholar] [CrossRef]
  84. Agarwal, M.; Hao, Y.; Kapoor, A.; Dong, C.H.; Fujii, H.; Zheng, X.; Zhu, J.K. A R2R3 type MYB transcription factor is involved in the cold regulation of CBF genes and in acquired freezing tolerance. J. Biol Chem. 2006, 281, 37636–37645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. 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] [Green Version]
  86. Knight, M.R.; Knight, H. Low-temperature perception leading to gene expression and cold tolerance in higher plants. New Phytol. 2012, 195, 737–751. [Google Scholar] [CrossRef]
  87. Gilmour, S.J.; Fowler, S.G.; Thomashow, M.F. Arabidopsis transcriptional activators CBF1, CBF2, and CBF3 have matching functional activities. Plant Mol. Biol. 2004, 54, 767–781. [Google Scholar] [CrossRef]
  88. Li, S.; Liu, X.; Wang, S.; Hao, D.; Xi, J. Proteomics dissection of cold responsive proteins based on PEG fractionation in Arabidopsis. Chem. Res. Chin. Univ. 2014, 30, 272–278. [Google Scholar] [CrossRef]
  89. Jaglo-Ottosen, K.R.; Gilmour, S.J.; Zarka, D.G.; Schabenberger, O.; Thomashow, M.F. Arabidopsis CBF1 overexpression induces COR genes and enhances freezing tolerance. Science 1998, 280, 104–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Liu, Q.; Kasuga, M.; Sakuma, Y.; Abe, H.; Miura, S.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Two transcription factors, DREB1 and DREB2, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought-and low-temperature-responsive gene expression, respectively, in Arabidopsis. Plant Cell 1998, 10, 1391–1406. [Google Scholar] [CrossRef] [Green Version]
  91. Chinnusamy, V.; Ohta, M.; Kanrar, S.; Lee, B.H.; Hong, X.; Agarwal, M.; Zhu, J.K. ICE1: A regulator of cold-induced transcriptome and freezing tolerance in Arabidopsis. Genes Dev. 2003, 17, 1043–1054. [Google Scholar] [CrossRef] [Green Version]
  92. Kim, Y.S.; Lee, M.; Lee, J.-H.; Lee, H.-J.; Park, C.-M. The unified ICE–CBF pathway provides a transcriptional feedback control of freezing tolerance during cold acclimation in Arabidopsis. Plant Mol. Biol. 2015, 89, 187–201. [Google Scholar] [CrossRef]
  93. Maruyama, K.; Takeda, M.; Kidokoro, S.; Yamada, K.; Sakuma, Y.; Urano, K.; Fujita, M.; Yoshiwara, K.; Matsukura, S.; Morishita, Y. Metabolic pathways involved in cold acclimation identified by integrated analysis of metabolites and transcripts regulated by DREB1A and DREB2A. Plant Physiol. 2009, 150, 1972–1980. [Google Scholar] [CrossRef] [Green Version]
  94. Shi, Y.; Tian, S.; Hou, L.; Huang, X.; Zhang, X.; Guo, H.; Yang, S. 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] [Green Version]
  95. Ishitani, M.; Xiong, L.; Stevenson, B.; Zhu, J.-K. Genetic analysis of osmotic and cold stress signal transduction in Arabidopsis: Interactions and convergence of abscisic acid-dependent and abscisic acid-independent pathways. Plant Cell 1997, 9, 1935–1949. [Google Scholar] [PubMed] [Green Version]
  96. Lee, B.H.; Henderson, D.A.; Zhu, J.-K. The Arabidopsis cold-responsive transcriptome and its regulation by ICE1. Plant Cell 2005, 17, 3155–3175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Park, S.; Lee, C.M.; Doherty, C.J.; Gilmour, S.J.; Kim, Y.; Thomashow, M.F. Regulation of the Arabidopsis CBF regulon by a complex low-temperature regulatory network. Plant J. 2015, 82, 193–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. 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] [Green Version]
  99. Jia, Y.; Ding, Y.; Shi, Y.; Zhang, X.; Gong, Z.; Yang, S. The cbfs triple mutants reveal the essential functions of CBFs in cold acclimation and allow the definition of CBF regulons in Arabidopsis. New Phytol. 2016, 212, 345–353. [Google Scholar] [CrossRef] [Green Version]
  100. Tian, J.; Song, Y.; Du, Q.; Yang, X.; Ci, D.; Chen, J.; Xie, J.; Li, B.; Zhang, D. Population genomic analysis of gibberellin-responsive long non-coding RNAs in Populus. J. Exp. Bot. 2016, 67, 2467–2482. [Google Scholar] [CrossRef] [Green Version]
  101. Ito, Y.; Katsura, K.; Maruyama, K.; Taji, T.; Kobayashi, M.; Seki, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Functional analysis of rice DREB1/CBF-type transcription factors involved in cold-responsive gene expression in transgenic rice. Plant Cell Physiol. 2006, 47, 141–153. [Google Scholar] [CrossRef] [Green Version]
  102. Dong, C.H.; Agarwal, M.; Zhang, Y.; Xie, Q.; Zhu, J.K. The negative regulator of plant cold responses, HOS1, is a RING E3 ligase that mediates the ubiquitination and degradation of ICE1. Proc. Natl. Acad. Sci. USA 2006, 103, 8281–8286. [Google Scholar] [CrossRef] [Green Version]
  103. Miura, K.; Jin, J.B.; Lee, J.; Yoo, C.Y.; Stirm, V.; Miura, T.; Ashworth, E.N.; Bressan, R.A.; Yun, D.J.; Hasegawa, P.M. SIZ1-mediated sumoylation of ICE1 controls CBF3/DREB1A expression and freezing tolerance in Arabidopsis. Plant Cell 2007, 19, 1403–1414. [Google Scholar] [CrossRef] [Green Version]
  104. Ding, Y.; Li, H.; Zhang, X.; Xie, Q.; Gong, Z.; Yang, S. OST1 kinase modulates freezing tolerance by enhancing ICE1 stability in Arabidopsis. Dev. Cell 2015, 32, 278–289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Jung, J.H.; Park, C.M. HOS1-mediated activation of FLC via chromatin remodeling under cold stress. Plant Signal. Behav. 2013, 8, e27342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Miura, K.; Ohta, M.; Nakazawa, M.; Ono, M.; Hasegawa, P.M. ICE1 Ser403 is necessary for protein stabilization and regulation of cold signaling and tolerance. Plant J. 2011, 67, 269–279. [Google Scholar] [CrossRef] [Green Version]
  107. Agarwal, P.K.; Agarwal, P.; Reddy, M.K.; Sopory, S.K. Role of DREB transcription factors in abiotic and biotic stress tolerance in plants. Plant Cell Rep. 2006, 25, 1263–1274. [Google Scholar] [CrossRef]
  108. 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] [Green Version]
  109. Vogel, J.T.; Zarka, D.G.; Van Buskirk, H.A.; 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]
  110. Kidokoro, S.; Kim, J.-S.; Ishikawa, T.; Suzuki, T.; Shinozaki, K.; Yamaguchi-Shinozaki, K. DREB1A/CBF3 is repressed by transgene-induced DNA methylation in the Arabidopsis ice1-1 mutant. Plant Cell 2020, 32, 1035–1048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Thomashow, M.; Torii, K.U. SCREAMing Twist on the Role of ICE1 in Freezing Tolerance. Plant Cell 2020, 32, 816–819. [Google Scholar] [CrossRef] [Green Version]
  112. Rodriguez, M.C.; Petersen, M.; Mundy, J. Mitogen-activated protein kinase signaling in plants. Annu. Rev. Plant Biol. 2010, 61, 621–649. [Google Scholar] [CrossRef]
  113. Li, H.; Ding, Y.; Shi, Y.; Zhang, X.; Zhang, S.; Gong, Z.; Yang, S. MPK3- and MPK6-Mediated ICE1 Phosphorylation Negatively Regulates ICE1 Stability and Freezing Tolerance in Arabidopsis. Dev. Cell 2017, 43, 630–642.e634. [Google Scholar] [CrossRef]
  114. Zhao, C.; Wang, P.; Si, T.; Hsu, C.C.; Wang, L.; Zayed, O.; Yu, Z.; Zhu, Y.; Dong, J.; Tao, W.A.; et al. MAP Kinase Cascades Regulate the Cold Response by Modulating ICE1 Protein Stability. Dev. Cell 2017, 43, 618–629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Voinnet, O. Origin, biogenesis, and activity of plant microRNAs. Cell 2009, 136, 669–687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Zhang, L.; Yao, L.; Zhang, N.; Yang, J.; Zhu, X.; Tang, X.; Calderon-Urrea, A.; Si, H. Lateral Root Development in Potato Is Mediated by Stu-mi164 Regulation of NAC Transcription Factor. Front. Plant Sci. 2018, 9, 383. [Google Scholar] [CrossRef] [PubMed]
  117. Song, G.; Zhang, R.; Zhang, S.; Li, Y.; Gao, J.; Han, X.; Chen, M.; Wang, J.; Li, W.; Li, G. Response of microRNAs to cold treatment in the young spikes of common wheat. BMC Genom. 2017, 18, 212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Bai, B.; Bian, H.; Zeng, Z.; Hou, N.; Shi, B.; Wang, J.; Zhu, M.; Han, N. miR393-Mediated Auxin Signaling Regulation is Involved in Root Elongation Inhibition in Response to Toxic Aluminum Stress in Barley. Plant Cell Physiol. 2017, 58, 426–439. [Google Scholar] [CrossRef] [PubMed]
  119. Sunkar, R.; Zhu, J.K. Novel and stress-regulated microRNAs and other small RNAs from Arabidopsis. Plant Cell 2004, 16, 2001–2019. [Google Scholar] [CrossRef] [Green Version]
  120. Kumar, R. Role of microRNAs in biotic and abiotic stress responses in crop plants. Appl. Biochem. Biotechnol. 2014, 174, 93–115. [Google Scholar] [CrossRef]
  121. Karimi, M.; Ghazanfari, F.; Fadaei, A.; Ahmadi, L.; Shiran, B.; Rabei, M.; Fallahi, H. The Small-RNA Profiles of Almond (Prunus dulcis Mill.) Reproductive Tissues in Response to Cold Stress. PLoS ONE 2016, 11, e0156519. [Google Scholar] [CrossRef]
  122. Megha, S.; Basu, U.; Kav, N.N.V. Regulation of low temperature stress in plants by microRNAs. Plant Cell Environ. 2018, 41, 1–15. [Google Scholar] [CrossRef]
  123. Bej, S.; Basak, J. MicroRNAs: The Potential Biomarkers in Plant Stress Response. Am. J. Plant Sci. 2014, 5, 43925. [Google Scholar] [CrossRef] [Green Version]
  124. Sunkar, R.; Li, Y.F.; Jagadeeswaran, G. Functions of microRNAs in plant stress responses. Trends Plant Sci. 2012, 17, 196–203. [Google Scholar] [CrossRef] [PubMed]
  125. Liu, Y.; Wang, K.; Li, D.; Yan, J.; Zhang, W. Enhanced Cold Tolerance and Tillering in Switchgrass (Panicum virgatum L.) by Heterologous Expression of Osa-miR393a. Plant Cell Physiol. 2017, 58, 2226–2240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Zeng, X.; Xu, Y.; Jiang, J.; Zhang, F.; Ma, L.; Wu, D.; Wang, Y.; Sun, W. Identification of cold stress responsive microRNAs in two winter turnip rape (Brassica rapa L.) by high throughput sequencing. BMC Plant Biol. 2018, 18, 52. [Google Scholar] [CrossRef] [PubMed]
  127. Yang, Y.; Zhang, X.; Su, Y.; Zou, J.; Wang, Z.; Xu, L.; Que, Y. miRNA alteration is an important mechanism in sugarcane response to low-temperature environment. BMC Genom. 2017, 18, 833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Wang, R.; Zhang, Y.; Kieffer, M.; Yu, H.; Kepinski, S.; Estelle, M. HSP90 regulates temperature-dependent seedling growth in Arabidopsis by stabilizing the auxin co-receptor F-box protein TIR1. Nat. Commun. 2016, 7, 10269. [Google Scholar] [CrossRef]
  129. Aslam, M.; Sugita, K.; Qin, Y.; Rahman, A. Aux/IAA14 Regulates microRNA-Mediated Cold Stress Response in Arabidopsis Roots. Int. J. Mol. Sci. 2020, 21, 8441. [Google Scholar] [CrossRef]
  130. Fischer, B.B.; Eggen, R.I.; Niyogi, K.K. Characterization of singlet oxygen-accumulating mutants isolated in a screen for altered oxidative stress response in Chlamydomonas reinhardtii. BMC Plant Biol. 2010, 10, 279. [Google Scholar] [CrossRef] [Green Version]
  131. Asada, K.; Kiso, K.; Yoshikawa, K. Univalent reduction of molecular oxygen by spinach chloroplasts on illumination. J. Biol. Chem. 1974, 249, 2175–2181. [Google Scholar] [CrossRef]
  132. Alves, F.R.R.; de Melo, H.C.; Crispim-Filho, A.J.; Costa, A.C.; Nascimento, K.J.T.; Carvalho, R.F. Physiological and biochemical responses of photomorphogenic tomato mutants (cv. Micro-Tom) under water withholding. Acta Physiol. Plant 2016, 38, 155. [Google Scholar] [CrossRef] [Green Version]
  133. Stockwell, B.R.; Friedmann Angeli, J.P.; Bayir, H.; Bush, A.I.; Conrad, M.; Dixon, S.J.; Fulda, S.; Gascon, S.; Hatzios, S.K.; Kagan, V.E.; et al. Ferroptosis: A Regulated Cell Death Nexus Linking Metabolism, Redox Biology, and Disease. Cell 2017, 171, 273–285. [Google Scholar] [CrossRef] [Green Version]
  134. Demidchik, V. Mechanisms of oxidative stress in plants: From classical chemistry to cell biology. Environ. Exp. Bot. 2015, 109, 212–228. [Google Scholar] [CrossRef]
  135. Pinhero, R.G.; Rao, M.V.; Paliyath, G.; Murr, D.P.; Fletcher, R.A. Changes in Activities of Antioxidant Enzymes and Their Relationship to Genetic and Paclobutrazol-Induced Chilling Tolerance of Maize Seedlings. Plant Physiol. 1997, 114, 695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Fukushima, A.; Iwasa, M.; Nakabayashi, R.; Kobayashi, M.; Nishizawa, T.; Okazaki, Y.; Saito, K.; Kusano, M. Effects of Combined Low Glutathione with Mild Oxidative and Low Phosphorus Stress on the Metabolism of Arabidopsis thaliana. Front. Plant Sci. 2017, 8, 1464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 2002, 7, 405–410. [Google Scholar] [CrossRef]
  138. Ramel, F.; Birtic, S.; Ginies, C.; Soubigou-Taconnat, L.; Triantaphylides, C.; Havaux, M. Carotenoid oxidation products are stress signals that mediate gene responses to singlet oxygen in plants. Proc. Natl. Acad. Sci. USA 2012, 109, 5535–5540. [Google Scholar] [CrossRef] [Green Version]
  139. Boscaiu, M.; Sánchez, M.; Bautista, I.; Donat, P.; Lidón, A.; Llinares, J.; Llul, C.; Mayoral, O.; Vicente, O. Phenolic Compounds as Stress Markers in Plants from Gypsum Habitats; University of Agricultural Sciences and Veterinary Medicine: Cluj-Napoca, Romania, 2010; Volume 67, pp. 1843–5394. [Google Scholar]
  140. Asada, K. Production and Scavenging of Reactive Oxygen Species in Chloroplasts and Their Functions. Plant Physiol. 2006, 141, 391. [Google Scholar] [CrossRef] [Green Version]
  141. Prasad, T.K.; Anderson, M.D.; Martin, B.A.; Stewart, C.R. Evidence for Chilling-Induced Oxidative Stress in Maize Seedlings and a Regulatory Role for Hydrogen Peroxide. Plant Cell 1994, 6, 65–74. [Google Scholar] [CrossRef]
  142. Prasad, T.K.; Anderson, M.D.; Stewart, C.R. Acclimation, Hydrogen Peroxide, and Abscisic Acid Protect Mitochondria against Irreversible Chilling Injury in Maize Seedlings. Plant Physiol. 1994, 105, 619–627. [Google Scholar] [CrossRef] [Green Version]
  143. Davletova, S.; Rizhsky, L.; Liang, H.; Shengqiang, Z.; Oliver, D.J.; Coutu, J.; Shulaev, V.; Schlauch, K.; Mittler, R. Cytosolic Ascorbate Peroxidase 1 Is a Central Component of the Reactive Oxygen Gene Network of Arabidopsis. Plant Cell 2005, 17, 268–281. [Google Scholar] [CrossRef] [Green Version]
  144. Kawakami, S.; Matsumoto, Y.; Matsunaga, A.; Mayama, S.; Mizuno, M. Molecular cloning of ascorbate peroxidase in potato tubers and its response during storage at low temperature. Plant Sci. 2002, 163, 829–836. [Google Scholar] [CrossRef]
  145. Sato, Y.; Masuta, Y.; Saito, K.; Murayama, S.; Ozawa, K. Enhanced chilling tolerance at the booting stage in rice by transgenic overexpression of the ascorbate peroxidase gene, OsAPXa. Plant Cell Rep. 2011, 30, 399–406. [Google Scholar] [CrossRef] [PubMed]
  146. Maruta, T.; Tanouchi, A.; Tamoi, M.; Yabuta, Y.; Yoshimura, K.; Ishikawa, T.; Shigeoka, S. Arabidopsis chloroplastic ascorbate peroxidase isoenzymes play a dual role in photoprotection and gene regulation under photooxidative stress. Plant Cell Physiol. 2010, 51, 190–200. [Google Scholar] [CrossRef] [PubMed]
  147. Knox, J.P.; Dodge, A.D. Isolation and activity of the photodynamic pigment hypericin. Plant Cell Environ. 1985, 8, 19–25. [Google Scholar] [CrossRef]
  148. Zhao, Y.; Luo, L.; Xu, J.; Xin, P.; Guo, H.; Wu, J.; Bai, L.; Wang, G.; Chu, J.; Zuo, J.; et al. Malate transported from chloroplast to mitochondrion triggers production of ROS and PCD in Arabidopsis thaliana. Cell Res. 2018, 28, 448–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Gutteridge, J.M.; Halliwell, B. Comments on review of Free Radicals in Biology and Medicine, second edition, by Barry Halliwell and John, M.C. Gutteridge. Free Radic. Biol. Med. 1992, 12, 93–95. [Google Scholar] [CrossRef]
  150. Chmielowska-Bak, J.; Izbianska, K.; Ekner-Grzyb, A.; Bayar, M.; Deckert, J. Cadmium Stress Leads to Rapid Increase in RNA Oxidative Modifications in Soybean Seedlings. Front. Plant Sci. 2017, 8, 2219. [Google Scholar] [CrossRef] [Green Version]
  151. Xu, Y.; Yu, W.; Ma, Q.; Zhou, H.; Jiang, C. Toxicity of sulfadiazine and copper and their interaction to wheat (Triticum aestivum L.) seedlings. Ecotoxicol. Environ. Saf. 2017, 142, 250–256. [Google Scholar] [CrossRef]
  152. Öpik, H.; Rolfe, S.A.; Street, H.E. The Physiology of Flowering Plants, 4th ed.; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2005; 392p. [Google Scholar]
  153. Verslues, P.E.; Zhu, J.K. Before and beyond ABA: Upstream sensing and internal signals that determine ABA accumulation and response under abiotic stress. Biochem Soc. Trans. 2005, 33, 375–379. [Google Scholar] [CrossRef] [Green Version]
  154. Toh, S.; Imamura, A.; Watanabe, A.; Nakabayashi, K.; Okamoto, M.; Jikumaru, Y.; Hanada, A.; Aso, Y.; Ishiyama, K.; Tamura, N.; et al. High temperature-induced abscisic acid biosynthesis and its role in the inhibition of gibberellin action in Arabidopsis seeds. Plant Physiol. 2008, 146, 1368–1385. [Google Scholar] [CrossRef] [Green Version]
  155. Feurtado, J.A.; Ambrose, S.J.; Cutler, A.J.; Ross, A.R.; Abrams, S.R.; Kermode, A.R. Dormancy termination of western white pine (Pinus monticola Dougl. Ex, D. Don) seeds is associated with changes in abscisic acid metabolism. Planta 2004, 218, 630–639. [Google Scholar] [CrossRef]
  156. Zhao, K.; Zhou, Y.; Li, Y.; Zhuo, X.; Ahmad, S.; Han, Y.; Yong, X.; Zhang, Q. Crosstalk of PmCBFs and PmDAMs Based on the Changes of Phytohormones under Seasonal Cold Stress in the Stem of Prunus mume. Int. J. Mol. Sci. 2018, 19, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Mega, R.; Meguro-Maoka, A.; Endo, A.; Shimosaka, E.; Murayama, S.; Nambara, E.; Seo, M.; Kanno, Y.; Abrams, S.R.; Sato, Y. Sustained low abscisic acid levels increase seedling vigor under cold stress in rice (Oryza sativa L.). Sci. Rep. 2015, 5, 13819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Liu, C.T.; Wang, W.; Mao, B.G.; Chu, C.C. Cold stress tolerance in rice: Physiological changes, molecular mechanism, and future prospects. Yi Chuan 2018, 40, 171–185. [Google Scholar] [CrossRef] [PubMed]
  159. Xiong, L.; Ishitani, M.; Lee, H.; Zhu, J.K. The Arabidopsis LOS5/ABA3 locus encodes a molybdenum cofactor sulfurase and modulates cold stress- and osmotic stress-responsive gene expression. Plant Cell 2001, 13, 2063–2083. [Google Scholar]
  160. Xiong, L.; Lee, H.; Ishitani, M.; Zhu, J.K. Regulation of osmotic stress-responsive gene expression by the LOS6/ABA1 locus in Arabidopsis. J. Biol. Chem. 2002, 277, 8588–8596. [Google Scholar] [CrossRef] [Green Version]
  161. Fu, J.; Wu, Y.; Miao, Y.; Xu, Y.; Zhao, E.; Wang, J.; Sun, H.; Liu, Q.; Xue, Y.; Xu, Y.; et al. Improved cold tolerance in Elymus nutans by exogenous application of melatonin may involve ABA-dependent and ABA-independent pathways. Sci. Rep. 2017, 7, 39865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Baron, K.N.; Schroeder, D.F.; Stasolla, C. Transcriptional response of abscisic acid (ABA) metabolism and transport to cold and heat stress applied at the reproductive stage of development in Arabidopsis thaliana. Plant Sci. 2012, 188-189, 48–59. [Google Scholar] [CrossRef]
  163. Zeevaart, J.A.; Creelman, R.A. Metabolism and Physiology of Abscisic Acid. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1988, 39, 439–473. [Google Scholar] [CrossRef]
  164. Eremina, M.; Rozhon, W.; Poppenberger, B. Hormonal control of cold stress responses in plants. Cell Mol. Life Sci. 2016, 73, 797–810. [Google Scholar] [CrossRef]
  165. Huang, X.; Shi, H.; Hu, Z.; Liu, A.; Amombo, E.; Chen, L.; Fu, J. ABA Is Involved in Regulation of Cold Stress Response in Bermudagrass. Front. Plant Sci. 2017, 8, 1613. [Google Scholar] [CrossRef] [Green Version]
  166. Nakashima, K.; Yamaguchi-Shinozaki, K.; Shinozaki, K. The transcriptional regulatory network in the drought response and its crosstalk in abiotic stress responses including drought, cold, and heat. Front. Plant Sci. 2014, 5, 170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Shinozaki, K.; Yamaguchi-Shinozaki, K. Molecular responses to dehydration and low temperature: Differences and cross-talk between two stress signaling pathways. Curr. Opin. Plant Biol. 2000, 3, 217–223. [Google Scholar] [CrossRef]
  168. Baker, S.S.; Wilhelm, K.S.; Thomashow, M.F. The 5’-region of Arabidopsis thaliana cor15a has cis-acting elements that confer cold-, drought-and ABA-regulated gene expression. Plant Mol. Biol. 1994, 24, 701–713. [Google Scholar] [CrossRef] [PubMed]
  169. Yamaguchi-Shinozaki, K.; Shinozaki, K. A novel cis-acting element in an Arabidopsis gene is involved in responsiveness to drought, low-temperature, or high-salt stress. Plant Cell 1994, 6, 251–264. [Google Scholar]
  170. Maruyama, K.; Todaka, D.; Mizoi, J.; Yoshida, T.; Kidokoro, S.; Matsukura, S.; Takasaki, H.; Sakurai, T.; Yamamoto, Y.Y.; Yoshiwara, K. Identification of cis-acting promoter elements in cold-and dehydration-induced transcriptional pathways in Arabidopsis, rice, and soybean. DNA Res. 2012, 19, 37–49. [Google Scholar] [CrossRef] [PubMed]
  171. Ashraf, M.A.; Rahman, A. Hormonal Regulation of Cold Stress Response. In Cold Tolerance in Plants: Physiological, Molecular and Genetic Perspectives; Wani, S.H., Herath, V., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 65–88. [Google Scholar]
  172. Chen, C.-C.; Liang, C.-S.; Kao, A.-L.; Yang, C.-C. HHP1, a novel signalling component in the cross-talk between the cold and osmotic signalling pathways in Arabidopsis. J. Exp. Bot. 2010, 61, 3305–3320. [Google Scholar] [CrossRef] [Green Version]
  173. Gray, W.M.; Ostin, A.; Sandberg, G.; Romano, C.P.; Estelle, M. High temperature promotes auxin-mediated hypocotyl elongation in Arabidopsis. Proc. Natl. Acad. Sci. USA 1998, 95, 7197–7202. [Google Scholar] [CrossRef] [Green Version]
  174. Koini, M.A.; Alvey, L.; Allen, T.; Tilley, C.A.; Harberd, N.P.; Whitelam, G.C.; Franklin, K.A. High temperature-mediated adaptations in plant architecture require the bHLH transcription factor PIF4. Curr. Biol. 2009, 19, 408–413. [Google Scholar] [CrossRef] [Green Version]
  175. Shibasaki, K.; Uemura, M.; Tsurumi, S.; Rahman, A. Auxin response in Arabidopsis under cold stress: Underlying molecular mechanisms. Plant Cell 2009, 21, 3823–3838. [Google Scholar] [CrossRef] [Green Version]
  176. Hanzawa, T.; Shibasaki, K.; Numata, T.; Kawamura, Y.; Gaude, T.; Rahman, A. Cellular auxin homeostasis under high temperature is regulated through a sorting NEXIN1-dependent endosomal trafficking pathway. Plant Cell 2013, 25, 3424–3433. [Google Scholar] [CrossRef] [Green Version]
  177. Rahman, A. Auxin: A regulator of cold stress response. Physiol. Plant 2013, 147, 28–35. [Google Scholar] [CrossRef] [PubMed]
  178. Bellstaedt, J.; Trenner, J.; Lippmann, R.; Poeschl, Y.; Zhang, X.; Friml, J.; Quint, M.; Delker, C. A Mobile Auxin Signal Connects Temperature Sensing in Cotyledons with Growth Responses in Hypocotyls. Plant Physiol. 2019, 180, 757–766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Ashraf, M.A.; Rahman, A. Cold stress response in Arabidopsis thaliana is mediated by GNOM ARF-GEF. Plant J. 2019, 97, 500–516. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Du, H.; Wu, N.; Fu, J.; Wang, S.; Li, X.; Xiao, J.; Xiong, L. 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] [PubMed] [Green Version]
  181. 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] [Green Version]
  182. Hong, J.H.; Savina, M.; Du, J.; Devendran, A.; Kannivadi Ramakanth, K.; Tian, X.; Sim, W.S.; Mironova, V.V.; Xu, J. A Sacrifice-for-Survival Mechanism Protects Root Stem Cell Niche from Chilling Stress. Cell 2017, 170, 102–113.e114. [Google Scholar] [CrossRef] [Green Version]
  183. Urao, T.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Plant histidine kinases: An emerging picture of two-component signal transduction in hormone and environmental responses. Sci. STKE 2001, 2001, re18. [Google Scholar] [CrossRef]
  184. Jeon, J.; Kim, N.Y.; Kim, S.; Kang, N.Y.; Novak, O.; Ku, S.J.; Cho, C.; Lee, D.J.; Lee, E.J.; Strnad, M.; et al. 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] [Green Version]
  185. Zwack, P.J.; Compton, M.A.; Adams, C.I.; Rashotte, A.M. Cytokinin response factor 4 (CRF4) is induced by cold and involved in freezing tolerance. Plant Cell Rep. 2016, 35, 573–584. [Google Scholar] [CrossRef]
  186. Jeon, J.; Cho, C.; Lee, M.R.; Van Binh, N.; Kim, J. CYTOKININ RESPONSE FACTOR2 (CRF2) and CRF3 Regulate Lateral Root Development in Response to Cold Stress in Arabidopsis. Plant Cell 2016, 28, 1828–1843. [Google Scholar] [CrossRef] [Green Version]
  187. Wang, H.; Huang, J.; Liang, X.; Bi, Y. Involvement of hydrogen peroxide, calcium, and ethylene in the induction of the alternative pathway in chilling-stressed Arabidopsis callus. Planta 2012, 235, 53–67. [Google Scholar] [CrossRef] [PubMed]
  188. Yu, X.M.; Griffith, M.; Wiseman, S.B. Ethylene induces antifreeze activity in winter rye leaves. Plant Physiol. 2001, 126, 1232–1240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Ciardi, J.A.; Deikman, J.; Orzolek, M.D. Increased ethylene synthesis enhances chilling tolerance in tomato. Physiol. Plant. 1997, 101, 333–340. [Google Scholar] [CrossRef]
  190. Klay, I.; Gouia, S.; Liu, M.; Mila, I.; Khoudi, H.; Bernadac, A.; Bouzayen, M.; Pirrello, J. Ethylene Response Factors (ERF) are differentially regulated by different abiotic stress types in tomato plants. Plant Sci. 2018, 274, 137–145. [Google Scholar] [CrossRef]
  191. Bolt, S.; Zuther, E.; Zintl, S.; Hincha, D.K.; Schmulling, T. ERF105 is a transcription factor gene of Arabidopsis thaliana required for freezing tolerance and cold acclimation. Plant Cell Environ. 2017, 40, 108–120. [Google Scholar] [CrossRef]
  192. Catala, R.; Salinas, J. The Arabidopsis ethylene overproducer mutant eto1-3 displays enhanced freezing tolerance. Plant Signal. Behav. 2015, 10, e989768. [Google Scholar] [CrossRef] [Green Version]
  193. Castorina, G.; Persico, M.; Zilio, M.; Sangiorgio, S.; Carabelli, L.; Consonni, G. The maize lilliputian1 (lil1) gene, encoding a brassinosteroid cytochrome P450 C-6 oxidase, is involved in plant growth and drought response. Ann. Bot. 2018, 122, 227–238. [Google Scholar] [CrossRef]
  194. Krishna, P.; Prasad, B.D.; Rahman, T. Brassinosteroid Action in Plant Abiotic Stress Tolerance. Methods Mol. Biol. 2017, 1564, 193–202. [Google Scholar] [CrossRef]
  195. Ahammed, G.J.; Xia, X.J.; Li, X.; Shi, K.; Yu, J.Q.; Zhou, Y.H. Role of brassinosteroid in plant adaptation to abiotic stresses and its interplay with other hormones. Curr. Protein. Pept. Sci. 2015, 16, 462–473. [Google Scholar] [CrossRef]
  196. Chung, Y.; Kwon, S.I.; Choe, S. Antagonistic regulation of Arabidopsis growth by brassinosteroids and abiotic stresses. Mol. Cells 2014, 37, 795–803. [Google Scholar] [CrossRef] [Green Version]
  197. 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]
  198. Kim, S.Y.; Kim, B.H.; Lim, C.J.; Lim, C.O.; Nam, K.H. Constitutive activation of stress-inducible genes in a brassinosteroid-insensitive 1 (bri1) mutant results in higher tolerance to cold. Physiol. Plant 2010, 138, 191–204. [Google Scholar] [CrossRef] [PubMed]
  199. Divi, U.K.; Krishna, P. Brassinosteroid: A biotechnological target for enhancing crop yield and stress tolerance. N. Biotechnol. 2009, 26, 131–136. [Google Scholar] [CrossRef] [PubMed]
  200. Eremina, M.; Unterholzner, S.J.; Rathnayake, A.I.; Castellanos, M.; Khan, M.; Kugler, K.G.; May, S.T.; Mayer, K.F.; Rozhon, W.; Poppenberger, B. Brassinosteroids participate in the control of basal and acquired freezing tolerance of plants. Proc. Natl. Acad. Sci. USA 2016, 113, E5982–E5991. [Google Scholar] [CrossRef] [Green Version]
  201. Kim, J.M.; Sasaki, T.; Ueda, M.; Sako, K.; Seki, M. Chromatin changes in response to drought, salinity, heat, and cold stresses in plants. Front. Plant Sci. 2015, 6, 114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Pandey, G.; Sharma, N.; Sahu, P.P.; Prasad, M. Chromatin-Based Epigenetic Regulation of Plant Abiotic Stress Response. Curr. Genom. 2016, 17, 490–498. [Google Scholar] [CrossRef] [Green Version]
  203. To, T.K.; Nakaminami, K.; Kim, J.M.; Morosawa, T.; Ishida, J.; Tanaka, M.; Yokoyama, S.; Shinozaki, K.; Seki, M. Arabidopsis HDA6 is required for freezing tolerance. Biochem. Biophys. Res. Commun. 2011, 406, 414–419. [Google Scholar] [CrossRef] [PubMed]
  204. Lim, C.J.; Park, J.; Shen, M.; Park, H.J.; Cheong, M.S.; Park, K.S.; Baek, D.; Bae, M.J.; Ali, A.; Jan, M.; et al. The histone-modifying complex PWR/HOS15/HD2C epigenetically regulates cold tolerance. Plant Physiol. 2020, 184, 1097–1111. [Google Scholar] [CrossRef]
  205. Zhu, J.; Jeong, J.C.; Zhu, Y.; Sokolchik, I.; Miyazaki, S.; Zhu, J.K.; Hasegawa, P.M.; Bohnert, H.J.; Shi, H.; Yun, D.J.; et al. Involvement of Arabidopsis HOS15 in histone deacetylation and cold tolerance. Proc. Natl. Acad. Sci. USA 2008, 105, 4945–4950. [Google Scholar] [CrossRef] [Green Version]
  206. Park, J.; Lim, C.J.; Shen, M.; Park, H.J.; Cha, J.Y.; Iniesto, E.; Rubio, V.; Mengiste, T.; Zhu, J.K.; Bressan, R.A.; et al. Epigenetic switch from repressive to permissive chromatin in response to cold stress. Proc. Natl. Acad. Sci. USA 2018, 115, E5400–E5409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Hu, Y.; Zhang, L.; Zhao, L.; Li, J.; He, S.; Zhou, K.; Yang, F.; Huang, M.; Jiang, L.; Li, L. Trichostatin A selectively suppresses the cold-induced transcription of the ZmDREB1 gene in maize. PLoS ONE 2011, 6, e22132. [Google Scholar] [CrossRef] [PubMed]
  208. Hu, Y.; Zhang, L.; He, S.; Huang, M.; Tan, J.; Zhao, L.; Yan, S.; Li, H.; Zhou, K.; Liang, Y.; et al. Cold stress selectively unsilences tandem repeats in heterochromatin associated with accumulation of H3K9ac. Plant Cell Environ. 2012, 35, 2130–2142. [Google Scholar] [CrossRef] [PubMed]
  209. Kindgren, P.; Ard, R.; Ivanov, M.; Marquardt, S. Transcriptional read-through of the long non-coding RNA SVALKA governs plant cold acclimation. Nat. Commun. 2018, 9, 4561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. An, F.M.; Hsiao, S.R.; Chan, M.T. Sequencing-based approaches reveal low ambient temperature-responsive and tissue-specific microRNAs in phalaenopsis orchid. PLoS ONE 2011, 6, e18937. [Google Scholar] [CrossRef] [Green Version]
  211. Arenas-Huertero, C.; Pérez, B.; Rabanal, F.; Blanco-Melo, D.; De la Rosa, C.; Estrada-Navarrete, G.S.F.; Covarrubias, A.A.; Reyes, J.L. Conserved and novel miRNAs in the legume Phaseolus vulgaris in response to stress. Plant Mol. Biol. 2009, 70, 385–401. [Google Scholar] [CrossRef]
  212. Barakat, A.; Sriram, A.; Park, J.; Zhebentyayeva, T.; Main, D.; Abbott, A. Genome wide identification of chilling responsive microRNAs in Prunus persica. BMC Genom. 2012, 13, 481. [Google Scholar] [CrossRef] [Green Version]
  213. Barrera-Figueroa, B.E.; Gao, L.; Wu, Z.; Zhou, X.; Zhu, J.; Jin, H.; Liu, R.; Zhu, J.K. High throughput sequencing reveals novel and abiotic stress-regulated microRNAs in the inflorescences of rice. BMC Plant Biol. 2012, 12, 132. [Google Scholar] [CrossRef] [Green Version]
  214. Byun, Y.J.; Kim, H.J.; Lee, D.H. LongSAGE analysis of the early response to cold stress in Arabidopsis leaf. Planta 2009, 229, 1181–1200. [Google Scholar] [CrossRef]
  215. Chen, L.; Zhang, Y.; Ren, Y.; Xu, J.; Zhang, Z.; Wang, Y. Genome-wide identification of cold-responsive and new microRNAs in Populus tomentosa by high-throughput sequencing. Biochem. Biophys. Res. Commun. 2012, 417, 892–896. [Google Scholar] [CrossRef]
  216. Giacomelli, J.I.; Weigel, D.; Chan, R.L.; Manavella, P.A. Role of recently evolved miRNA regulation of sunflower HaWRKY6 in response to temperature damage. New Phytol. 2012, 195, 766–773. [Google Scholar] [CrossRef]
  217. Jung, H.J.; Kang, H. Expression and functional analyses of microRNA417 in Arabidopsis thaliana under stress conditions. Plant Physiol. Biochem. 2007, 45, 805–811. [Google Scholar] [CrossRef] [PubMed]
  218. Kim, W.; Ahn, H.J.; Chiou, T.J.; Ahn, J.H. The role of the miR399-PHO2 module in the regulation of flowering time in response to different ambient temperatures in Arabidopsis thaliana. Mol. Cells 2011, 32, 83–88. [Google Scholar] [CrossRef] [PubMed]
  219. Lee, H.; Yoo, S.J.; Lee, J.H.; Kim, W.; Yoo, S.K.; Fitzgerald, H.; Carrington, J.C.; Ahn, J.H. Genetic framework for flowering-time regulation by ambient temperature-responsive miRNAs in Arabidopsis. Nucleic Acids. Res. 2010, 38, 3081–3093. [Google Scholar] [CrossRef] [PubMed]
  220. Liu, H.H.; Tian, X.; Li, Y.J.; Wu, C.A.; Zheng, C.C. Microarray-based analysis of stress-regulated microRNAs in Arabidopsis thaliana. RNA 2008, 14, 836–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Lu, S.; Sun, Y.H.; Chiang, V.L. Stress-responsive microRNAs in Populus. Plant J. 2008, 55, 131–151. [Google Scholar] [CrossRef]
  222. Lv, D.K.; Bai, X.; Li, Y.; Ding, X.D.; Ge, Y.; Cai, H.; Ji, W.; Wu, N.; Zhu, Y.M. Profiling of cold-stress-responsive miRNAs in rice by microarrays. Gene 2010, 459, 39–47. [Google Scholar] [CrossRef]
  223. Reyes, J.L.; Chua, N.H. ABA induction of miR159 controls transcript levels of two MYB factors during Arabidopsis seed germination. Plant, J. 2007, 49, 592–606. [Google Scholar] [CrossRef]
  224. Shen, J.; Xie, K.; Xiong, L. Global expression profiling of rice microRNAs by one-tube stem-loop reverse transcription quantitative PCR revealed important roles of microRNAs in abiotic stress responses. Mol. Genet. Genom. 2010, 284, 477–488. [Google Scholar] [CrossRef]
  225. Tang, Z.; Zhang, L.; Xu, C.; Yuan, S.; Zhang, F.; Zheng, Y.; Zhao, C. Uncovering small RNA-mediated responses to cold stress in a wheat thermosensitive genic male-sterile line by deep sequencing. Plant Physiol. 2012, 159, 721–738. [Google Scholar] [CrossRef] [Green Version]
  226. Thiebaut, F.; Rojas, C.A.; Almeida, K.L.; Grativol, C.; Domiciano, G.C.; Lamb, C.R.; Engler, J.A.; Hemerly, A.S.; Ferreira, P.C. Regulation of miR319 during cold stress in sugarcane. Plant Cell Environ. 2012, 35, 502–512. [Google Scholar] [CrossRef]
  227. Xia, K.; Wang, R.; Ou, X.; Fang, Z.; Tian, C.; Duan, J.; Wang, Y.; Zhang, M. OsTIR1 and OsAFB2 downregulation via OsmiR393 overexpression leads to more tillers, early flowering and less tolerance to salt and drought in rice. PLoS ONE 2012, 7, e30039. [Google Scholar] [CrossRef] [PubMed]
  228. Zhang, J.; Xu, Y.; Huan, Q.; Chong, K. Deep sequencing of Brachypodium small RNAs at the global genome level identifies microRNAs involved in cold stress response. BMC Genom. 2009, 10, 449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Zhou, X.; Wang, G.; Sutoh, K.; Zhu, J.K.; Zhang, W. Identification of cold-inducible microRNAs in plants by transcriptome analysis. Biochim. Biophys. Acta 2008, 1779, 780–788. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Significant events during cold stress response in plants.
Figure 1. Significant events during cold stress response in plants.
Agronomy 12 00702 g001
Figure 2. A simplified model of calcium signaling in plants. Calcium has a unique ability to adjust itself in the cytosol at the time of stress or optimal growth challenges. Its occurrence per period, time span, magnitude and location oscillates with the kind and the degree of stress. Ca2+ signals are further decode by calcium sensors resulting in the transcriptional regulation and ultimately stress response.
Figure 2. A simplified model of calcium signaling in plants. Calcium has a unique ability to adjust itself in the cytosol at the time of stress or optimal growth challenges. Its occurrence per period, time span, magnitude and location oscillates with the kind and the degree of stress. Ca2+ signals are further decode by calcium sensors resulting in the transcriptional regulation and ultimately stress response.
Agronomy 12 00702 g002
Figure 3. The schematic representation of various components involved in cold stress signaling and response. Cold stress-induced membrane modification and changes in calcium ion concentration leads to the downstream signaling events. Calcium ion signal is perceived by various calcium sensors (e.g, Calmodulin (CaM), Calcineurin B-like proteins (CBLs), CBL-interacting protein kinases (CIPKs), Calcium-dependent protein kinases (CDPKs) and CRLK present in cytosol, plasma membrane and membranes of chloroplast, mitochondria, vacuole and nucleus, etc. These signals further activate other components of cold stress response machinery such as the MAPK cascade, which ultimately regulates the expression level of ICE1, CBF, and other cold regulated gene expression.
Figure 3. The schematic representation of various components involved in cold stress signaling and response. Cold stress-induced membrane modification and changes in calcium ion concentration leads to the downstream signaling events. Calcium ion signal is perceived by various calcium sensors (e.g, Calmodulin (CaM), Calcineurin B-like proteins (CBLs), CBL-interacting protein kinases (CIPKs), Calcium-dependent protein kinases (CDPKs) and CRLK present in cytosol, plasma membrane and membranes of chloroplast, mitochondria, vacuole and nucleus, etc. These signals further activate other components of cold stress response machinery such as the MAPK cascade, which ultimately regulates the expression level of ICE1, CBF, and other cold regulated gene expression.
Agronomy 12 00702 g003
Figure 4. A diagrammatic representation of the synchronous effect of ROS, high light and cold stress inside the cell environment. As shown in figure chloroplast is the site for carbon reduction and shared energy cycle with the mitochondria. ROS accumulation leads to programmed cell death (PCD), an interesting way to escape oxidative damage. A malate shuttle between chloroplast and mitochondria stabilizes the levels of ROS. Numerous antioxidant systems prevail in the cell and are effective in scavenging ROS.
Figure 4. A diagrammatic representation of the synchronous effect of ROS, high light and cold stress inside the cell environment. As shown in figure chloroplast is the site for carbon reduction and shared energy cycle with the mitochondria. ROS accumulation leads to programmed cell death (PCD), an interesting way to escape oxidative damage. A malate shuttle between chloroplast and mitochondria stabilizes the levels of ROS. Numerous antioxidant systems prevail in the cell and are effective in scavenging ROS.
Agronomy 12 00702 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Aslam, M.; Fakher, B.; Ashraf, M.A.; Cheng, Y.; Wang, B.; Qin, Y. Plant Low-Temperature Stress: Signaling and Response. Agronomy 2022, 12, 702. https://doi.org/10.3390/agronomy12030702

AMA Style

Aslam M, Fakher B, Ashraf MA, Cheng Y, Wang B, Qin Y. Plant Low-Temperature Stress: Signaling and Response. Agronomy. 2022; 12(3):702. https://doi.org/10.3390/agronomy12030702

Chicago/Turabian Style

Aslam, Mohammad, Beenish Fakher, Mohammad Arif Ashraf, Yan Cheng, Bingrui Wang, and Yuan Qin. 2022. "Plant Low-Temperature Stress: Signaling and Response" Agronomy 12, no. 3: 702. https://doi.org/10.3390/agronomy12030702

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