Next Article in Journal
Bone Regeneration of a 3D-Printed Alloplastic and Particulate Xenogenic Graft with rhBMP-2
Next Article in Special Issue
HbSnRK2.6 Functions in ABA-Regulated Cold Stress Response by Promoting HbICE2 Transcriptional Activity in Hevea brasiliensis
Previous Article in Journal
Genetic Biomarkers in Chronic Myeloid Leukemia: What Have We Learned So Far?
Previous Article in Special Issue
Wood Formation under Severe Drought Invokes Adjustment of the Hormonal and Transcriptional Landscape in Poplar
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Non-Coding RNAs in Response to Drought Stress

by
Temesgen Assefa Gelaw
1,2 and
Neeti Sanan-Mishra
1,*
1
Plant RNAi Biology Group, International Centre for Genetic Engineering and Biotechnology, New Delhi 110067, India
2
Department of Biotechnology, College of Natural and Computational Science, Debre Birhan University, Debre Birhan P.O. Box 445, Ethiopia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(22), 12519; https://doi.org/10.3390/ijms222212519
Submission received: 21 October 2021 / Revised: 11 November 2021 / Accepted: 15 November 2021 / Published: 20 November 2021
(This article belongs to the Special Issue Mechanisms of Drought, Temperature and Salinity Tolerance in Plants)

Abstract

:
Drought stress causes changes in the morphological, physiological, biochemical and molecular characteristics of plants. The response to drought in different plants may vary from avoidance, tolerance and escape to recovery from stress. This response is genetically programmed and regulated in a very complex yet synchronized manner. The crucial genetic regulations mediated by non-coding RNAs (ncRNAs) have emerged as game-changers in modulating the plant responses to drought and other abiotic stresses. The ncRNAs interact with their targets to form potentially subtle regulatory networks that control multiple genes to determine the overall response of plants. Many long and small drought-responsive ncRNAs have been identified and characterized in different plant varieties. The miRNA-based research is better documented, while lncRNA and transposon-derived RNAs are relatively new, and their cellular role is beginning to be understood. In this review, we have compiled the information on the categorization of non-coding RNAs based on their biogenesis and function. We also discuss the available literature on the role of long and small non-coding RNAs in mitigating drought stress in plants.

1. Introduction

About 80–95% fresh biomass of non-woody plants is occupied by water, which plays an important role in many aspects of plant life. Several abiotic factors, such as low rainfall, salinity, very high or very low temperature and high intensity of light lead to water deficit in plants [1]. The reduction in water exerts stress leading to changes in the physiological, morphological, ecological, biochemical and molecular characteristics [2,3]. These changes can affect overall plant development resulting in yield reduction and/or plant loss [4]. In arid and semi-arid environments, drought is one of the most important stress factors for plants [1,4,5]. The continuous increase in environmental temperature has increased the probability of occurrence, duration and severity of drought, thereby making it challenging to meet the global food demands of the ever-increasing world population [6,7].
Drought stress tolerance is a quantitatively controlled trait in plants [8,9]. It causes changes in gene-expression patterns, water transport and osmotic balance, oxidative homeostasis and repair mechanisms. This affects the vital metabolic processes, chlorophyll synthesis and photosynthesis, decelerates seed germination, reduces stomatal movement, inhibits root development and limits nutrient uptake [3,6,10]. As sessile organisms, plants have evolved several mechanisms to withstand water stress and/or drought by inducing changes at the developmental and functional levels [5,9,11,12]. The resistance mechanisms include strategies for avoiding, escaping, tolerating and recovering from drought [13]. To tide over short periods of drought or ‘drought escape’ plants have the ability to regulate growth [10,11]. Once the stress is over, plants resume growth to overcome drought injury and this is known as drought recovery [13].
An important universal physiological process to overcome drought stress involves regulating stomata movement to control respiration, transpiration, photosynthesis and temperature [11,14,15]. Plants can also endure severe water-stress scenarios through osmotic adjustments and accumulation of dehydration-induced proteins [10,12,16,17]. The other changes include regulating the onset of senescence and fine-tuning of phytohormones [2,9,10,12,16] (Figure 1). Plants also modulate the redox pathway by balancing the production of antioxidant enzymes, such as including superoxide dismutase, peroxidase and ascorbate peroxidase, to scavenge the reactive oxygen species (ROS) produced during drought stress [18]. This also helps to maintain organelle stability, protect chloroplast membranes and stabilize the PSII system [19]. Therefore, it is important to identify the factors that regulate the genetic components and govern the nature of plant response.
Relatively recent studies have shown that long and small non-coding RNAs (ncRNAs) are important modulators of drought tolerance in plants [12,20,21,22,23,24]. The networking between ncRNAs and their target genes is, in turn, controlled by various other enzymatic components in the cell [9,25]. The advances in high-throughput analysis, such as RNA sequencing (RNA-Seq) and bioinformatics, have accelerated scientific research [26]. Sequence analysis has proved to be an important tool to explore the differences in response to stress between sensitive and tolerant plants, such as sorghum [27], tomato [28], coffee [29], cassava [30], peanut [31], Populus [32], Trifolium [33], wheat [34], rice [35] and maize [36,37]. This has led to the identification of stress-responsive gene expression; however, our knowledge about the regulatory processes is still limited. This review highlights important updates on the available literature on the role of long and small ncRNAs in response to drought stress response in plants.

2. Long Non-Coding RNAs

Over the last decade, long ncRNAs (lncRNAs), usually >200 nucleotides (nt) in size, have emerged as a pivot of genome regulation, adding a new layer of epigenetic control [38,39,40], but their clear evolutionary origins and functional specialization are still inexplicit. This group of ncRNAs lacks polypeptide-coding potential and possesses significant biochemical versatility, with each lncRNA having specific functions [39,41,42]. High-throughput RNA-Seq based investigations have primarily contributed to the identification of lncRNAs in many organisms [43]. Many investigations have been performed in humans and animals [44,45,46]; however, in plants, only a few molecules have been well characterized [39,47,48].
In eukaryotes, transcription followed by processing of the nascent RNA precedes the formation of messenger (mRNA). The biogenesis of a fully mature mRNA requires the coordinated action of enzymes that perform 7-methyl Guanosine (m7G) capping, splicing, polyadenylation, maturation, export and surveillance [49,50]. In a similar manner, primary transcripts of lncRNAs, which are produced by RNA Polymerases II or III and IV or V, undergo polyadenylation, m7G capping and splicing [24,51]. Most plant lncRNAs are polyadenylated, but in mammals and yeast, some non-polyadenylated lncRNAs have been found [52]; therefore, the presence of non-polyadenylated functional lncRNAs in plants cannot be completely ruled out [53]. The non-polyadenylated lncRNAs are processed by Ribonuclease P to generate free ends that are stabilized by the capping of small nucleolar RNA (snoRNA). In some cases, the snoRNAs have been found at both 3′ and 5′ ends. Reports have also shown that Ribonuclease P processed free ends can self-ligate to form circular structures [39,40,54,55].
Based on their biogenesis and location, the lncRNAs can be classified into several subgroups [56], as described below.
(a) Long intergenic ncRNAs (lincRNAs) are 200 to 2000 nt in length. These transcripts are derived from the intergenic region and have also been referred to as large/long intervening ncRNAs, very long intervening ncRNAs and macroRNAs [24,57]. They are polyadenylated, weakly spliced, exhibit tissue-specific expression and possess a trans-regulatory role [39,58]. These lncRNAs are characterized by rapid turnover rates, which present a challenge in understanding their functional significance [39,59]. The lincRNAs are localized at least 5 kb away from protein-coding regions and are, therefore, said to lie in gene deserts [60]. These have been sub-classified based on their association with specific regions [61,62], such as enhancer RNAs or eRNAs [52,60], upstream antisense RNA (uaRNA), promoter-associated long RNA (PALR) or promoter upstream transcripts (PROMPTs) [63] and telomeric repeat-containing RNA (TERRA). The PROMPTs and eRNAs are short-lived lncRNAs that have been identified mostly in humans.
(b) Transposable element (TE)-derived lncRNAs (TE-lncRNAs) are generated from the TEs [57]. These transcripts can sometimes act as precursors to microRNAs (miRNAs) and small interfering RNAs (siRNAs) [64,65,66]. In Arabidopsis, rice, maize and tomato TE-lncRNAs were reported, and their association with response to abiotic stresses was revealed [67,68,69,70]. TEs, also known as jumping genes, were first discovered in maize in the early 1950s [71]. They have the ability to copy/cut and paste themselves into other genome regions [72]. Based on the mechanism of transposition, they have been classified as Class I or RNA mediated/retro-elements and Class II or DNA elements. Class II classical TEs transpose via a cut-and-paste mechanism, while Class I TEs transpose through a reverse-transcription-based copy-and-paste mechanism. The DNA copy thus generated can get integrated anywhere in the plant genome [65,72,73]. In the maize genome, more than 85% of DNA is derived from TEs [72,74], so, correspondingly, a large number of TE-lncRNAs have been predicted to be present in maize [75,76].
(c) Intron-derived lncRNAs (incRNAs) originate from within the introns of protein-coding genes [77] and include totally intronic RNA (TIN) or partially intronic RNA (PIN). These transcripts therefore, are regulated by various transcription activation pathways [78]. The incRNAs may contain poly(A) modifications and are usually stable but they may not be highly conserved across different plant species [44,79]. It was reported that few miRNAs and snoRNAs originate from the intronic regions, so, initially, it was considered that the incRNAs may act as progenitors of the small ncRNAs; however, recent studies have confirmed their independent existence [80,81,82].
(d) Natural antisense transcripts (NATs) originate from coding regions (both exon and intron) in eukaryotic genomes and are amongst the widespread lncRNAs. They possess both cis and trans-action to regulate gene expression by silencing [83,84]. Cis-NATs are transcribed from the inverse strand of the target genomic locus to regulate the corresponding sense transcript [83]. Trans-NATs arise from a locus away from that of the target gene [83,85]. The binding of NATs triggers the production of specific siRNAs, which also exhibit a predominantly trans mode of action.
(e) Circular lncRNAs (circncRNAs) are highly conserved but low in abundance and are more stable than linear lncRNAs, as they cannot be degraded easily [86]. They were first characterized as non-polyadenylated circular RNAs in plant viroids [87]. CircncRNAs mostly arise in the nucleus from the back-splicing of exons in pre-mRNAs [39,86,88,89], while some arise in the cytoplasm. It is postulated that the failure of intronic lariat debranching during canonical splicing plays a role in the biogenesis of circncRNAs [88,90,91,92]. Most circncRNAs may consist of one or more extra exons and are categorized as extra-exon circular ncRNAs (eecircncRNAs), and others may be derived from the intron region of the parent gene and called circular intronic RNAs (circincRNAs) or intron retained circular ncRNAs; however, some arise from overlapping regions and are called exon-intron circncRNAs (eicincRNAs) [90,93]. Since circncRNAs are derived from the internal exon regions, they can affect the splicing of their linear counterparts. They have been shown to regulate cell development by acting as endogenous target mimics (eTM) of miRNAs, miRNA sponges [43,91], protein scaffolds or templates for protein translation. The circncRNAs present in exosomes were shown to regulate the proliferation of the respective cells [92].

2.1. Function of lncRNAs

It is clear that a large number of lncRNAs are transcribed in plant cells, but their molecular mechanism is largely unknown [94,95,96,97,98]. They mainly play a regulatory role by facilitating gene silencing to control transcriptional regulation and genome imprinting. These regulations are associated with diverse biological processes, such as root organogenesis [99], photo-morphogenesis [100], control of flowering time, reproduction, nutrient homeostasis [95] and so on [58,67,101]. Their expression levels vary significantly in different tissues and over different developmental stages. The lncRNAs also undergo dynamic regulatory adjustments during the response to abiotic stress [96,102,103,104,105,106] and pathogen invasion [101].
It has been generalized that the Pol IV transcribed lncRNAs serve as precursors for siRNAs, while Pol V transcribed lncRNAs act by modulating the chromatin framework [107]. The lncRNAs can act in cis or trans orientation, forming simple and complex networks. For instance, in Arabidopsis, ~1400 light-responsive NATs were identified, and they could act in both the same and opposite directions [106,108]. They may act as signal molecules guide molecules, precursors for miRNAs and siRNAs, regulators of pre-mRNA splicing and modulators of chromatin [70,107,109,110,111]. Some of the identified functions of lncRNAs are listed in Table 1.

2.1.1. lncRNAs as Target Mimics

The lncRNAs can regulate transcription directly or by acting as target mimics of the small ncRNAs [57,124,125]. Some lncRNAs, such as IPS1 (induced by phosphate starvation) and ASCO (alternative splicing competitor), contain sequences complementary to that of the miRNAs, so they can act as decoys or sponges or eTMs by competing for miRNA binding [59,112,125,126]. The mimic sites within the lncRNAs are non-cleavable and block the miRNA function, thus conferring translational regulation in cis orientation. Several of these target mimics are thought to have a role in plant growth and development [127].
The discovery of the IPS1 gene in Arabidopsis introduced the concept of eTMs and unveiled the novel cellular mechanism behind the regulation of the miRNAs [128]. IPS1 contains a region complementary to miR399, but it has a loop at the cleavage point of miR399. Thus, IPS1:miR399 forms a stable pair and quenches the silencing activity of the miRNA. Both IPS1 and miR399 expression are induced upon phosphate starvation and IPS1 expression seems to be required for fine-tuning of miR399 activity [112]. Subsequently, it was shown that lncRNA23468 functions as a decoy for miR482b to compete with transcripts of NBS-LRR genes [129].
The ASCO-lncRNA binds to transcripts encoding nuclear alternative splicing regulators, AtNSRa and AtNSRb to regulate lateral root development [111]. In addition, lncRNA16397 targets slGRX22 (glutaredoxin gene) to induce the expression of GRX21 and reduce ROS accumulation [130]. The dual regulators act to modulate gene expression during Phytophthora infestans infection in tomatoes.
Reports have also indicated that lncRNAs can be targeted and cleaved by the miRNAs [118,131,132]. A computational study by Fan et al. [59] found 466 maize lncRNAs as targets of 165 miRNAs and 86 lncRNAs as decoys for 58 miRNAs. In Populus, about 51 lncRNAs were reported as putative miRNA targets and 20 lncRNAs were reported as target mimics of the known miRNAs in response to drought stress [118]. In Cleistogenes, RNA-Seq analysis identified 52 lncRNAs as target mimics for miRNAs [97]. These analyses clearly showed that lncRNAs are associated with the miRNA nodes and supported their regulatory role in plants.

2.1.2. lncRNAs in DNA Modification

Numerous reports have indicated that, in plants, the lncRNAs participate in the modification of DNA at different developmental stages [105] such as reproduction [133], embryogenesis [134] and organogenesis [92] under stress conditions. The classical example is provided by cold-induced incRNA, COOLAIR, which inhibits FLC (Flowering Locus C) during vernalization to regulate the flowering time in Arabidopsis [135]. The repression is achieved by enrichment of Polycomb repressive complex 2 (PRC2) and subsequent trimethylation of Histone H3 Lysine 27 (H3K27me3) at the FLC locus [136].
Epigenetic silencing via DNA methylation and histone modification is an important mechanism for regulating gene expression. It is specifically significant for controlling transposons, repetitive sequences and centromeric regions and for imprinting [48,70,137]. The lncRNAs can also guide gene silencing through siRNA-dependent DNA methylation [47]. The role of small RNA-directed DNA methylation (RdDM) and heterochromatinization has been well studied in plants [138,139,140]. The plant-specific RNA polymerases, RNA Pol IV and V play a crucial role in this process [24,51,141,142,143,144]. Briefly, Pol IV, along with the CLASSY chromatin remodeling factors (CCRFs) and homeodomain transcription factors, such as DTF1/ SHH1, transcribes transposons and repetitive sequences. The transcripts are converted to double-stranded RNAs (dsRNAs) by the action of RNA-dependent RNA polymerase-2 (RDR2) and the dsRNAs are processed into small ncRNAs, specifically siRNA duplexes by Dicer-like 3 (DCL3) enzyme [138,139]. These siRNAs are loaded in Argonaute 4 (AGO4)-containing complex to guide RdDM. In an alternate pathway, siRNAs are generated through Pol I–RDR6 transcription and are loaded into the AGO6 complex. At loci where Pol V is producing nascent transcripts, the siRNA-guided AGO4,6 complex interacts with the larger subunit of RNA Pol V, NRPE1 [145,146]. This complex is stabilized by the KTF1 (yeast transcription elongation factor, SPT5 homolog) to subsequently establish DNA methylation through domains rearranged methyltransferase 2 (DRM2) [141,146]. The methylated state of DNA is maintained through cell divisions through pathways catalyzed by methyltransferase1 (MET1) or chromomethylase3 (CMT3) [117,147].
It was shown that RdDM regulates the repetitive intergenic elements and their expression in maize. The RdDM function is supported by mediator of paramutation 1 (MOP1-1) in maize, which is an ortholog of At-RDR2 [148]. It was shown that MOP1 copies the RNA Pol IV transcript for processing into siRNAs [144]. In another study, 110 maize lincRNAs and 46 genic lncRNAs were predicted as precursors for Mop1-sensitive siRNAs [149]. In addition, 26 lincRNAs and 97 genic lncRNAs were predicted as precursors for shRNA, while one lincRNA and two genic lncRNAs were predicted as precursors for miRNA. RdDM is a complex pathway that has also been implicated with short-term and long-term stress memory [105], so further investigations are required to understand the role of lincRNAs in regulating RdDM functions in maize in response to drought and other abiotic stresses.

3. Small Non-Coding RNAs

The small ncRNAs comprise a number of categories among which the miRNAs and siRNAs constitute the major groups. They function as key regulators of transcriptional and post-transcriptional gene expression [139,150,151,152] and are therefore implicated in the control of various physiological and developmental processes in plants, such as growth, organ formation, phase transition, nutrient balance and stress response [10,22,153,154,155,156]. Several online tools and databases have been developed that have enabled the prediction, documentation and analysis of the small ncRNAs and their targets [48,57,124]. Deep sequencing and degradome analyses, coupled with advanced tools and databases, have driven the identification of various small ncRNAs in response to single or combined abiotic stresses [20,157,158,159].

3.1. Small Interfering RNAs

Overall, siRNAs are generally 21–24 nt in length and are produced by the sequential processing of long dsRNAs in a phased or non-overlapping manner. They may arise either from endogenous sources, such as TEs, repetitive elements and centromere, or exogenous sources, such as invading viruses or aberrant inverted repeats [139,160]. The siRNAs can target endogenous as well as exogenous sequences serving as the first line of host defense [161]. The long dsRNAs are processed by DCLs into mature siRNAs, which get associated with AGO protein to form the catalytic core of the RNA-induced silencing complex (RISC) to facilitate gene silencing [139,162]. The siRNA strand that directs the RISC complex is called the guide strand, while the other strand is known as the passenger strand. The passenger strand is excluded and undergoes degradation, while the guide strand directs RISC to its target transcript for cleavage. The identity of the guide and passenger strand is regulated on the basis of the thermodynamic stability of 5′ end [163,164]. The siRNAs also mediate transcriptional gene silencing through the RNA induced transcriptional silencing (RITS) complex [139].
The siRNAs are involved in regulating gene expression, maintaining genome stability and aiding plant defense. In Arabidopsis, DCL2 and DCL4 are involved in production of primary siRNAs from aberrant dsRNAs. DCL2 processes 22 nt siRNAs that contribute to the antiviral defense and plant development while DCL4 processes 21 nt siRNAs to initiate primary defense against invasion of viruses and transgenes [113,139]. The DCL3 processes 24 nt siRNAs to direct methylation of DNA sequences resulting in chromatin modification and transcriptional gene silencing [142,145]. The functions of DCL2 and DCL4 are partially redundant and they are also involved in biogenesis of secondary or transitive siRNAs. The secondary siRNAs are processed from dsRNA produced by the action of RDR6 and SGS3 on single stranded RNA templates that are primed by primary siRNAs [139].
Depending on their site of origin, the siRNAs are classified as repeat-associated siRNA (rasiRNA), trans-acting siRNA (tasiRNA), natural-antisense siRNA (nat-siRNA), heterochromatic siRNA (hc-siRNA) and vi-siRNA (viral siRNAs).
(a) Ra-siRNAs are derived from TEs and repetitive DNAs [165] and mainly function in the silencing of retrotransposons and various abiotic stress factors, including drought [166]. Studies in maize and Arabidopsis have indicated complex feedback regulatory loops between rasiRNA and their target RNAs [166,167].
(b) TasiRNAs are derived by phased cleavage of dsRNA, which is produced after miRNA-mediated cleavage of the TAS gene-derived transcripts. In Arabidopsis, TAS1 and TAS2 transcripts are targeted by miR173, TAS3 transcripts are recognized by miR390 and TAS4 is targeted by miR828 [168]. They play a crucial regulatory role in development through post-transcriptional silencing [169,170]. TAS1, TAS2 and TAS3 were downregulated in response to drought and salinity stress [171]. The tasiRNA-ARF (auxin response factor) module is involved in regulating flower morphogenesis under drought and salt stress [172]. In Sorghum bicolor, two TAS3 gene homologs were identified to regulate the response to drought stress [170].
(c) Nat-siRNAs are a class of functional siRNAs, which originate from within the annealed regions of the natural antisense transcript (NAT) pairs [173]. Scientific evidence has indicated that NATs and Nat-siRNAs are involved in regulating various biological processes of plants and animals, such as phosphate homeostasis [174], stress response [175,176], chromatin remodeling and RNA editing [176,177,178,179].
(d) Hc-siRNAs are derived from heterochromatic intergenic regions including repeats and transposons [180]. The hc-siRNAs recognize the nascent Pol V-dependent transcript via base-pair complementarity and guide the DNA methylation and histone modification machinery to the loci for transcriptional gene silencing [180,181,182]. Their role has been reported in plants in response to several biotic [183,184,185,186] and abiotic stress factors [187,188].
(e) Vi-siRNAs are derived from dsRNA replicative intermediates of viruses to induce specific antiviral immunity [189]. They are generally processed from the sense strand of the viral genome [190]. Most of the vi-siRNAs have 5′ monophosphate, which indicates that vi-siRNAs can be produced by the viral RDR [191]. The role of vi-siRNAs has been reported in response to viral pathogen response in different plants, such as Arabidopsis [192], tomato [193], soybean [194], tobacco [195] and so on.

3.2. MicroRNAs

The miRNAs are processed from long primary transcripts that are transcribed from the genome. The steps in their biogenesis are complex and intricately regulated, as it involves the coordination of several proteins [196]. The primary transcripts (pri-miRNA) are sequentially processed by the DCL1 containing microprocessor complex into precursor miRNAs (pre-miRNA) and then into mature miRNAs. The steps in miRNA biogenesis are illustrated in Figure 2. Several other proteins, such as HYL and SE, are required for accurate DCL1 function [197]. The mature miRNA duplex is then methylated at the ends by HEN1 and transported to the cytoplasm, where it gets associated with the AGO containing RISC to form a functional complex, which can bring about transcript cleavage or suppress translation [198,199,200].
The miRNAs regulate various aspects of plant growth and development (Table 2) by regulating tissue or organ differentiation and development, shoot branching, root branching, lateral root development, panicle formation, flower development, seed development, primordial development, apical dominance, etc. [150,156,199,200,201,202,203,204,205,206]. The miRNAs also play an important role in promoting adaptation and tolerance to fluctuations in environmental conditions [207,208,209,210]. Moreover, miRNAs act in a coordinated manner by controlling the network of key genes, transcription factors and phytohormones [208,211,212,213,214].
It was seen that mutants in the miRNA biogenesis pathway exhibited an impaired response to abscisic acid (ABA), auxin and cytokinins [213,215], thus indicating the overlapping of regulatory hubs in plants. Later, it was shown that miR159 and miR164 modulated the levels of gibberellic acid (GA) and auxin, respectively [216,217,218]. The transcripts for auxin receptors, TIR1 (transport inhibitor response-1) and F-box protein 2 are targeted by miR393 [219,220,221,222]. The miRNA expression levels are also modulated by hormones, as exemplified by the downregulation of miR167 after treatment with ARF [223].
The first direct evidence that miRNAs are involved in plant stress responses came from the work of Jones-Rhoades and Bartel in the year 2004. Abiotic stress-regulated miRNAs were first reported in the model plant Arabidopsis thaliana [224] and, by now, the stress-responsive miRNAs have been reported in almost all plant species [225,226,227,228,229]. Functional studies have also supported this role for miRNAs. For example, overexpression of miR393 reduced plant growth in drought stress by downregulating the auxin signals [230]. There are also reports on the functional involvement of the miRNA passenger strand (miRNA*) in various responses. For example, miR169g* and miR172b* were downregulated in tomato leaves under varying phosphate deficient conditions [199,231].
Plant miRNA activity is precisely controlled by the regulation of expression of miRNA genes, processing of mature miRNAs and function of miRNAs. The first level of control involves the development and tissue-specific regulation of pri-miRNA transcription in response to hormonal and environmental cues by a variety of transcription factors [232,233], such as ARF, LFY, MYC2, etc. The transcripts of many of these transcription factors are regulated by the miRNAs, indicating the existence of complex cellular feedback loops [234].
The second level of control is achieved by regulating the processing or biogenesis of mature miRNAs. This is indicated by differences in the levels of pri/pre-miRNA and mature miRNAs and by the presence of DCL3 dependent 24 nt long miRNAs [169,235]. The regulation of DCL1 transcripts by miR162 and AGO1 transcripts by miR168 also adds to the spatial or temporal differences in miRNA activities [236]. In 2008, it was reported that overexpression of SINEs (short interspaced elements) resulted in phenotypes similar to that of miRNA-deficient mutants. Later, it was discovered that stem-loops of SINEs mimic the pre-miRNAs to bind and quench HYL1 [237].
The third level of control can be achieved by sorting miRNAs in different AGO complexes. According to the most popular hypothesis, the 5′-terminal nucleotide of miRNA guide strand determines the selection of the specific AGO containing RISC and, hence, the subsequent mode of action [238,239]. For instance, Uridine at the 5′ end supports preferential sorting with AGO1, whereas adenosine at the 5′ end favors sorting with AGO2 and AGO4 [240].
Table 2. List of conserved miRNAs and their key target genes that function in plant growth and development.
Table 2. List of conserved miRNAs and their key target genes that function in plant growth and development.
miRNAsTarget GeneFunctionsReferences
miR156/157SPLPhase transition from vegetative to reproductive phase; flowering[203]
miR159MYB familyDevelopment of male reproductive organs[216]
miR160ARF10, ARF16Controls root development and gravitropism[213]
miR165/166HD-ZIPIIILeaf development and polarity; lamina expansion[202]
miR166RDD1Grain size and weight[204]
miR167ARF10, ARF16, ARF17Floral patterning; controls anther and ovule development[241]
ARF6, ARF8Stamen and gynoecium and maturation; seed development[214]
miR168AGOsLeaf polarity[200]
miR169NF-YAFloral organ identity[242]
miR172AP2Floral patterning and floral organ development; regulates the inner whorl organ differentiation[243]
miR319TCPLeaf morphogenesis[226]
miR390ARF2, ARF3, ARF4Leaf development, adaxial identity of leaf blade, lateral organ development and leaf senescence[172]
miR394Leaf Curling Responsiveness (LCR)Regulation of leaf curling, shoot meristem differentiation and maintenance in abscisic acid–dependent manner[244]
miR396Growth Regulating Factors (GRFs)Adaxial–abaxial polarity of leaf and cell proliferation[245]
miR399PHO2Control of flowering time[227]
miR408PlantacyaninRoot development[246]
miR444MADS boxFloral patterning and development control[247]
miR824MADS-box geneFormation of stomatal complexes in meristems[248]
miR824AGL16Stomatal development
miR848IAA28Root development and lateral root development[249]
miR1218NAC3Organ separation[250]

4. Role of Long and Small Non-Coding RNAs during Drought Stress

The regulatory functions of plant lncRNAs and miRNAs in plant stress response have been comprehensively studied [67,106,109,251,252]. These two classes of ncRNAs also participate in response to water deficit and drought through complex cellular pathways involving chromatin modulation, target mimicry, transcriptional regulation, hormonal signaling and by directly regulating drought-responsive genes [57,117,158,230,253,254].

4.1. lncRNAs in Drought Stress

Genome-wide transcriptome studies have identified several drought-responsive lncRNAs in different plant species [96,97,115,255,256,257,258]. For example, studies on the identification of drought-responsive lncRNAs in grass families detected 664 potential candidates in maize [102], 98 in rice [119], 19 in foxtail millet [259] and 1597 in switchgrass [117]. The various reports on the identification of drought-responsive lncRNAs are presented in Table 3. Most lncRNAs regulate the drought response by acting on genes participating in ethylene and ABA synthesis or signaling, calcium signaling, starch and sucrose synthesis and several other metabolic processes.
The studies on cassava identified 51 drought-specific differentially expressed lncRNAs and qRT-PCR validation of selected molecules among them revealed the up- regulation of lincRNA101, lincRNA391 and lincRNA356. Other lncRNAs, such as lincRNA64, lincRNA350, lincRNA182 and lincRNA392, were downregulated under drought stress. The TCONS_00060863 and TCONS_00097416 lncRNAs were shown to regulate ABA and ethylene signaling pathways, respectively, under drought stress [116]. In switchgrass, drought stress upregulated the lncRNAs XLOC_053020, XLOC_014465 and XLOC_033252 to control ABA synthesis, XLOC_074836 to regulate ethylene signaling and XLOC_005809 to control trehalose phosphate synthase [117].
In rice, 98 drought-responsive NAT-lncRNAs were identified by using RNA-Seq analysis. These included two important drought-responsive lncRNAs viz NAT Os02g0250700-01 and NAT Os02g0180800-01, which targets the late embryogenesis abundant protein and cinnamoyl CoA reductase genes, respectively [119]. Studies in maize identified that the lncRNAs expressing at the R1 stage (silking stage) had a critical role in drought stress tolerance [122]. The possible role of lncRNAs as positive regulators of drought stress tolerance in Arabidopsis was identified with the discovery of a novel nucleus localized 755 nt long drought-induced lincRNA (DRIR). The DRIR overexpressing Arabidopsis lines had higher drought tolerance than wild-type seedlings [113]. This lincRNA was a nuclear-localized and controlled transcription of several drought stress-responsive genes, including ABA signaling genes (ABI5, P5CS1, RD29A and RD29B), aquaporin genes (NIP1 and TIP4), annexin gene (ANNAT7), fucosyltransferase4 (FUT4) gene and transcription factor genes (NAC3 and WARKY8) [113].

4.2. miRNAs in Drought Stress

Several studies have also shown the role of miRNAs in regulating plant response to drought stress (Table 4). A number of miRNAs, such as miR156, miR158, miR159, miR165, miR167, miR168, miR169, miR171, miR319, miR393, miR394 and miR396, were upregulated in response to drought stress in Arabidopsis [268]. The upregulation of Arabidopsis miR393, miR319 and miR397 in response to dehydration was reported earlier [224]. In drought-stressed rice seedlings, genome-wide analysis was carried out across different developmental stages, from tillering to inflorescence formation, using a microarray platform [269]. This analysis identified 30 miRNA gene families that were differentially regulated. Among these, 16 miRNA families, namely miR156, miR159, miR168, miR170, miR171, miR172, miR319, miR396, miR397, miR408, miR529, miR896, miR1030, miR1035, miR1050, miR1088 and miR1126, were significantly downregulated. Meanwhile, 14 miRNAs, namely miR159, miR169, miR171, miR319, miR395, miR474, miR845, miR851, miR854, miR896, miR901, miR903, miR1026 and miR1125, were significantly upregulated under drought stress. Another report showed that miR164c, miR319b and miR1861d were downregulated, while miR166h, miR172d and miR408 were upregulated under drought stress in rice [270]. In rice, pre-miRNA expression profiling indicated that miR171f was involved in the progression of rice root development and growth and response to drought stress [119]. In a different study, it was shown that miR169g was strongly upregulated and miR393 was transiently induced by drought stress in rice [271]. Inoculation of rice plants with P. indica caused upregulation of miR396, and this resulted in the downregulation of growth-regulating factor (GRF), which lowered the rate of transpiration and enabled the plants to tolerate drought [272].
In Medicago truncatula, miR169 was downregulated only in the roots while miR398a,b and miR408 were strongly upregulated in both shoots and roots under drought stress [299]. In a Populus plant, miR156, miR159, miR171, miR319, miR395 and miR474 were upregulated in response to drought stress [300]. In Populus tomentosa, about 152 conserved miRNAs were identified and the expression of 17 conserved and nine novel miRNAs was investigated in response to drought stress [301]. In Vitis vinifera, 12 novel and species-specific miRNA candidates were reported in response to drought stress. Moreover, 70 conserved miRNAs were identified and 28 novel miRNAs were predicted in a drought-resistant grapevine [302].
Differential regulation of miRNAs in response to drought stress has been well studied in maize [253,286,288,302,303,304]. For example, miR398 was upregulated after treatment with polyethylene glycol and downregulated under soil drought [305]. The downregulation of miR167 during drought stress upregulated its target PLD (Phospholipase D), which is involved in controlling ABA response and stomatal movement [21]. Similarly, the downregulation of miR159 in drought triggered the expression of HD-ZIP, ARF and GA-MYB transcription factors, which contributed to greater adventitious and lateral root formation. Moreover, miR474 was upregulated in drought to inhibit proline dehydrogenase (PDH) [306], while miR827 was upregulated during drought stress to act on NADP-binding and SPX (SYG1/Pho81/XPR) transcripts to activate stress signal transduction pathways [305].
It was shown that miR156 interacts with the ABA-dependent strigolactone signaling pathways in tomatoes under drought stress. The study identified miR156 as a mediator of stomatal movements and the findings indicated a cause–effect link between miR156 accumulation and regulation of water relations and stomatal functioning [307]. In sugarcane, miR169* was shown to target various transcripts such as Elongation Factor 1-alpha (EF 1α) in response to water depletion [297]. It was identified that miR529, miR535 and miR156 regulate transcripts of Squamosa-promoter binding protein-like (SPL) to control organ development and morphogenesis during stress. Similarly, miR159 targets MYB33 and miR172 targets AP2 (Apetala 2) to regulate plant development in response to drought stress [152].
Functional studies have provided an insight into the role of miRNAs in regulating the response to drought stress. In Arabidopsis, overexpression of ath-miR169a [308] and gma-169c [309], which targets the Nuclear factor Y-A (NFY-A) resulted in increased drought stress sensitivity. In contrast, similar studies in tomatoes have reported that plants overexpressing sly-miR169c show negative regulation of stomatal movement, reduced leaf water loss and transpiration rate, and improved drought tolerance [310]. Overexpression of Osa-miR393, which targets the auxin-responsive OsTIR1 and OsAFB2, lowered the tolerance of rice plants to salt and drought stress [230]. In another example, Osa-miR319 overexpression in creeping bentgrass led to greater tolerance to salinity and drought, by decreasing the expression of its putative target genes: AsPCF5, AsPCF6, AsPCF8, AsTCP14 and AsNAC60 [283]. In Populus ussuriensis, overexpression of Pu-miR172d significantly decreased stomatal density by directly repressing the expression of PuGTL1 and PuSDD1. This resulted in increased water use efficiency and drought tolerance by reducing net photosynthetic rate, stomatal conductance and rate of transpiration [311]. This study showed that Pu-miR172d-PuGTL1-PuSDD1 module played an important role in stomatal differentiation and acted as a potential target for creating drought-tolerant plants. Similar studies in other plants showed that overexpression of miR156 in Alfalfa [312], miR408 in chickpea [282], Osa-miR319a in creeping bentgrass [283], miR169 in tomato [310] and miR159 in potato [313] resulted in enhanced drought stress tolerance. Recently, it was shown that miR535 overexpressing and CRISPR/Cas9 knockout rice showed enhanced stress tolerance when tested in presence of sodium chloride, polyethylene glycol, abscisic acid and dehydration stresses parameters [314].

4.3. Interaction between Long and Small ncRNAs in Drought Stress

Studies on the mechanism of action of lncRNAs have revealed their complex interaction with the small ncRNAs. Together, they form complex regulatory hubs for controlling various drought responsive pathways at the transcription, post-transcription and epigenome levels. Studies in Cassava showed that 11 drought-specific differentially expressed lncRNAs acted as target mimics for miR156, miR164, miR169 and miR172 [116]. Under drought stress, lincRNA340 acts as a target mimic of miR169 to enhance the expression of its target gene NFY [115]. The lncRNA, TCONS_00068353 acted as a target mimic for miR156k and miR172c to control several abiotic stress-responsive genes [116].
Many plant TEs contain stress-responsive cis-acting elements and produce lncRNAs in response to specific stress [315], and many of these are possible sources of small ncRNAs that can regulate both TE and non-TE transcripts based on sequence complementarity. In maize, eight drought-responsive lncRNAs acted as precursors of miRNAs [102]. It was shown that TE-derived epigenetically activated siRNAs (easiRNAs) participated in transcriptional silencing. In rice, TE-siRNA815 could induce a de novo DNA methylation process via the RdDM pathway [316]. The stress-downregulated Osa-miR820 originates from CACTA-TE [317] and targets de novo DNA methyltransferase (DRM2) transcripts. Overexpression of Osa-miR820 enhanced salt tolerance in rice plants [318]. It was also shown that ZmNAC111 expression is repressed by miniature inverted-repeat transposable element (MITE) through RdDM and H3K9 dimethylation during drought tolerance [121]. Overexpression of the ZmaNAC111 gene boosted drought tolerance in maize seedlings [319]. This phenomenon has unveiled functional crosstalk between small ncRNAs and the TEs, indicating that novel stress-responsive regulatory networks may be operative in plants [70,320].

5. Conclusions and Perspectives

The steadily increasing world population has challenged the agricultural sector to produce a substantial amount of crops. However, crop productivity all over the world is anticipating challenges by the ever-changing climate, variable weather conditions and environmental stresses. The limited availability of water and global warming has increased the incidence of drought, making it a major contributor to agricultural losses. To tackle this problem and produce enough food to feed the growing world population, it is important to generate crops that can survive underwater limiting conditions and can evade drought stress. This process can be aided by a thorough understanding of plant responses to water deficit and drought stress.
The exciting discovery of RNA-mediated gene silencing has highlighted the role of long and small ncRNAs in maintaining the homeostasis of gene expression. Advances in RNA-Seq analysis, computational analysis and functional genomic studies have enabled the discovery of several long and small ncRNAs and facilitated the understanding of their regulations. However, their functional characterization and annotation are limited to select plant species. Though the studies on ncRNAs are still in their infancy, their discovery has unraveled a novel mechanism of gene regulation. The small ncRNAs, such as miRNAs, regulate various aspects of plant biology, while the long ncRNAs have a role in regulating the miRNAs by acting as target mimics, sponges or decoys. The ncRNAs normally work in highly complex and intricately connected networks to regulate plant growth and development. The small ncRNAs belong to large families where specific members may be associated with a definite development stage or response.
In the last few years, substantial progress has been made in deciphering the mechanisms of ncRNAs. It has been shown that the small ncRNAs have the ability to move systemically within the plant’s vasculature or locally from one cell to another. This was demonstrated by micro-grafting miR399 overexpressing Arabidopsis shoots on wild-type roots. The chimeric plants accumulated very high levels of mature miR399 species in the wild-type roots, where the primary transcripts were virtually absent. The chimeric plants showed downregulation of PHO2 in the wild-type roots and Pi accumulation in the shoots. This indicated a role for the miRNAs in long-distance signaling for maintaining nutrient balance [321] The miR399 could not only move through the phloem tissues, but the transported molecules retained their biological activity in the recipient tissues. In another report, both ath-miR399d and its star sequence were identified as the mobile elements. During phosphate starvation, translocation by miR827 and miR2111a between shoots and roots was also demonstrated [322]. The long-distance mobility of miRNA species reflects on their potential in root–shoot communications during stress responses [323,324]. The miRNA shuttles may be operative in response to drought stress, as well. Indications towards this come from studies on gma-miR172, which is induced under salt and drought treatments. The miR172 cleaves/inhibits the transcript encoding AP2/EREBP-type transcription factor (SSAC1) to relieve inhibition of thiamine biosynthesis gene (THI1) that encodes a positive regulator of salt stress tolerance [325].
There is no doubt that the ncRNAs play a crucial role in regulating plant growth and stress responses. Many important issues remain to be answered, such as how do the ncRNAs move from the cells where they are produced and move into the recipient cells? How are the ncRNAs transported, and in what way are they protected from nucleolytic degradation during movement? Are there specific proteins or chemical tags which help them in such transfers?
The information related to long ncRNAs is still emerging, and there is still a lot more to discover with respect to their functions and regulations. Dedicated and systematic efforts will be required to understand how the ncRNAs networks operate in different crop plants over spatiotemporal boundaries and identify their association with response to drought and related stresses. It will be a lot more exciting to understand if they have any role in influencing inter-organ communications and stress responses. In this context, genetic screens and transgenic approaches will aid in unraveling their novel functionalities and features. It is envisaged that such studies will open up opportunities for designing efficient strategies for development of stress-tolerant crops.

Author Contributions

T.A.G. wrote the manuscript and drew the figures. N.S.-M. edited and improved the manuscript. Both authors have approved the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by grants from ICGEB. TAG acknowledges the Arturo Falaschi ICGEB Fellowships Programme.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Salehi-Lisar, S.Y.; Bakhshayeshan-Agdam, H. Drought stress tolerance in plants. In Drought Stress in Plants: Causes, Consequences and Tolerance; Hossain, M.A., Wani, S., Bhattacharjee, S., Burritt, D., Tran, L.S., Eds.; Springer: Cham, Switzerland, 2016; Volume 1. [Google Scholar] [CrossRef]
  2. Bhargava, S.; Sawant, K. Drought stress adaptation: Metabolic adjustment and regulation of gene expression. Plant Breed. 2013, 132, 21–32. [Google Scholar] [CrossRef]
  3. Krannich, C.T.; Maletzki, L.; Kurowsky, C.; Horn, R. Network candidate genes in breeding for drought tolerant crops. Int. J. Mol. Sci. 2015, 16, 16378–16400. [Google Scholar] [CrossRef]
  4. Wang, C.; Linderholm, H.W.; Song, Y.; Wang, F.; Liu, Y.; Tian, J.; Xu, J.; Song, Y.; Ren, G. Impacts of drought on maize and soybean production in northeast China during the past five decades. Int. J. Environ. Res. Public Health 2020, 17, 2459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Arbona, V.; Manzi, M.; de Ollas, C.; Gómez-Cadenas, A. Metabolomics as a tool to investigate abiotic stress tolerance in plants. Int. J. Mol. Sci. 2013, 14, 4885–4911. [Google Scholar] [CrossRef] [PubMed]
  6. Todaka, D.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Recent advances in the dissection of drought-stress regulatory networks and strategies for development of drought-tolerant transgenic rice plants. Front. Plant Sci. 2015, 6, 1–20. [Google Scholar] [CrossRef] [Green Version]
  7. Tigchelaar, M.; Battisti, D.S.; Naylor, R.L.; Ray, D.K. Future warming increases probability of globally synchronized maize production shocks. Proc. Natl. Acad. Sci. USA 2018, 115, 6644–6649. [Google Scholar] [CrossRef] [Green Version]
  8. Anupama, A.; Bhugra, S.; Lall, B.; Chaudhury, S.; Chugh, A. Assessing the correlation of genotypic and phenotypic responses of indica rice varieties under drought stress. Plant Physiol. Biochem. 2018, 127, 343–354. [Google Scholar] [CrossRef]
  9. Loka, D.A.; Oosterhuis, D.M.; Ritchie, G.L. Water-Deficit Stress in Cotton. Stress Physiol. Cott. 2011, 7, 37–72. [Google Scholar]
  10. Khan, A.; Pan, X.; Najeeb, U.; Tan, D.K.Y.; Fahad, S.; Zahoor, R.; Luo, H. Coping with drought: Stress and adaptive mechanisms and management through cultural and molecular alternatives in cotton as vital constituents for plant stress resilience and fitness. Biol. Res. 2018, 51, 1–17. [Google Scholar] [CrossRef]
  11. Manavalan, L.P.; Guttikonda, S.K.; Tran, L.S.P.; Nguyen, H.T. Physiological and molecular approaches to improve drought resistance in soybean. Plant Cell Physiol. 2009, 50, 1260–1276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Golldack, D.; Luking, I.; Yang, O. Plant tolerance to drought and salinity: Stress regulating transcription factors and their functional significance in the cellular transcriptional network. Plant Cell Rep. 2011, 30, 1383–1391. [Google Scholar] [CrossRef]
  13. Fang, Y.; Liao, K.; Du, H.; Xu, Y.; Song, H.; Li, X.; Xiong, L. A stress-responsive NAC transcription factor SNAC3 confers heat and drought tolerance through modulation of reactive oxygen species in rice. J. Exp. Bot. 2015, 66, 6803–6817. [Google Scholar] [CrossRef] [Green Version]
  14. Feia, X.; Lia, J.; Konga, L.; Hua, H.; Tiana, J.; Liua, Y.; Weia, A. MiRNAs and their target genes regulate the antioxidant system of Zanthoxylum bungeanum under drought stress. Plant Phys. Biochem. 2020, 150, 196–203. [Google Scholar] [CrossRef]
  15. Martin-StPaul, N.; Delzon, S.; Cochard, H. Plant resistance to drought depends on timely stomatal closure. Ecol. Lett. 2017, 20, 1437–1447. [Google Scholar] [CrossRef] [PubMed]
  16. Luo, L.J. Breeding for water-saving and drought-resistance rice (WDR) in China. J. Exp. Bot. 2010, 61, 3509–3517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Isah, T. Stress and defense responses in plant secondary metabolites production. Biol. Res. 2019, 52, 39. [Google Scholar] [CrossRef] [Green Version]
  18. Gill, S.S.; Tuteja, N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiol. Biochem. 2010, 48, 909–930. [Google Scholar] [CrossRef] [PubMed]
  19. Lima, C.S.; Ferreira-Silvab, S.L.; Carvalhoa, F.E.L.; Netoc, M.C.L.; Aragaod, R.M.; Silvae, E.N.; Sousa, R.M.J.; Silveira, G. Antioxidant protection and PSII regulation mitigate photo-oxidative stress induced by drought followed by high light in cashew plants. Environ. Exp. Bot. 2018, 149, 59–69. [Google Scholar] [CrossRef] [Green Version]
  20. Xu, J.; Hou, Q.M.; Khare, T.; Verma, S.K.; Kumar, V. Exploring miRNAs for developing climate-resilient crops: A perspective review. Sci. Tot. Environ. 2019, 653, 91–104. [Google Scholar] [CrossRef]
  21. Ding, Y.; Tao, Y.; Zhu, C. Emerging roles of microRNAs in the mediation of drought stress response in plants. J. Exp. Bot. 2013, 64, 3077–3086. [Google Scholar] [CrossRef]
  22. Koroban, N.V.; Kudryavtseva, A.V.; Krasnov, G.S.; Sadritdinova, A.F.; Fedorova, M.S.; Snezhkina, A.V.; Bolsheva, N.L.; Muravenko, O.V.; Dmitriev, A.A.; Melnikova, N.V. The role of microRNAs in abiotic stress response in plants. Mol. Biol. 2016, 50, 337–343. [Google Scholar] [CrossRef]
  23. Shengli, L.; Yanjie, N.; Qian, H.E.; Jinyu, W.; Yuzhen, C.; Cunfu, L.U. Genome-wide Identification of microRNAs that respond to drought stress in seedlings of tertiary relict Ammopiptanthus mongolicus. Hortic. Plant J. 2017, 3, 209–218. [Google Scholar] [CrossRef]
  24. Yu, Y.; Zhang, Y.; Chen, X.; Chen, Y. Plant noncoding RNAs: Hidden players in development and stress responses. Annu. Rev. Cell Dev. Biol. 2019, 35, 407–431. [Google Scholar] [CrossRef] [PubMed]
  25. Hu, J.; Zeng, T.; Xia, Q.; Qian, Q.; Yang, C.; Ding, Y.; Chen, L.; Wang, W. Unravelling miRNA regulation in yield of rice (Oryza sativa) based on differential network model. Sci. Rep. 2018, 8, 1–10. [Google Scholar] [CrossRef]
  26. Goodwin, S.; McPherson, J.D.; McCombie, W.R. Coming of age: Ten years of next-generation sequencing technologies. Nat. Rev. Genet. 2016, 17, 333–351. [Google Scholar] [CrossRef] [PubMed]
  27. Fracasso, A.; Trindade, L.M.; Amaducci, S. Drought stress tolerance strategies revealed by RNA-Seq in two sorghum genotypes with contrasting WUE. BMC Plant Biol. 2016, 16, 1–18. [Google Scholar] [CrossRef] [PubMed]
  28. Zhong, S.; Fei, Z.; Chen, Y.R.; Huang, M.; Vrebalov, J.; Mcquinn, R.; Gapper, N.; Liu, B.; Xiang, J.; Shao, Y.; et al. Single-base resolution methylomes of tomato fruit development reveal epigenome modifications associated with ripening. Nat. Biotechnol. 2013, 31, 154–159. [Google Scholar] [CrossRef]
  29. Mofatto, L.S.; Carneiro, F.A.; Vieira, N.G.; Duarte, K.E.; Vidal, R.O.; Alekcevetch, J.C.; Cotta, M.G.; Verdeil, J.-L.; Lapeyre-Montes, F.; Lartaud, M.; et al. Identification of candidate genes for drought tolerance in coffee by high-throughput sequencing in the shoot apex of different Coffea arabica cultivars. BMC Plant Biol. 2016, 16, 94. [Google Scholar] [CrossRef] [Green Version]
  30. Hu, W.; Xia, Z.Q.; Yan, Y.; Ding, Z.H.; Tie, W.W.; Wang, L.Z.; Zou, M.; Wei, Y.; Lu, C.; Hou, X.; et al. Genome-wide gene phylogeny of CIPK family in cassava and expression analysis of partial drought-induced genes. Front. Plant Sci. 2015, 6, 1–15. [Google Scholar] [CrossRef] [Green Version]
  31. Brasileiro, A.C.; Morgante, C.V.; Araujo, A.C.; Leal-Bertioli, S.C.; Silva, A.K.; Martins, A.C.; Vinson, C.C.; Santos, C.M.R.; Bonfim, O.; Togawa, R.C.; et al. Transcriptome profiling of wild Arachis from water-limited environments uncovers drought tolerance candidate genes. Plant Mol. Biol. Rep. 2015, 33, 1876–1892. [Google Scholar] [CrossRef]
  32. Tang, S.; Liang, H.; Yan, D.; Zhao, Y.; Han, X.; Carlson, J.E.; Xia, X.; Yin, W. Populus euphratica: The transcriptomic response to drought stress. Plant Mol. Biol. 2013, 83, 539–557. [Google Scholar] [CrossRef]
  33. Yates, S.A.; Swain, M.T.; Hegarty, M.J.; Chernukin, I.; Lowe, M.; Allison, G.G.; Ruttink, T.; Abberton, M.T.; Jenkins, G.; Skot, L. De novo assembly of red clover transcriptome based on RNA-Seq data provides insight into drought response, gene discovery and marker identification. BMC Genom. 2014, 15, 453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hua, Y.; Zhang, C.; Shi, W.; Chen, H. High-throughput sequencing reveals microRNAs and their targets in response to drought stress in wheat (Triticum aestivum L.). Biotechnol. Biotechnol. Equip. 2019, 33, 465–471. [Google Scholar] [CrossRef] [Green Version]
  35. Mutum, R.D.; Kumar, S.; Balyan, S.; Kansal, S.; Mathur, S.; Raghuvanshi, S. Identification of novel miRNAs from drought tolerant rice variety Nagina 22. Sci. Rep. 2016, 6, 1–15. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, X.; Elling, A.A.; Li, X.; Li, N.; Peng, Z.; He, G.; Sun, H.; Qi, Y.; Liu, X.S.; Denga, X.W. Genome-wide and organ-specific landscapes of epigenetic modifications and their relationships to mRNA and small RNA transcriptomes in maize. Plant Cell 2009, 21, 1053–1069. [Google Scholar] [CrossRef] [Green Version]
  37. Xu, J.; Yuan, Y.; Xu, Y.; Zhang, G.; Guo, X.; Wu, F.; Wang, Q.; Rong, T.; Pan, G.; Cao, M.; et al. Identification of candidate genes for drought tolerance by whole-genome resequencing in maize. BMC Plant Biol. 2014, 14, 83. [Google Scholar] [CrossRef] [Green Version]
  38. Espinosa, J.M. On the Origin of lncRNAs: Missing link found. Trends Genet. 2017, 33, 660–662. [Google Scholar] [CrossRef]
  39. Wu, H.; Yang, L.; Chen, L.L. The diversity of long noncoding RNAs and their generation. Trends Genet. 2017, 33, 540–552. [Google Scholar] [CrossRef]
  40. Zampetaki, A.; Albrecht, A.; Steinhofel, K. Long non-coding RNA structure and function: Is there a link? Front. Physiol. 2018, 9, 1–8. [Google Scholar] [CrossRef]
  41. Gloss, B.S.; Dinger, M.E. The specificity of long noncoding RNA expression. Biochim. Biophys. Acta 2016, 1859, 16–22. [Google Scholar] [CrossRef] [PubMed]
  42. Hirose, T.; Nakagawa, S. Clues to long noncoding RNA taxonomy. Biochim. Biophys. Acta 2016, 1859, 1–2. [Google Scholar] [CrossRef]
  43. Liu, J.; Wang, H.; Chua, N.H. Long noncoding RNA transcriptome of plants. Plant Biotechnol. J. 2015, 13, 319–328. [Google Scholar] [CrossRef] [PubMed]
  44. Derrien, T.; Johnson, R.; Bussotti, G.; Tanzer, A.; Djebali, S.; Tilgner, H.; Guernec, G.; Martin, D.; Merkel, A.; Knowles, D.G.; et al. The GENCODE v7 catalog of human long noncoding RNAs: Analysis of their gene structure, evolution and expression. Genome Res. 2012, 22, 1775–1789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Bai, Y.; Dai, X.; Harrison, A.P.; Chen, M. RNA regulatory networks in animals and plants: A long noncoding RNA perspective. Brief. Funct. Genom. 2015, 14, 91–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Kazemzadeh, M.; Safaralizadeh, R.; Orang, A.V. LncRNAs: Emerging players in gene regulation and disease pathogenesis. J. Genet. 2015, 94, 771–784. [Google Scholar] [CrossRef]
  47. Wang, H.L.V.; Chekanova, J.A. Long noncoding RNAs in plants. Adv. Exp. Med. Biol. 2017, 1008, 33–154. [Google Scholar] [CrossRef]
  48. Liu, X.; Hao, L.; Li, D.; Zhu, L.; Hu, S. Long non-coding RNAs and their biological roles in plants. Genom. Proteom. Bioinform. 2015, 13, 137–147. [Google Scholar] [CrossRef] [Green Version]
  49. Moore, M.J.; Proudfoot, N.J. Pre-mRNA processing reaches back to transcription and ahead to translation. Cell 2009, 136, 688–700. [Google Scholar] [CrossRef] [Green Version]
  50. Bentley, D.L. Coupling mRNA processing with transcription in time and space. Nat. Rev. Genet. 2014, 15, 163–175. [Google Scholar] [CrossRef] [Green Version]
  51. Wierzbicki, A.T.; Haag, J.R.; Pikaard, C.S. Noncoding transcription by RNA Polymerase Pol IVb/Pol V mediates transcriptional silencing of overlapping and adjacent genes. Cell 2008, 135, 635–648. [Google Scholar] [CrossRef] [Green Version]
  52. Andersson, R.; Gebhard, C.; Miguel-Escalada, I.; Hoof, I.; Bornholdt, J.; Boyd, M.; Chen, Y.; Zhao, X.; Schmidl, C.; Suzuki, T.; et al. An atlas of active enhancers across human cell types and tissues. Nature 2014, 507, 455–461. [Google Scholar] [CrossRef] [PubMed]
  53. Wilusz, J.E.; Freier, S.M.; Spector, D.L. 3′ end processing of a long nuclear-retained noncoding RNA yields a tRNA-like cytoplasmic RNA. Cell 2008, 135, 919–932. [Google Scholar] [CrossRef] [Green Version]
  54. Zhang, X.O.; Yin, Q.F.; Wang, H.B.; Zhang, Y.; Chen, T.; Zheng, P.; Lu, X.; Chen, L.L.; Yang, L. Species-specific alternative splicing leads to unique expression of sno-lncRNAs. BMC Genom. 2014, 15, 287. [Google Scholar] [CrossRef] [Green Version]
  55. Sibley, C.R.; Blazquez, L.; Ule, J. Lessons from non-canonical splicing. Nat. Rev. Genet. 2016, 17, 407–421. [Google Scholar] [CrossRef] [PubMed]
  56. Lucero, L.; Ferrero, L.; Fonouni-Farde, C.; Ariel, F. Functional classification of plant long noncoding RNAs: A transcript is known by the company it keeps. New Phytol. 2021, 229, 1251–1260. [Google Scholar] [CrossRef] [PubMed]
  57. Jha, U.C.; Nayyar, H.; Jha, R.; Khurshid, M.; Zhou, M.; Mantri, N.; Siddique, K.H.M. Long non-coding RNAs: Emerging players regulating plant abiotic stress response and adaptation. BMC Plant Biol. 2020, 20, 1–20. [Google Scholar] [CrossRef]
  58. Statello, L.; Guo, C.J.; Chen, L.L.; Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef]
  59. Fan, C.; Hao, Z.; Yan, J.; Li, G. Genome-wide identification and functional analysis of lincRNAs acting as miRNA targets or decoys in maize. BMC Genom. 2015, 16, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. De Santa, F.; Barozzi, I.; Mietton, F.; Ghisletti, S.; Polletti, S.; Tusi, B.K.; Muller, H.; Ragoussis, J.; Wei, C.L.; Natoli, G. A large fraction of extragenic RNA Pol II transcription sites overlap enhancers. PLoS Biol. 2010, 8, e1000384. [Google Scholar] [CrossRef] [Green Version]
  61. Mattick, J.S.; Rinn, J.L. Discovery and annotation of long noncoding RNAs. Nat. Struct. Mol. Biol. 2015, 22, 5–7. [Google Scholar] [CrossRef]
  62. St. Laurent, G.; Wahlestedt, C.; Kapranov, P. The Landscape of long noncoding RNA classification. Trends Genet. 2015, 31, 239–251. [Google Scholar] [CrossRef] [Green Version]
  63. Preker, P.; Almvig, K.; Christensen, M.S.; Valen, E.; Mapendano, C.K.; Sandelin, A.; Jensen, T.H. PROMoter uPstream Transcripts share characteristics with mRNAs and are produced upstream of all three major types of mammalian promoters. Nucleic Acids Res. 2011, 39, 7179–7193. [Google Scholar] [CrossRef]
  64. Li, L.; Eichten, S.R.; Shimizu, R.; Petsch, K.; Yeh, C.T.; Wu, W.; Chettoor, A.M.; Givan, S.A.; Cole, R.A.; Fowler, J.E.; et al. Genome-wide discovery and characterization of maize long non-coding RNAs. Genome Biol. 2014, 15, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ali, A.; Han, K.; Liang, P. Role of transposable elements in gene regulation in the human genome. Life 2021, 11, 118. [Google Scholar] [CrossRef] [PubMed]
  66. Burgess, D.; Li, H.; Zhao, M.; Kim, S.Y.; Lisch, D. Silencing of mutator elements in maize involves distinct populations of small RNAs and distinct patterns of DNA methylation. Genetics 2020, 215, 379–391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Chen, L.; Zhu, Q.H.; Kaufmann, K. Long non-coding RNAs in plants: Emerging modulators of gene activity in development and stress responses. Planta 2020, 252, 1–14. [Google Scholar] [CrossRef]
  68. Wang, X.; Ai, G.; Zhang, C.; Cui, L.; Wang, J.; Li, H.; Zhang, J.; Ye, Z. Expression and diversification analysis reveals transposable elements play important roles in the origin of Lycopersicon-specific lncRNAs in tomato. New Phytol. 2016, 209, 1442–1455. [Google Scholar] [CrossRef]
  69. Wang, D.; Qu, Z.; Yang, L.; Zhang, Q.; Liu, Z.H.; Do, T.; Adelson, D.L.; Wang, Z.Y.; Searle, I.; Zhu, J.K. Transposable elements (TEs) contribute to stress-related long intergenic noncoding RNAs in plants. Plant J. 2017, 90, 133–146. [Google Scholar] [CrossRef] [Green Version]
  70. Cho, J. Transposon-derived non-coding RNAs and their function in plants. Front. Plant Sci. 2018, 9, 1–6. [Google Scholar] [CrossRef] [PubMed]
  71. McClintock, B. Chromosome organization and genic expression. Cold Spring Harb. Symp. Quant. Biol. 1951, 16, 13–47. [Google Scholar] [CrossRef]
  72. Bourgeois, Y.; Boissinot, S. On the population dynamics of junk: A review on the population genomics of transposable elements. Genes 2019, 10, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Pedersen, I.M.; Zisoulis, D.G. Transposable elements and miRNA: Regulation of genomic stability and plasticity. Mob. Genet. Elem. 2016, 6, 1–7. [Google Scholar] [CrossRef] [Green Version]
  74. Schnable, P.S.; Ware, D.; Fulton, R.S.; Stein, J.C.; Wei, F.; Pasternak, S.; Liang, C.; Zhang, J.; Fulton, L.; Graves, T.A.; et al. The B73 maize genome: Complexity, diversity and dynamics. Science 2009, 326, 1112–1115. [Google Scholar] [CrossRef] [Green Version]
  75. Lv, Y.; Hu, F.; Zhou, Y.; Wu, F.; Gaut, B.S. Maize transposable elements contribute to long non-coding RNAs that are regulatory hubs for abiotic stress response. BMC Genom. 2019, 20, 1–17. [Google Scholar] [CrossRef]
  76. Noshay, J.M.; Anderson, S.N.; Zhou, P.; Ji, L.; Ricci, W.; Lu, Z.; Stitzer, M.C.; Crisp, P.A.; Hirsch, C.N.; Zhang, X.; et al. Monitoring the interplay between transposable element families and DNA methylation in maize. PLoS Genet. 2019, 15, e1008291. [Google Scholar] [CrossRef] [Green Version]
  77. Ma, L.; Bajic, V.B.; Zhang, Z. On the classification of long non-coding RNAs. RNA Biol. 2013, 10, 924–933. [Google Scholar] [CrossRef]
  78. Prensner, J.R.; Iyer, M.K.; Balbin, O.A.; Dhanasekaran, S.M.; Cao, Q.; Brenner, J.C.; Laxman, B.; Asangani, I.A.; Grasso, C.S.; Kominsky, H.D.; et al. Transcriptome sequencing across a prostate cancer cohort identifies PCAT-1, an unannotated lincRNA implicated in disease progression. Nat. Biotechnol. 2011, 29, 742–749. [Google Scholar] [CrossRef] [Green Version]
  79. Tani, H.; Imamachi, N.; Mizutani, R.; Imamura, K.; Kwon, Y.; Miyazaki, S.; Maekawa, S.; Suzuki, Y.; Akimitsu, N. Genome-wide analysis of long noncoding RNA turnover. Methods Mol. Biol. 2015, 1262, 305–320. [Google Scholar] [CrossRef]
  80. Louro, R.; Smirnova, A.S.; Verjovski-Almeida, S. Long intronic noncoding RNA transcription: Expression noise or expression choice? Genomics 2009, 93, 291–298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Kim, Y.K.; Kim, V.N. Processing of intronic microRNAs. EMBO J. 2007, 26, 775–783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Zhou, H.; Lin, K. Excess of microRNAs in large and very 5′ biased introns. Biochem. Biophys. Res Commun. 2008, 368, 709–715. [Google Scholar] [CrossRef]
  83. Wight, M.; Werner, A. The functions of natural antisense transcripts. Essays Biochem. 2013, 54, 91–101. [Google Scholar] [CrossRef] [PubMed]
  84. Yu, X.; Yang, J.; Li, X.; Liu, X.; Sun, C.; Wu, F.; He, Y. Global analysis of cis-natural antisense transcripts and their heat-responsive nat-siRNAs in Brassica rapa. BMC Plant Biol. 2013, 13, 1. [Google Scholar] [CrossRef] [Green Version]
  85. Lavorgna, G.; Dahary, D.; Lehner, B.; Sorek, R.; Sanderson, C.M.; Casari, G. In search of antisense. Trends Biochem. Sci. 2004, 29, 88–94. [Google Scholar] [CrossRef] [PubMed]
  86. Li, X.; Yang, L.; Chen, L.L. The biogenesis, functions and challenges of circular RNAs. Mol. Cell 2018, 71, 428–442. [Google Scholar] [CrossRef] [Green Version]
  87. Sanger, H.L.; Klotz, G.; Riesner, D.; Gross, H.J.; Kleinschmidt, A.K. Viroids are single stranded covalently closed circular RNA molecules existing as highly base paired rod like structures. Proc. Natl. Acad. Sci. USA 1976, 73, 3852–3856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Zhang, Y.; Zhang, X.O.; Chen, T.; Xiang, J.F.; Yin, Q.F.; Xing, Y.H.; Zhu, S.; Yang, L.; Chen, L.L. Circular intronic long noncoding RNAs. Mol. Cell 2013, 51, 792–806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Chu, Q.; Bai, P.; Zhu, X.; Zhang, X.; Mao, L.; Zhu, Q.H.; Fan, L.; Ye, C.Y. Characteristics of plant circular RNAs. Brief. Bioinform. 2018, 21, 135–143. [Google Scholar] [CrossRef]
  90. Liu, X.; Wang, X.; Li, J.; Hu, S.; Deng, Y.; Yin, H.; Bao, X.; Zhang, Q.C.; Wang, G.; Wang, B.; et al. Identification of mecciRNAs and their roles in the mitochondrial entry of proteins. Sci. China Life Sci. 2020, 63, 1429–1449. [Google Scholar] [CrossRef]
  91. Qin, T.; Li, J.; Zhang, K.Q. Structure, regulation and function of linear and circular long non-coding RNAs. Front. Genet. 2020, 11, 150. [Google Scholar] [CrossRef] [Green Version]
  92. Vicens, Q.; Westhof, E. Biogenesis of circular RNAs. Cell 2014, 159, 13–14. [Google Scholar] [CrossRef] [Green Version]
  93. Meng, S.; Zhou, H.; Feng, Z.; Xu, Z.; Tang, Y.; Li, P.; Wu, M. CircRNA: Functions and properties of a novel potential biomarker for cancer. Mol. Cancer 2017, 16, 1. [Google Scholar] [CrossRef]
  94. Di, C.; Yuan, J.; Wu, Y.; Li, J.; Lin, H.; Hu, L.; Zhang, T.; Qi, Y.; Gerstein, M.B.; Guo, Y.; et al. Characterization of stress-responsive lncRNAs in Arabidopsis thaliana by integrating expression, epigenetic and structural features. Plant J. 2014, 80, 848–861. [Google Scholar] [CrossRef]
  95. Yuan, J.; Zhang, Y.; Dong, J.; Sun, Y.; Lim, B.L.; Liu, D.; Lu, Z.J. Systematic characterization of novel lncRNAs responding to phosphate starvation in Arabidopsis thaliana. BMC Genom. 2016, 17, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Weidong, Q.; Hongping, C.; Zuozhen, Y.; Biaolin, H.; Xiangdong, L.; Bing, A.; Yuan, L.; Yu, H.; Jiankun, X.; Fantao, Z. Systematic characterization of long non-coding RNAs and their responses to drought stress in Dongxiang wild rice. Rice Sci. 2020, 27, 21–31. [Google Scholar] [CrossRef]
  97. Yan, Q.; Wu, F.; Yan, Z.; Li, J.; Ma, T.; Zhang, Y.; Zhao, Y.; Wang, Y.; Zhang, J. Differential co-expression networks of long non-coding RNAs and mRNAs in Cleistogenes songorica under water stress and during recovery. BMC Plant Biol. 2019, 19, 1. [Google Scholar] [CrossRef] [Green Version]
  98. Liu, J.; Jung, C.; Xu, J.; Wang, H.; Deng, S.; Bernad, L.; Arenas-Huertero, C.; Chua, N.H. Genome-wide analysis uncovers regulation of long intergenic noncoding RNAs in Arabidopsis CW. Plant Cell 2012, 24, 4333–4345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Li, S.; Yamada, M.; Han, X.; Ohler, U.; Benfey, P.N. High-resolution expression map of the Arabidopsis root reveals alternative splicing and lincRNA regulation. Dev. Cell 2016, 9, 508–522. [Google Scholar] [CrossRef] [Green Version]
  100. Wang, Y.; Fan, X.; Lin, F.; He, G.; Terzaghi, W.; Zhu, D.; Deng, X.W. Arabidopsis noncoding RNA mediates control of photomorphogenesis by red light. Proc. Natl. Acad. Sci. USA 2014, 111, 10359–10364. [Google Scholar] [CrossRef] [Green Version]
  101. Seo, J.S.; Sun, H.X.; Park, B.S.; Huang, C.H.; Yeh, S.D.; Jung, C.; Chua, N.H. ELF18-induced long-noncoding RNA associates with mediator to enhance expression of innate immune response genes in Arabidopsis. Plant Cell 2017, 29, 1024–01038. [Google Scholar] [CrossRef] [Green Version]
  102. Zhang, W.; Han, Z.; Guo, Q.; Liu, Y.; Zheng, Y.; Fangli, W.; Jin, W. Identification of maize long non-coding RNAs responsive to drought stress. PLoS ONE 2014, 9, e98958. [Google Scholar] [CrossRef] [Green Version]
  103. Chen, R.; Li, M.; Zhang, H.; Duan, L.; Sun, X.; Jiang, Q.; Zhang, H.; Hu, Z. Continuous salt stress-induced long non-coding RNAs and DNA methylation patterns in soybean roots. BMC Genom. 2019, 20, 1–12. [Google Scholar] [CrossRef] [Green Version]
  104. Unver, T.; Tombuloglu, H. Barley long non-coding RNAs (lncRNA) responsive to excess boron. Genomics 2020, 112, 1947–1955. [Google Scholar] [CrossRef]
  105. de Urquiaga, M.C.O.; Thiebaut, F.; Hemerly, A.S.; Ferreira, P.C.G. From trash to luxury: The potential role of plant lncRNA in DNA methylation during abiotic stress. Front. Plant Sci. 2021, 11, 1–10. [Google Scholar] [CrossRef] [PubMed]
  106. Zhao, M.; Wang, T.; Sun, T.; Yu, X.; Tian, R.; Zhang, W.H. Identification of tissue-specific and cold-responsive lncRNAs in Medicago truncatula by high-throughput RNA sequencing. BMC Plant Biol. 2020, 1, 116. [Google Scholar] [CrossRef] [Green Version]
  107. Ariel, F.; Romero-Barrios, N.; Jégu, T.; Benhamed, M.; Crespi, M. Battles and hijacks: Noncoding transcription in plants. Trends Plant Sci. 2015, 20, 362–371. [Google Scholar] [CrossRef]
  108. Wang, H.; Chung, P.J.; Liu, J.; Jang, I.C.; Kean, M.J.; Xu, J.; Chua, N.H. Genome-wide identification of long noncoding natural antisense transcripts and their responses to light in Arabidopsis. Genome Res. 2014, 24, 444–453. [Google Scholar] [CrossRef] [Green Version]
  109. Wang, T.; Zhao, M.; Zhang, X.; Liu, M.; Yang, C.; Chen, Y.; Chen, R.; Wen, J.; Mysore, K.S.; Zhang, W.H. Novel phosphate deficiency-responsive long non-coding RNAs in the legume model plant Medicago truncatula. J. Exp. Bot. 2017, 68, 5937–5948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Geisler, S.; Coller, J. RNA in unexpected places: Long non-coding RNA functions in diverse cellular contexts. Nat. Rev. Mol. Cell Biol. 2013, 14, 69–712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Bardou, F.; Ariel, F.; Simpson, C.G.; Romero-Barrios, N.; Laporte, P.; Balzergue, S.; Brown, J.W.S.; Crespi, M. Long noncoding RNA modulates alternative splicing regulators in Arabidopsis. Dev. Cell 2014, 30, 166–176. [Google Scholar] [CrossRef] [Green Version]
  112. Franco-Zorrilla, J.M.; Valli, A.; Todesco, M.; Mateos, I.; Puga, M.I.; Rubio-Somoza, I.; Leyva, A.; Weigel, D.; García, J.A.; Paz-Ares, J. Target mimicry provides a new mechanism for regulation of microRNA activity. Nat. Genet. 2007, 39, 1033–1037. [Google Scholar] [CrossRef] [PubMed]
  113. Qin, T.; Zhao, H.; Cui, P.; Albesher, N.; Xionga, L. A nucleus-localized long non-coding RNA enhances drought and salt stress tolerance. Plant Physiol. 2017, 175, 1321–1336. [Google Scholar] [CrossRef] [Green Version]
  114. Tan, X.; Li, S.; Hu, L.; Zhang, C. Genome-wide analysis of long non-coding RNAs (lncRNAs) in two contrasting rapeseed (Brassica napus L.) genotypes subjected to drought stress and re-watering. BMC Plant Biol. 2020, 20, 1–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Li, S.; Yu, X.; Lei, N.; Cheng, Z.; Zhao, P.; He, Y.; Wang, W.; Peng, M. Genome-wide identification and functional prediction of cold and/or drought-responsive lncRNAs in cassava. Sci. Rep. 2016, 7, 1–17. [Google Scholar] [CrossRef]
  116. Ding, Z.; Wu, C.; Tie, W.; Yan, Y.; He, G.; Hu, W. Strand-specific RNA-seq based identification and functional prediction of lncRNAs in response to melatonin and simulated drought stresses in cassava. Plant Physiol. Biochem. 2019, 140, 96–104. [Google Scholar] [CrossRef]
  117. Zhang, C.; Tang, G.; Peng, X.; Sun, F.; Liu, S.; Xi, Y. Long non-coding RNAs of switchgrass (Panicum virgatum L.) in multiple dehydration stresses. BMC Plant Biol. 2018, 18, 1–15. [Google Scholar] [CrossRef] [Green Version]
  118. Shuai, P.; Liang, D.; Tang, S.; Zhang, Z.; Ye, C.Y.; Su, Y.; Xia, X.; Yin, W. Genome-wide identification and functional prediction of novel and drought-responsive lincRNAs in Populus trichocarpa. J. Exp. Bot. 2014, 65, 4975–4983. [Google Scholar] [CrossRef]
  119. Chung, P.J.; Jung, H.; Jeong, D.H.; Ha, S.H.; Choi, Y.D.; Kim, J.K. Transcriptome profiling of drought responsive noncoding RNAs and their target genes in rice. BMC Genom. 2016, 17, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Cagirici, H.B.; Alptekin, B.; Budak, H. RNA sequencing and co-expressed long non-coding RNA in modern and wild wheats. Sci. Rep. 2017, 7, 1–16. [Google Scholar] [CrossRef] [Green Version]
  121. Mao, H.; Wang, H.; Liu, S.; Li, Z.; Yang, X.; Yan, J.; Li, J.; Tran, L.S.P.; Qin, F. A transposable element in a NAC gene is associated with drought tolerance in maize seedlings. Nat. Commun. 2015, 6, 1–13. [Google Scholar] [CrossRef] [Green Version]
  122. Pang, J.; Zhang, X.; Ma, X.; Zhao, J. Spatio-temporal transcriptional dynamics of maize long non-coding RNAs responsive to drought stress. Genes 2019, 10, 138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Du, Q.; Wang, K.; Zou, C.; Xu, C.; Li, W.X. The PILNCR1-miR399 regulatory module is important for low phosphate tolerance in maize. Plant Physiol. 2018, 177, 1743–1753. [Google Scholar] [CrossRef] [Green Version]
  124. Waseem, M.; Liu, Y.; Xia, R. Long non-coding RNAs, the dark matter: An emerging regulatory component in plants. Int. J. Mol. Sci. 2021, 22, 86. [Google Scholar] [CrossRef] [PubMed]
  125. Wu, H.J.; Wang, Z.M.; Wang, M.; Wang, X.J. Widespread long noncoding RNAs as endogenous target mimics for microRNAs in plants. Plant Physiol. 2013, 161, 1875–1884. [Google Scholar] [CrossRef] [Green Version]
  126. Wang, J.; Yu, W.; Yang, Y.; Li, X.; Chen, T.; Liu, T.; Ma, N.; Yang, X.; Liu, R.; Zhang, B. Genome-wide analysis of tomato long non-coding RNAs and identification as endogenous target mimic for microRNA in response to TYLCV infection. Sci. Rep. 2015, 5, 1–16. [Google Scholar] [CrossRef] [Green Version]
  127. Todesco, M.; Rubio-Somoza, I.; Paz-Ares, J.; Weigel, D. A collection of target mimics for comprehensive analysis of MicroRNA function in Arabidopsis thaliana. PLoS Genet. 2010, 6, e1001031. [Google Scholar] [CrossRef] [Green Version]
  128. Borah, P.; Das, A.; Milner, M.J.; Ali, A.; Bentley, A.R.; Pandey, R. Long non-coding RNAs as endogenous target mimics and exploration of their role in low nutrient stress tolerance in plants. Genes 2018, 9, 459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Jiang, N.; Cui, J.; Shi, Y.; Yang, G.; Zhou, X.; Hou, X.; Meng, J.; Luan, Y. Tomato lncRNA23468 functions as a competing endogenous RNA to modulate NBS-LRR genes by decoying miR482b in the tomato-Phytophthora infestans interaction. Hortic. Res. 2019, 6, 1. [Google Scholar] [CrossRef] [Green Version]
  130. Cui, J.; Luan, Y.; Jiang, N.; Bao, H.; Meng, J. Comparative transcriptome analysis between resistant and susceptible tomato allows the identification of lncRNA16397 conferring resistance to Phytophthora infestans by co-expressing glutaredoxin. Plant J. 2017, 89, 577–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Jalali, S.; Bhartiya, D.; Lalwani, M.K.; Sivasubbu, S.; Scaria, V. Systematic transcriptome wide analysis of lncRNA-miRNA interactions. PLoS ONE 2013, 8, e53823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Juan, L.; Wang, G.; Radovich, M.; Schneider, B.P.; Clare, S.E.; Wang, Y.; Liu, Y. Potential roles of microRNAs in regulating long intergenic noncoding RNAs. BMC Med. Genom. 2013, 6, 1–11. [Google Scholar] [CrossRef] [Green Version]
  133. Ding, J.; Shen, J.; Mao, H.; Xie, W.; Li, X.; Zhang, Q. RNA-directed DNA methylation is involved in regulating photoperiod- sensitive male sterility in rice. Mol. Plant. 2012, 5, 1210–1216. [Google Scholar] [CrossRef] [Green Version]
  134. Chen, Y.; Li, X.; Su, L.; Chen, X.; Zhang, S.; Xu, X.; Zhang, Z.; Chen, Y.; XuHan, X.; Lin, Y.; et al. Genome-wide identification and characterization of long non-coding RNAs involved in the early somatic embryogenesis in Dimocarpus longan Lour. BMC Genom. 2018, 19, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Csorba, T.; Questa, J.I.; Sun, Q.; Dean, C. Antisense COOLAIR mediates the coordinated switching of chromatin states at FLC during vernalization. Proc. Natl. Acad. Sci. USA 2014, 111, 16160–16165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Heo, J.B.; Sung, S. Vernalization-mediated epigenetic silencing by a long intronic noncoding RNA. Science 2011, 80, 76–79. [Google Scholar] [CrossRef] [Green Version]
  137. Zhou, M.; Palanca, A.M.S.; Law, J.A. Locus-specific control of the de novo DNA methylation pathway in Arabidopsis by the CLASSY family. Nat. Genet. 2018, 50, 865–873. [Google Scholar] [CrossRef]
  138. Huang, C.F.; Zhu, J.K. RNA splicing factors and RNA-directed DNA methylation. Biology 2014, 3, 243–254. [Google Scholar] [CrossRef] [PubMed]
  139. Sanan-Mishra, N.; Abdul Kader Jailani, A.; Mandal, B.; Mukherjee, S.K. Secondary siRNAs in plants: Biosynthesis, various functions and applications in virology. Front. Plant Sci. 2021, 12, 1–32. [Google Scholar] [CrossRef]
  140. Matzke, M.; Kanno, T.; Daxinger, L.; Huettel, B.; Matzke, A.J. RNA-mediated chromatin-based silencing in plants. Curr. Opin. Cell Biol. 2009, 21, 367–376. [Google Scholar] [CrossRef]
  141. Gao, Z.; Liu, H.L.; Daxinger, L.; Pontes, O.; He, X.; Qian, W.; Lin, H.; Xie, M.; Lorkovic, Z.J.; Zhang, S.; et al. An RNA polymerase II-and AGO4-associated protein acts in RNA-directed DNA methylation. Nature 2010, 465, 106–109. [Google Scholar] [CrossRef] [Green Version]
  142. Zheng, B.; Wang, Z.; Li, S.; Yu, B.; Liu, J.Y.; Chen, X. Intergenic transcription by RNA polymerase II coordinates Pol IV and Pol V in siRNA-directed transcriptional gene silencing in Arabidopsis. Genes Dev. 2009, 23, 2850–2860. [Google Scholar] [CrossRef] [Green Version]
  143. Law, J.A.; Vashisht, A.A.; Wohlschlegel, J.A.; Jacobsen, S.E. SHH1, a homeodomain protein required for DNA methylation, as well as RDR2, RDM4 and chromatin remodeling factors, associate with RNA Polymerase IV. PLoS Genet. 2011, 7, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Haag, J.R.; Pikaard, C.S. Multi-subunit RNA polymerases IV and V: Purveyors of non-coding RNA for plant gene silencing. Nat. Rev. Mol. Cell Biol. 2011, 12, 483–492. [Google Scholar] [CrossRef]
  145. Wierzbicki, A.T.; Ream, T.S.; Haag, J.R.; Pikaard, C.S. RNA polymerase v transcription guides ARGONAUTE4 to chromatin. Nat. Genet. 2009, 41, 630–634. [Google Scholar] [CrossRef] [PubMed]
  146. Zhong, X.; Hale, C.J.; Law, J.A.; Johnson, L.M.; Feng, S.; Tu, A.; Jacobsen, S.E. DDR complex facilitates global association of RNA polymerase v to promoters and evolutionarily young transposons. Nat. Struct. Mol. Biol. 2012, 9, 870–875. [Google Scholar] [CrossRef] [Green Version]
  147. Ramirez-Prado, J.S.; Piquerez, S.J.M.; Bendahmane, A.; Hirt, H.; Raynaud, C.; Benhamed, M. Modify the histone to win the battle: Chromatin dynamics in plant-pathogen interactions. Front. Plant Sci. 2018, 9, 1–18. [Google Scholar] [CrossRef] [PubMed]
  148. Nobuta, K.; Lu, C.; Shrivastava, R.; Pillay, M.; De Paoli, E.; Accerbi, M.; Arteaga-Vazquez, M.; Sidorenko, L.; Jeong, D.H.; Yen, Y.; et al. Distinct size distribution of endogenous siRNAs in maize: Evidence from deep sequencing in the mop1-1 mutant. Proc. Natl. Acad. Sci. USA 2008, 105, 14958–14963. [Google Scholar] [CrossRef] [Green Version]
  149. Boerner, S.; McGinnis, K.M. Computational identification and functional predictions of long noncoding RNA in Zea mays. PLoS ONE 2012, 7, e43047. [Google Scholar] [CrossRef] [PubMed]
  150. Li, S.; Castillo-González, C.; Yu, B.; Zhang, X. The functions of plant small RNAs in development and in stress responses. Plant J. 2017, 90, 654–670. [Google Scholar] [CrossRef] [Green Version]
  151. Zhao, C.; Sun, X.; Li, L. Biogenesis and function of extracellular miRNAs. ExRNA 2019, 1, 1–9. [Google Scholar] [CrossRef] [Green Version]
  152. Nadarajah, K.; Kumar, I.S. Drought response in rice: The miRNA story. Int. J. Mol. Sci. 2019, 20, 3766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Betti, F.; Ladera-Carmona, M.J.; Perata, P.; Loreti, E. RNAi mediated hypoxia stress tolerance in plants. Int. J. Mol. Sci. 2020, 21, 9394. [Google Scholar] [CrossRef]
  154. Brant, E.J.; Budak, H. Plant small non-coding RNAs and their roles in biotic stresses. Front. Plant Sci. 2018, 9, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. César de Lima, J.; Loss-Morais, G.; Margis, R. MicroRNAs play critical roles during plant development and in response to abiotic stresses. Genet. Mol. Biol. 2012, 35, 1069–1077. [Google Scholar] [CrossRef] [Green Version]
  156. Djami-Tchatchou, A.T.; Sanan-Mishra, N.; Ntushelo, K.; Dubery, I.A. Functional roles of microRNAs in agronomically important plants-potential as targets for crop improvement and protection. Front. Plant Sci. 2017, 8, 378. [Google Scholar] [CrossRef] [Green Version]
  157. Celton, J.M.; Gaillard, S.; Bruneau, M.; Pelletier, S.; Aubourg, S.; Martin-Magniette, M.L.; Navarro, L.; Laurens, F.; Renou, J.P. Widespread anti-sense transcription in apple is correlated with siRNA production and indicates a large potential for transcriptional and/or post-transcriptional control. New Phytol. 2014, 203, 287–299. [Google Scholar] [CrossRef]
  158. Zhang, N.; Yang, J.; Wang, Z.; Wen, Y.; Wang, J.; He, W.; Liu, B.; Si, H.; Wang, D. Identification of novel and conserved microRNAs related to drought stress in potato by deep sequencing. PLoS ONE 2014, 9, e95489. [Google Scholar] [CrossRef]
  159. Zhang, L.; Chia, J.-M.; Kumari, S.; Stein, J.C.; Liu, Z.; Narechania, A.; Maher, C.A.; Guill, K.; McMullen, M.D.; Ware, D. A genome-wide characterization of microRNA genes in maize. PLoS Genet. 2009, 5, e1000716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Lippman, Z.; Martienssen, R. The role of RNA interference in heterochromatic silencing. Nature 2004, 431, 364–370. [Google Scholar] [CrossRef]
  161. Agrawal, N.; Dasaradhi, P.V.; Mohmmed, A.; Malhotra, P.; Bhatnagar, R.K.; Mukherjee, S.K. RNA interference: Biology, mechanism, and applications. Microbiol. Mol. Biol. Rev. 2003, 67, 657–685. [Google Scholar] [CrossRef] [Green Version]
  162. MacRae, I.J.; Li, F.; Zhou, K.; Cande, W.Z.; Doudna, J.A. Structure of dicer and mechanistic implications for RNAi. Cold Spring Harb. Symp. Quant. Biol. 2006, 71, 73–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Elbashir, S.M.; Martinez, J.; Patkaniowska, A.; Lendeckel, W.; Tuschl, T. Functional anatomy of siRNAs for mediating efficient RNAi in Drosophila melanogaster embryo lysate. EMBO J. 2001, 20, 6877–6888. [Google Scholar] [CrossRef] [Green Version]
  164. Schwarz, D.S.; Hutvágner, G.; Du, T.; Xu, Z.; Aronin, N.; Zamore, P.D. Asymmetry in the assembly of the RNAi enzyme complex. Cell 2003, 115, 199–208. [Google Scholar] [CrossRef] [Green Version]
  165. Guo, Q.; Liu, Q.; Smith, N.A.; Liang, G.; Wang, M.B. RNA silencing in Plants: Mechanisms, technologies and applications in horticultural crops. Curr. Genom. 2016, 17, 476–489. [Google Scholar] [CrossRef] [PubMed]
  166. Barber, W.T.; Zhang, W.; Win, H.; Varala, K.K.; Dorweiler, J.E.; Hudson, M.E.; Moose, S.P. Repeat associated small RNAs vary among parents and following hybridization in maize. Proc. Natl. Acad. Sci. USA 2012, 109, 10444–10449. [Google Scholar] [CrossRef] [Green Version]
  167. Mallory, A.C.; Vaucheret, H. ARGONAUTE 1 homeostasis invokes the coordinate action of the microRNA and siRNA pathways. EMBO Rep. 2009, 10, 521–526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Rajagopalan, R.; Vaucheret, H.; Trejo, J.; Bartel, D.P. A diverse and evolutionarily fluid set of microRNAs in Arabidopsis thaliana. Genes Dev. 2006, 20, 3407–3425. [Google Scholar] [CrossRef] [Green Version]
  169. Vazquez, F.; Vaucheret, H.; Rajagopalan, R.; Lepers, C.; Gasciolli, V.; Mallory, A.C.; Hilbert, J.L.; Bartel, D.P.; Crete, P. Endogenous trans-acting siRNAs regulate the accumulation of Arabidopsis mRNAs. Mol. Cell 2004, 16, 69–79. [Google Scholar] [CrossRef] [PubMed]
  170. Katiyar, A.; Smita, S.; Muthusamy, S.K.; Chinnusamy, V.; Pandey, D.M.; Bansal, K.C. Identification of novel drought-responsive microRNAs and trans-acting siRNAs from Sorghum bicolor (L.) Moench by high-throughput sequencing analysis. Front. Plant Sci. 2015, 6, 506. [Google Scholar] [CrossRef] [Green Version]
  171. Matsui, A.; Mizunashi, K.; Tanaka, M.; Kaminuma, E.; Nguyen, A.H.; Nakajima, M.; Kim, J.M.; Nguyen, D.; Van Toyoda, T.; Seki, M. TasiRNA-ARF pathway moderates floral architecture in Arabidopsis plants subjected to drought stress. BioMed Res. Int. 2014, 2014, 303451. [Google Scholar] [CrossRef] [Green Version]
  172. Hunter, C.; Willmann, M.R.; Wu, G.; Yoshikawa, M.; de Gutierrez-Nava, M.L.L.; Poethig, R.S. Trans-acting siRNA-mediated repression of ETTIN and ARF4 regulates heteroblasty in Arabidopsis. Development 2006, 133, 2973–2981. [Google Scholar] [CrossRef] [Green Version]
  173. Yu, D.; Meng, Y.; Zuo, Z.; Xue, J.; Wang, H. NATpipe: An integrative pipeline for systematical discovery of natural antisense transcripts (NATs) and phase-distributed nat-siRNAs from de novo assembled transcriptomes. Sci. Rep. 2016, 6, 8–13. [Google Scholar] [CrossRef] [Green Version]
  174. Jabnoune, M.; Secco, D.; Lecampion, C.; Robaglia, C.; Shu, Q.; Poirier, Y. A rice cis-natural antisense RNA acts as a translational enhancer for its cognate mRNA and contributes to phosphate homeostasis and plant fitness. Plant Cell 2013, 25, 4166–4182. [Google Scholar] [CrossRef] [Green Version]
  175. Wang, X.J.; Gaasterland, T.; Chua, N.H. Genome-wide prediction and identification of cis-natural antisense transcripts in Arabidopsis thaliana. Genome Biol. 2005, 6, 4. [Google Scholar] [CrossRef] [Green Version]
  176. Lapidot, M.; Pilpel, Y. Genome-wide natural antisense transcription: Coupling its regulation to its different regulatory mechanisms. EMBO Rep. 2006, 7, 1216–1222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Faghihi, M.A.; Wahlestedt, C. Regulatory roles of natural antisense transcripts. Nat. Rev. Mol. Cell Biol. 2009, 10, 637–643. [Google Scholar] [CrossRef]
  178. Faghihi, M.A.; Zhang, M.; Huang, J.; Modarresi, F.; Van der Brug, M.P.; Nalls, M.A.; Cookson, M.R.; St-Laurent, G.; Wahlestedt, C. Evidence for natural antisense transcript-mediated inhibition of microRNA function. Genome Biol. 2010, 11, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Lu, T.; Zhu, C.; Lu, G.; Guo, Y.; Zhou, Y.; Zhang, Z.; Zhao, Y.; Li, W.; Lu, Y.; Tang, W.; et al. Strand-specific RNA-seq reveals widespread occurrence of novel cis-natural antisense transcripts in rice. BMC Genom. 2012, 13, 1. [Google Scholar] [CrossRef] [Green Version]
  180. Castel, S.E.; Martienssen, R.A. RNA interference in the nucleus: Roles for small RNAs in transcription, epigenetics and beyond. Nat. Rev. Genet. 2013, 14, 100–112. [Google Scholar] [CrossRef] [PubMed]
  181. Law, J.A.; Jacobsen, S.E. Establishing, maintaining and modifying DNA methylation patterns in plants and animals. Nat. Rev. Genet. 2010, 11, 204–220. [Google Scholar] [CrossRef] [PubMed]
  182. Azevedo, J.; Cooke, R.; Lagrange, T. Taking RISCs with AGO hookers. Curr. Opin. Plant Biol. 2011, 14, 594–600. [Google Scholar] [CrossRef] [PubMed]
  183. Lopez, A.; Ramírez, V.; García-Andrade, J.; Flors, V.; Vera, P. The RNA silencing enzyme RNA polymerase V is required for plant immunity. PLoS Genet. 2011, 7, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Dowen, R.H.; Pelizzola, M.; Schmitz, R.J.; Lister, R.; Dowen, J.M.; Nery, J.R.; Dixon, J.E.; Ecker, J.R. Widespread dynamic DNA methylation in response to biotic stress. Proc. Natl. Acad. Sci. USA 2012, 109, 32. [Google Scholar] [CrossRef] [Green Version]
  185. Yu, A.; Lepère, G.; Jay, F.; Wang, J.; Bapaume, L.; Wang, Y.; Abraham, A.L.; Penterman, J.; Fischer, R.L.; Voinnet, O.; et al. Dynamics and biological relevance of DNA demethylation in Arabidopsis antibacterial defense. Proc. Natl. Acad. Sci. USA 2013, 110, 2389–2394. [Google Scholar] [CrossRef] [Green Version]
  186. Medina, C.; Da Rocha, M.; Magliano, M.; Raptopoulo, A.; Marteu, N.; Lebrigand, K.; Abad, P.; Favery, B.; Jaubert-Possamai, S. Characterization of siRNAs clusters in Arabidopsis thaliana galls induced by the root-knot nematode Meloidogyne incognita. BMC Genom. 2018, 19, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Chinnusamy, V.; Zhu, J.K. RNA-directed DNA methylation and demethylation in plants. Sci. China C Life Sci. 2009, 52, 331–343. [Google Scholar] [CrossRef] [Green Version]
  188. Ku, Y.S.; Wong, J.W.H.; Mui, Z.; Liu, X.; Hui, J.H.L.; Chan, T.F.; Lam, H.M. Small RNAs in plant responses to abiotic stresses: Regulatory roles and study methods. Int. J. Mol. Sci. 2015, 16, 24532–24554. [Google Scholar] [CrossRef]
  189. Ding, S.W.; Lu, R. Virus-derived siRNAs and piRNAs in immunity and pathogenesis. Curr. Opin. Virol. 2011, 1, 533–544. [Google Scholar] [CrossRef] [Green Version]
  190. Sharma, N.; Sahu, P.P.; Puranik, S.; Prasad, M. Recent advances in plant-virus interaction with emphasis on small interfering RNAs (siRNAs). Mol. Biotechnol. 2013, 55, 63–77. [Google Scholar] [CrossRef]
  191. Szittya, G.; Moxon, S.; Pantaleo, V.; Toth, G.; Pilcher, R.L.R.; Moulton, V.; Burgyan, J.; Dalmay, T. Structural and functional analysis of viral siRNAs. PLoS Pathog. 2010, 6, e1000838. [Google Scholar] [CrossRef]
  192. Zhu, B.; Gao, H.; Xu, G.; Wu, D.; Song, S.; Jiang, H.; Zhu, S.; Qi, T.; Xie, D. Arabidopsis ALA1 and ALA2 mediate RNAi-based antiviral immunity. Front. Plant Sci. 2017, 8, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Glick, E.; Zrachya, A.; Levy, Y.; Mett, A.; Gidoni, D.; Belausov, E.; Citovsky, V.; Gafni, Y. Interaction with host SGS3 is required for suppression of RNA silencing by tomato yellow leaf curl virus V2 protein. Proc. Natl. Acad. Sci. USA 2008, 105, 157–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Yadav, R.K.; Chattopadhyay, D. Enhanced viral intergenic region-specific short interfering RNA accumulation and DNA methylation correlates with resistance against a geminivirus. Mol. Plant Microbe Interact. 2011, 24, 1189–1197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Donaire, L.; Wang, Y.; Gonzalez-Ibeas, D.; Mayer, K.F.; Aranda, M.A.; Llave, C. Deep-sequencing of plant viral small RNAs reveals effective and widespread targeting of viral genomes. Virology 2009, 392, 203–214. [Google Scholar] [CrossRef] [Green Version]
  196. Wang, H.; Wang, H.; Duan, X.; Liu, C.; Li, Z. Digital quantitative analysis of microRNA in single cell based on ligation-depended polymerase colony (Polony). Biosens. Bioelectron. 2017, 95, 146–151. [Google Scholar] [CrossRef]
  197. Dong, Z.; Han, M.H.; Fedoroff, N. The RNA-binding proteins HYL1 and SE promote accurate in vitro processing of pri-miRNA by DCL1. Proc. Natl. Acad. Sci. USA 2008, 105, 9970–9975. [Google Scholar] [CrossRef] [Green Version]
  198. Catalanotto, C.; Cogoni, C.; Zardo, G. MicroRNA in control of gene expression: An overview of nuclear functions. Int. J. Mol. Sci. 2016, 17, 1712. [Google Scholar] [CrossRef] [Green Version]
  199. Liu, W.W.; Meng, J.; Cui, J.; Luan, Y.S. Characterization and function of microRNAs in plants. Front. Plant Sci. 2017, 8, 1–7. [Google Scholar] [CrossRef] [Green Version]
  200. Vaucheret, H.; Vazquez, F.; Crete, P.; Bartel, D.P. The action of ARGONAUTE1 in the miRNA pathway and its regulation by the miRNA pathway are crucial for plant development. Genes 2004, 18, 1187–1197. [Google Scholar] [CrossRef] [Green Version]
  201. Rubio-Somoza, I.; Weigel, D. MicroRNA networks and developmental plasticity in plants. Trends Plant Sci. 2011, 16, 258–264. [Google Scholar] [CrossRef]
  202. Ochando, I.; Jover-Gil, S.; Jose’ Ripoll, J.; Candela, H.; Vera, A.; Ponce, M.R.; Martınez-Laborda, A.; Micol, J.L. Mutations in the microRNA complementarity site of the INCURVATA4 gene perturb meristem function and adaxialize lateral organs in Arabidopsis. Plant Physiol. 2006, 141, 607–619. [Google Scholar] [CrossRef] [Green Version]
  203. Jin, D.; Wang, Y.; Zhao, Y.; Chen, M. MicroRNAs and their cross-talks in plant development. J. Genet. Genom. 2013, 40, 161–170. [Google Scholar] [CrossRef] [PubMed]
  204. Liu, H.; Yu, H.; Tang, G.; Huang, T. Small but powerful: Function of microRNAs in plant development. Plant Cell Rep. 2018, 37, 515–528. [Google Scholar] [CrossRef] [PubMed]
  205. Li, C.; Zhang, B. MicroRNAs in control of plant development. J. Cell. Physiol. 2016, 231, 303–313. [Google Scholar] [CrossRef]
  206. Chen, C.; Zeng, Z.; Liu, Z.; Xia, R. Small RNAs, emerging regulators critical for the development of horticultural traits. Hortic. Res. 2018, 5, 1–14. [Google Scholar] [CrossRef] [Green Version]
  207. 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]
  208. Crisp, P.A.; Ganguly, D.; Eichten, S.R.; Borevitz, J.O.; Pogson, B.J. Reconsidering plant memory: Intersections between stress recovery, RNA turnover and epigenetics. Sci. Adv. 2016, 2, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Shriram, V.; Kumar, V.; Devarumath, R.M.; Khare, T.S.; Wani, S.H. MicroRNAs as potential targets for abiotic stress tolerance in plants. Front. Plant Sci. 2016, 7, 1–18. [Google Scholar] [CrossRef]
  210. Jatan, R.; Lata, C. Role of microRNAs in abiotic and biotic stress resistance in plants. Proc. Indian Natl. Sci. Acad. 2019, 85, 553–567. [Google Scholar] [CrossRef]
  211. Zhu, J.; Geisler, M. Keeping it all together: Auxin-actin crosstalk in plant development. J. Exp. Bot. 2015, 66, 4983–4998. [Google Scholar] [CrossRef] [Green Version]
  212. Hernández, Y.; Sanan-Mishra, N. miRNA mediated regulation of NAC transcription factors in plant development and environment stress response. Plant Gene 2017, 11, 190–198. [Google Scholar] [CrossRef]
  213. Wang, J.W.; Wang, L.J.; Mao, Y.B.; Cai, W.J.; Xue, H.W.; Chen, X.Y. Control of root cap formation by MicroRNA-targeted auxin response factors in Arabidopsis. Plant Cell 2005, 17, 2204–2216. [Google Scholar] [CrossRef] [Green Version]
  214. Wu, M.F.; Tian, Q.; Reed, J.W. Arabidopsis microRNA 167 controls patterns of ARF6 and ARF8 expression and regulates both female and male reproduction. Development 2006, 133, 4211–4218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Lu, C.; Fedoroff, N. A mutation in the Arabidopsis HYL1 gene encoding a dsRNA binding protein affects responses to abscisic acid, auxin and cytokinin. Plant Cell 2000, 12, 2351–2365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Tsuji, H.; Aya, K.; Ueguchi-Tanaka, M.; Shimada, Y.; Nakazono, M.; Watanabe, R.; Nishizawa, N.K.; Gomi, K.; Shimada, A.; Kitano, H.; et al. GAMYB controls different sets of genes and is differentially regulated by microRNA in aleurone cells and anthers. Plant J. 2006, 47, 427–444. [Google Scholar] [CrossRef]
  217. Achard, P.; Herr, A.; Baulcombe, D.C.; Harberd, N.P. Modulation of floral development by a gibberellin-regulated microRNA. Development 2004, 131, 3357–3365. [Google Scholar] [CrossRef] [Green Version]
  218. Guo, H.S.; Xie, Q.; Fei, J.F.; Chua, N.H. MicroRNA directs mRNA cleavage of the transcription factor NAC1 to down regulate auxin signals for Arabidopsis lateral root development. Plant Cell 2005, 17, 1376–1386. [Google Scholar] [CrossRef] [Green Version]
  219. Iglesias, M.J.; Terrile, M.C.; Windels, D.; Lombardo, M.C.; Bartoli, C.G.; Vazquez, F.; Estelle, M.; Casalongué, C.A. MiR393 regulation of auxin signaling and redox-related components during acclimation to salinity in Arabidopsis. PLoS ONE 2014, 9, e107678. [Google Scholar] [CrossRef] [Green Version]
  220. Chen, Z.; Hu, L.; Han, N.; Hu, J.; Yang, Y.; Xiang, T.; Zhang, X.; Wang, L. Overexpression of a miR393-resistant form of transport inhibitor response protein 1 (mTIR1) enhances salt tolerance by increased osmoregulation and Na+ exclusion in Arabidopsis thaliana. Plant Cell Physiol. 2015, 56, 73–83. [Google Scholar] [CrossRef] [Green Version]
  221. Guo, F.; Han, N.; Xie, Y.; Fang, K.; Yang, Y.; Zhu, M.; Wang, J.; Bian, H. The miR393a/target module regulates seed germination and seedling establishment under submergence in rice (Oryza sativa L.). Plant Cell Environ. 2016, 39, 2288–2302. [Google Scholar] [CrossRef]
  222. Jodder, J. miRNA-mediated regulation of auxin signaling pathway during plant development and stress responses. J. Biosci. 2020, 45, 110. [Google Scholar] [CrossRef]
  223. Sanan-Mishra, N.; Varanasi, S.P.R.M.; Mukherjee, S.K. Micro-regulators of auxin action. Plant Cell Rep. 2013, 32, 733–740. [Google Scholar] [CrossRef] [PubMed]
  224. 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] [PubMed] [Green Version]
  225. Deng, M.; Cao, Y.; Zhao, Z.; Yang, L.; Zhang, Y.; Dong, Y.; Fan, G. Discovery of MicroRNAs and their target genes related to drought in Paulownia ‘Yuza 1’ by high-throughput sequencing. Int. J. Genom. 2017, 2017, 3674682. [Google Scholar] [CrossRef] [Green Version]
  226. Yang, C.; Li, D.; Mao, D.; Liu, X.; JI, C.; Li, X.; Zhao, X.; Cheng, Z.; Chen, C.; Zhu, L. Overexpression of microRNA319 impacts leaf morphogenesis and leads to enhanced cold tolerance in rice (Oryza sativa L.). Plant Cell Environ. 2013, 36, 2207–2218. [Google Scholar] [CrossRef]
  227. Branscheid, A.; Sieh, D.; Pant, B.D.; May, P.; Devers, E.A.; Elkrog, A.; Schauser, L.; Scheible, W.-R.; Krajinski, F. Expression pattern suggests a role of miR399 in the regulation of the cellular response to local Pi increase during arbuscular mycorrhizal symbiosis. Mol. Plant Microbe Interact. 2010, 23, 915–926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Fang, X.; Zhao, Y.; Ma, Q.; Huang, Y.; Wang, P.; Zhang, J.; Nian, H.; Yang, C. Identification and comparative analysis of cadmium tolerance-associated miRNAs and their targets in two soybean genotypes. PLoS ONE 2013, 8, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Huang, J.; Yang, M.; Zhang, X. The function of small RNAs in plant biotic stress response. J. Integr. Plant Biol. 2016, 58, 312–327. [Google Scholar] [CrossRef] [Green Version]
  230. Xia, K.; Wang, R.; Ou, X.; Fang, Z.; Tian, C.; Duan, J.; Wang, Y.; Zhang, M. OsTIR1 and OsAFB2 down regulation 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]
  231. Gu, M.; Xu, K.; Chen, A.; Zhu, Y.; Tang, G.; Xu, G. Expression analysis suggests potential roles of microRNAs for phosphate and arbuscular mycorrhizal signaling in Solanum lycopersicum. Physiol. Plant. 2010, 138, 226–237. [Google Scholar] [CrossRef]
  232. Xie, Z.; Allen, E.; Fahlgren, N.; Calamar, A.; Givan, S.A.; Carrington, J.C. Expression of Arabidopsis miRNA genes. Plant Physiol. 2005, 138, 2145–2154. [Google Scholar] [CrossRef] [Green Version]
  233. Kawashima, C.G.; Yoshimoto, N.; Maruyama-Nakashita, A.; Tsuchiya, Y.N.; Saito, K.; Takahashi, H.; Dalmay, T. Sulphur starvation induces the expression of microRNA-395 and one of its target genes but in different cell types. Plant J. 2009, 57, 313–321. [Google Scholar] [CrossRef] [PubMed]
  234. Megraw, M.; Baev, V.; Rusinov, V.; Jensen, S.T.; Kalantidis, K.; Hatzigeorgiou, A.G. MicroRNA promoter element discovery in Arabidopsis. RNA 2006, 12, 1612–1619. [Google Scholar] [CrossRef] [Green Version]
  235. Parizotto, E.A.; Dunoyer, P.; Rahm, N.; Himber, C.; Voinnet, O. In vivo investigation of the transcription, processing, endonucleolytic activity and functional relevance of the spatial distribution of a plant miRNA. Genes Dev. 2004, 1818, 2237–2242. [Google Scholar] [CrossRef] [Green Version]
  236. Xie, Z.; Kasschau, K.D.; Carrington, J.C. Negative feedback regulation of Dicer-Like1 in Arabidopsis by microRNA-guided mRNA degradation. Curr. Biol. 2003, 13, 784–789. [Google Scholar] [CrossRef] [Green Version]
  237. Pouch-Pelissier, M.N.; Pelissier, T.; Elmayan, T.; Vaucheret, H.; Boko, D.; Jantsch, M.F.; Deragon, J.M. SINE RNA induces severe developmental defects in Arabidopsis thaliana and interacts with HYL1 (DRB1), a key member of the DCL1 complex. PLoS Genet. 2008, 4, e1000096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Montgomery, T.A.; Howell, M.D.; Cuperus, J.T.; Li, D.; Hansen, J.E.; Alexander, A.L.; Chapman, E.J.; Fahlgren, N.; Allen, E.; Carrington, J.C. Specificity of ARGONAUTE7-miR390 interaction and dual functionality in TAS3 trans-acting siRNA formation. Cell 2008, 133, 128–141. [Google Scholar] [CrossRef] [Green Version]
  239. Takeda, A.; Iwasaki, S.; Watanabe, T.; Utsumi, M.; Watanabe, Y. The mechanism selecting the guide strand from small RNA duplexes is different among Argonaute proteins. Plant Cell Physiol. 2008, 49, 493–500. [Google Scholar] [CrossRef] [Green Version]
  240. Vazquez, F.; Blevins, T.; Ailhas, J.; Boller, T.; Meins, F., Jr. Evolution of Arabidopsis miR genes generates novel microRNA classes. Nucleic Acids Res. 2008, 36, 6429–6438. [Google Scholar] [CrossRef] [Green Version]
  241. Ru, P.; Xu, L.; Ma, H.; Huang, H. Plant fertility defects induced by the enhanced expression of microRNA167. Cell Res. 2006, 16, 457–465. [Google Scholar] [CrossRef] [Green Version]
  242. Cartolano, M.; Castillo, R.; Efremova, N.; Kuckenberg, M.; Zethof, J.; Gerats, T.; Schwarz-Sommer, Z.; Vandenbussche, M. A conserved microRNA module exerts homeotic control over Petunia hybrida and Antirrhinum majus floral organ identity. Nat. Genet. 2007, 39, 901–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Mlotshwa, S.; Yang, Z.; Kim, Y.J.; Chen, X. Floral patterning defects induced by Arabidopsis APETALA2 and microRNA172 expression in Nicotiana benthamiana. Plant Mol. Biol. 2006, 61, 781–793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Song, J.B.; Gao, S.; Sun, D.; Li, H.; Shu, X.X.; Yang, Z.M. miR394 and LCR are involved in Arabidopsis salt and drought stress responses in an abscisic acid-dependent manner. BMC Plant Biol. 2013, 13, 210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Rodriguez, R.E.; Mecchia, M.A.; Debernardi, J.M.; Schommer, C.; Weigel, D.; Palatnik, J.F. Control of cell proliferation in Arabidopsis thaliana by microRNA miR396. Development 2010, 137, 103–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. .Song, Z.; Zhang, L.; Wang, Y.; Li, H.; Li, S.; Zhao, H.; Zhang, H. Constitutive expression of miR408 improves biomass and seed yield in Arabidopsis. Front. Plant Sci. 2018, 8, 1–14. [Google Scholar] [CrossRef]
  247. Gupta, S.K.; Rai, A.K.; Kanwar, S.S.; Chand, D.; Singh, N.K.; Sharma, T.R. The single functional blast resistance gene Pi54 activates a complex defense mechanism in rice. J. Exp. Bot. 2012, 63, 757–772. [Google Scholar] [CrossRef] [Green Version]
  248. Kutter, C.; Schob, H.; Stadler, M.; Meins, F.; Si-Ammour, A. MicroRNA-mediated regulation of stomatal development in Arabidopsis. Plant Cell 2007, 19, 2417–2429. [Google Scholar] [CrossRef] [Green Version]
  249. Rybel, B.D.; Vassileva, V.; Parizot, B.; Demeulenaere, M.; Grunewald, W.; Audenaert, D.; Campenhout, J.V.; Overvoorde, P.; Jansen, L.; Vanneste, S. A novel Aux/IAA28 signaling cascade activates GATA23-dependent specification of lateral root founder cell identity. Curr. Biol. 2010, 20, 1697–1706. [Google Scholar] [CrossRef]
  250. Talmor-Neiman, M.; Stav, R.; Frank, W.; Voss, B.; Arazi, T. Novel micro-RNAs and intermediates of micro-RNA biogenesis from moss. Plant J. 2006, 47, 25–37. [Google Scholar] [CrossRef]
  251. Liang, C.; Wang, W.; Ma, J.; Wang, J.; Zhou, F.; Li, W.; Yu, Y.; Zhang, L.; Huang, W.; Huang, X. Identification of differentially expressed microRNAs of sunflower seedlings under drought stress. Agron. J. 2020, 112, 2472–2484. [Google Scholar] [CrossRef]
  252. Banerjee, S.; Sirohi, A.; Ansari, A.A.; Gill, S.S. Role of small RNAs in abiotic stress responses in plants. Plant Gene 2017, 11, 180–189. [Google Scholar] [CrossRef]
  253. Das, R.; Mondal, S.K. Plant miRNAs: Biogenesis and its functional validation to combat drought stress with special focus on maize. Plant Gene 2021, 27, 100294. [Google Scholar] [CrossRef]
  254. Upadhyay, U.; Singh, P.; Verma, O.P. Role of microRNAs in regulating drought stress tolerance in maize. J. Pharmacogn. Phytochem. 2019, 8, 328–331. [Google Scholar]
  255. Niu, C.; Li, H.; Jiang, L.; Yan, M.; Li, C.; Geng, D.; Xie, Y.; Yan, Y.; Shen, X.; Chen, P.; et al. Genome-wide identification of drought-responsive microRNAs in two sets of malus from interspecific hybrid progenies. Hortic. Res. 2019, 6, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Chakraborty, A.; Viswanath, A.; Malipatil, R.; Rathore, A.; Thirunavukkarasu, N. Structural and functional characteristics of miRNAs in five strategic millet species and their utility in drought tolerance. Front. Genet. 2020, 11, 1–15. [Google Scholar] [CrossRef]
  257. Singh, U.; Khemka, N.; Rajkumar, M.S.; Garg, R.; Jain, M. PLncPRO for prediction of long non-coding RNAs (lncRNAs) in plants and its application for discovery of abiotic stress-responsive lncRNAs in rice and chickpea. Nucleic Acids Res. 2017, 45, 22. [Google Scholar] [CrossRef]
  258. Liu, X.; Zhang, X.; Sun, B.; Hao, L.; Liu, C.; Zhang, D.; Tang, H.; Li, C.; Li, Y.; Shi, Y.; et al. Genome-wide identification and comparative analysis of drought-related microRNAs in two maize inbred lines with contrasting drought tolerance by deep sequencing. PLoS ONE 2019, 14, e0219176. [Google Scholar] [CrossRef]
  259. Qi, X.; Xie, S.; Liu, Y.; Yi, F.; Yu, J. Genome-wide annotation of genes and noncoding RNAs of foxtail millet in response to simulated drought stress by deep sequencing. Plant Mol. Biol. 2013, 83, 459–473. [Google Scholar] [CrossRef]
  260. Muthusamy, M.; Uma, S.; Backiyarani, S.; Saraswathi, M.S. Genome-wide screening for novel, drought stress-responsive long non-coding RNAs in drought-stressed leaf transcriptome of drought-tolerant and -susceptible banana (Musa spp.) cultivars using Illumina high-throughput sequencing. Plant Biotechnol. Rep. 2015, 9, 279–286. [Google Scholar] [CrossRef]
  261. Wu, C.; Ding, Z.; Chen, M.; Yang, G.; Tie, W.; Yan, Y.; Zeng, J.; He, G.; Hu, W. Identification and functional prediction of lncRNAs in response to PEG and ABA treatment in cassava. Environ. Exp. Bot. 2019, 166, 103809. [Google Scholar] [CrossRef]
  262. Suksamran, R.; Saithong, T.; Thammarongtham, C.; Kalapanulak, S. Genomic and transcriptomic analysis identified novel putative cassava lncRNAs involved in cold and drought stress. Genes 2020, 11, 366. [Google Scholar] [CrossRef] [Green Version]
  263. Wang, J.; Lin, J.; Kan, J.; Wang, H.; Li, X.; Yang, Q.; Li, H.; Chang, Y. Genome-wide identification and functional prediction of novel drought-responsive lncRNAs in Pyrus betulifolia. Genes 2018, 9, 311. [Google Scholar] [CrossRef] [Green Version]
  264. Eom, S.H.; Lee, H.J.; Lee, J.H.; Wi, S.H.; Kim, S.K.; Hyun, T.K. Identification and functional prediction of drought-responsive long non-coding RNA in tomato. Agronomy 2019, 9, 629. [Google Scholar] [CrossRef] [Green Version]
  265. Lv, Y.; Liang, Z.; Ge, M.; Qi, W.; Zhang, T.; Lin, F.; Peng, Z.; Zhao, H. Genome-wide identification and functional prediction of nitrogen-responsive intergenic and intronic long non-coding RNAs in maize (Zea mays L.). BMC Genom. 2016, 17, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Han, Y.; Li, X.; Yan, Y.; Duan, M.-H.; Xu, J.-H. Identification, characterization and functional prediction of circular RNAs in maize. Mol. Genet. Genom. 2020, 295, 491–503. [Google Scholar] [CrossRef]
  267. Xu, J.; Wang, Q.; Freeling, M.; Zhang, X.; Xu, Y.; Mao, Y.; Tang, X.; Wu, F.; Lan, H.; Cao, M.; et al. Natural antisense transcripts are significantly involved in regulation of drought stress in maize. Nucleic Acids Res. 2017, 45, 5126–5141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. 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]
  269. Zhou, L.; Liu, Y.; Liu, Z.; Kong, D.; Duan, M.; Luo, L. Genome-wide identification and analysis of drought-responsive microRNAs in Oryza sativa. J. Exp. Bot. 2010, 61, 4157–4168. [Google Scholar] [CrossRef] [PubMed]
  270. Zhang, F.; Luo, X.; Zhou, Y.; Xie, J. Genome-wide identification of conserved microRNA and their response to drought stress in Dongxiang wild rice (Oryza rufipogon Griff.). Biotechnol. Lett. 2016, 38, 711–721. [Google Scholar] [CrossRef] [PubMed]
  271. Zhao, B.; Liang, R.; Ge, L.; Li, W.; Xiao, H.; Lin, H.; Ruan, K.; Jin, Y. Identification of drought-induced microRNAs in rice. Biochem. Biophys. Res. Commun. 2007, 354, 585–590. [Google Scholar] [CrossRef]
  272. Fard, E.M.; Ghabooli, M.; Mehri, N.; Bakhshi, B. Regulation of miR159 and miR396 mediated by Piriformospora indica confer drought tolerance in rice. J. Plant Mol. Breed. 2017, 5, 10–18. [Google Scholar] [CrossRef]
  273. Liu, X.; Dong, X.; Liu, Z.; Shi, Z.; Jiang, Y.; Qi, M.; Xu, T.; Li, T. Repression of ARF10 by microRNA160 plays an important role in the mediation of leaf water loss. Plant Mol. Biol. 2016, 92, 313–336. [Google Scholar] [CrossRef]
  274. Ramachandran, P.; Wang, G.; Augstein, F.; De Vries, J.; Carlsbecker, A. Continuous root xylem formation and vascular acclimation to water deficit involves endodermal ABA signaling via miR165. Development 2018, 145, 1–7. [Google Scholar] [CrossRef] [Green Version]
  275. Yan, J.; Zhao, C.; Zhou, J.; Yang, Y.; Wang, P.; Zhu, X.; Tang, G.; Bressan, R.A.; Zhu, J.-K. The miR165/166 mediated regulatory module plays critical roles in ABA homeostasis and response in Arabidopsis thaliana. PLoS Genet. 2016, 12, e1006416. [Google Scholar] [CrossRef] [PubMed]
  276. Zhang, J.; Zhang, H.; Srivastava, A.K.; Pan, Y.; Bai, J.; Fang, J.; Shi, H.; Zhu, J.-K. Knockdown of rice microRNA166 confers drought resistance by causing leaf rolling and altering stem xylem development. Plant Physiol. 2018, 176, 2082–2094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Kinoshita, N.; Wang, H.; Kasahara, H.; Liu, J.; MacPherson, C.; Machida, Y.; Kamiya, Y.; Hannah, M.A.; Chua, N.-H. IAA-Ala Resistant3, an evolutionarily conserved target of miR167, mediates Arabidopsis root architecture changes during high osmotic stress. Plant Cell 2012, 24, 3590–3602. [Google Scholar] [CrossRef] [Green Version]
  278. Du, Q.; Zhao, M.; Gao, W.; Sun, S.; Li, W.X. microRNA/microRNA* complementarity is important for the regulation pattern of NFYA5 by miR169 under dehydration shock in Arabidopsis. Plant J. 2017, 91, 22–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Ma, C.; Burd, S.; Lers, A. miR408 is involved in abiotic stress responses in Arabidopsis. Plant J. 2015, 84, 169–187. [Google Scholar] [CrossRef]
  280. Qiu, C.W.; Liu, L.; Feng, X.; Hao, P.F.; He, X.; Cao, F.; Wu, F. Genome-wide identification and characterization of drought stress responsive microRNAs in Tibetan wild barley. Int. J. Mol. Sci. 2020, 21, 2795. [Google Scholar] [CrossRef] [Green Version]
  281. Jatan, R.; Chauhan, P.S.; Lata, C. Pseudomonas putida modulates the expression of miRNAs and their target genes in response to drought and salt stresses in chickpea (Cicer arietinum L.). Genomics 2019, 1114, 509–519. [Google Scholar] [CrossRef]
  282. Hajyzadeh, M.; Turktas, M.; Khawar, K.M.; Unver, T. miR408 overexpression causes increased drought tolerance in chickpea. Gene 2015, 555, 186–193. [Google Scholar] [CrossRef] [PubMed]
  283. Zhou, M.; Li, D.; Li, Z.; Hu, Q.; Yang, C.; Zhu, L.; Luo, H. Constitutive expression of a miR319 gene alters plant development and enhances salt and drought tolerance in transgenic Creeping bentgrass. Plant Physiol. 2013, 161, 1375–1391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Li, C.; Li, Y.; Bai, L.; He, C.; Yu, X. Dynamic expression of miRNAs and their targets in the response to drought stress of grafted cucumber seedlings. Hortic. Plant J. 2016, 2, 41–49. [Google Scholar] [CrossRef] [Green Version]
  285. Li, B.; Qin, Y.; Duan, H.; Yin, W.; Xia, X. Genome-wide characterization of new and drought stress responsive microRNAs in Populus euphratica. J. Exp. Bot. 2011, 62, 3765–3779. [Google Scholar] [CrossRef] [Green Version]
  286. Wei, L.; Zhang, D.; Xiang, F.; Zhang, Z. Differentially expressed miRNAs potentially involved in the regulation of defense mechanism to drought stress in maize seedlings. Int. J. Plant Sci. 2009, 170, 979–989. [Google Scholar] [CrossRef]
  287. Seeve, C.M.; Sunkar, R.; Zheng, Y.; Liu, L.; Liu, Z.; McMullen, M.; Nelson, S.; Sharp, R.E.; Oliver, M.J. Water-deficit responsive microRNAs in the primary root growth zone of maize. BMC Plant Biol. 2019, 19, 1–16. [Google Scholar] [CrossRef] [Green Version]
  288. Aravind, J.; Rinku, S.; Pooja, B.; Shikha, M.; Kaliyugam, S.; Mallikarjuna, M.G.; Kumar, A.; Rao, A.R.; Nepolean, T. Identification, characterization and functional validation of drought-responsive microRNAs in subtropical maize inbreds. Front. Plant Sci. 2017, 8, 941. [Google Scholar] [CrossRef] [Green Version]
  289. Mica, E.; Gianfranceschi, L.; Pe, M.E. Characterization of five microRNA families in maize. J. Exp. Bot. 2006, 57, 2601–2612. [Google Scholar] [CrossRef] [Green Version]
  290. Lu-yang, H.; Xu-yang, L.; Xiao-Jing, Z.; Bao-Cheng, S.; Cheng, L.; Deng-Feng, Z.; Huai-Jun, T.; Chun-hui, L.; Yong-Xiang, L.; Yun-Su, S. Genome-wide identification and comparative analysis of drought related genes in roots of two maize inbred lines with contrasting drought tolerance by RNA sequencing. J. Integr. Agric. 2020, 19, 449–464. [Google Scholar] [CrossRef]
  291. Bakhshi, B.; Fard, E.M.; Salekdeh, G.H.; Bihamta, M.R. Evaluation of miR398 differential expression in rice under drought stress condition. Bull. Georg. Natl. Acad. Sci. 2014, 8, 87–92. [Google Scholar]
  292. Sepúlveda-García, E.B.; Pulido-Barajas, J.F.; Huerta-Heredia, A.A.; Peña-Castro, J.M.; Liu, R.; Barrera-Figueroa, B.E. Differential expression of maize and teosinte micrornas under submergence, drought and alternated stress. Plants 2020, 9, 1367. [Google Scholar] [CrossRef] [PubMed]
  293. Arshad, M.; Feyissa, B.A.; Amyot, L.; Aung, B.; Hannoufa, A. MicroRNA156 improves drought stress tolerance in alfalfa (Medicago sativa) by silencing SPL13. Plant Sci. 2017, 258, 122–136. [Google Scholar] [CrossRef] [PubMed]
  294. Wang, T.; Chen, L.; Zhao, M.; Tian, Q.; Zhang, W.H. Identification of drought-responsive microRNAs in Medicago truncatula by genome-wide high-throughput sequencing. BMC Genom. 2011, 12, 8–13. [Google Scholar] [CrossRef] [Green Version]
  295. Singh, S.; Kumarb, A.; Pandab, D.; Modi, M.K.; Sena, P. Identification and characterization of drought responsive miRNAs from a drought tolerant rice genotype of Assam. Plant Gene 2020, 21, 100213. [Google Scholar] [CrossRef]
  296. Fantao, Z.; Yuan, L.; Meng, Z.; Yi, Z.; Hongping, C.; Biaolin, H.; Jiankun, X. Identification and characterization of drought stress- responsive novel microRNAs in Dongxiang wild rice. Rice Sci. 2018, 25, 175–184. [Google Scholar] [CrossRef]
  297. Gentile, A.; Ferreira, T.H.; Mattos, R.S.; Dias, L.I.; Hoshino, A.A.; Carneiro, M.S.; Souza, G.M.; Tercílio, C., Jr.; Nogueira, R.M.; Endres, L.; et al. Effects of drought on the microtranscriptome of field-grown sugarcane plants. Planta 2013, 237, 783–798. [Google Scholar] [CrossRef] [Green Version]
  298. Balyan, S.; Kumar, M.; Mutum, R.D.; Raghuvanshi, U.; Agarwal, P.; Mathur, S.; Raghuvanshi, S. Identification of miRNA-mediated drought responsive multi-tiered regulatory network in drought tolerant rice, Nagina. Sci. Rep. 2017, 71, 1–17. [Google Scholar] [CrossRef]
  299. Trindade, I.; Capitao, C.; Dalmay, T.; Fevereiro, M.P.; Santos, D.M. miR398 and miR408 are up regulated in response to water deficit in Medicago truncatula. Planta 2010, 231, 705–716. [Google Scholar] [CrossRef]
  300. Shuai, P.; Liang, D.; Zhang, Z.; Yin, W.; Xia, X. Identification of drought-responsive and novel Populus trichocarpa microRNAs by high-throughput sequencing and their targets using degradome analysis. BMC Genom. 2013, 14, 1. [Google Scholar] [CrossRef] [Green Version]
  301. Ren, Y.; Chen, L.; Zhang, Y.; Kang, X.; Zhang, Z.; Wang, Y. Identification of novel and conserved Populus tomentosa microRNA as components of a response to water stress. Funct. Integr. Genom. 2012, 12, 327–339. [Google Scholar] [CrossRef]
  302. Pagliarani, C.; Vitali, M.; Ferrero, M.; Vitulo, N.; Incarbone, M.; Lovisolo, C.; Valle, G.; Schubert, A. The accumulation of miRNAs differentially modulated by drought stress is affected by grafting in grapevine. Plant Physiol. 2017, 173, 2180–2195. [Google Scholar] [CrossRef] [Green Version]
  303. Kong, Y.; Elling, A.A.; Chen, B.; Deng, X. Differential Expression of microRNAs in maize inbred and hybrid lines during salt and drought stress. Am. J. Plant Sci. 2010, 01, 69–76. [Google Scholar] [CrossRef] [Green Version]
  304. Li, J.; Fu, F.; An, M.; Zhou, S.; She, Y.; Li, W. Differential expression of microRNAs in response to drought stress in maize. J. Integr. Agric. 2013, 12, 1414–1422. [Google Scholar] [CrossRef]
  305. Ferdous, J.; Hussain, S.S.; Shi, B.J. Role of microRNAs in plant drought tolerance. Plant Biotechnol. J. 2015, 13, 293–305. [Google Scholar] [CrossRef] [PubMed]
  306. Pradhan, A.; Naik, N.; Sahoo, K.K. RNAi mediated drought and salinity stress tolerance in plants. Am. J. Plant Sci. 2015, 06, 1990–2008. [Google Scholar] [CrossRef] [Green Version]
  307. Visentin, I.; Pagliarani, C.; Deva, E.; Caracci, A.; Turečková, V.; Novák, O.; Lovisolo, C.; Schubert, A.; Cardinale, F. A novel strigolactone-miR156 module controls stomatal behaviour during drought recovery. Plant Cell Environ. 2020, 43, 1613–1624. [Google Scholar] [CrossRef] [PubMed]
  308. Li, W.X.; Oono, Y.; Zhu, J.; He, X.J.; Wu, J.M.; Iida, K.; Lu, X.Y.; Cui, X.; Jin, H.; Zhu, J.K. The Arabidopsis NFYA5 transcription factor is regulated transcriptionally and post-transcriptionally to promote drought resistance. Plant Cell 2008, 20, 2238–2251. [Google Scholar] [CrossRef] [Green Version]
  309. Yu, Y.; Ni, Z.; Wang, Y.; Wan, H.; Hu, Z.; Jiang, Q.; Sun, X.; Zhang, H. Overexpression of soybean miR169c confers increased drought stress sensitivity in transgenic Arabidopsis thaliana. Plant Sci. 2019, 285, 68–78. [Google Scholar] [CrossRef]
  310. Zhang, X.; Zou, Z.; Gong, P.; Zhang, J.; Ziaf, K.; Li, H.; Xiao, F.; Ye, Z. Over-expression of microRNA169 confers enhanced drought tolerance to tomato. Biotechnol. Lett. 2011, 33, 403–409. [Google Scholar] [CrossRef] [PubMed]
  311. Liu, Q.; Wang, Z.; Yu, S.; Li, W.; Zhang, M.; Yang, J.; Li, D.; Yang, J.; Li, C. Pu-miR172d regulates stomatal density and water-use efficiency via targeting PuGTL1 in poplar. J. Exp. Bot. 2021, 72, 1370–1383. [Google Scholar] [CrossRef]
  312. Arshad, M.; Gruber, M.Y.; Hannoufa, A. Transcriptome analysis of microRNA156 overexpression alfalfa roots under drought stress. Sci. Rep. 2018, 8, 1–13. [Google Scholar] [CrossRef] [PubMed]
  313. Pieczynski, M.; Marczewski, W.; Hennig, J.; Dolata, J.; Bielewicz, D.; Piontek, P.; Wyrzykowska, A.; Krusiewicz, D.; Strzelczyk-Zyta, D.; Konopka-Postupolska, D.; et al. Down regulation of CBP80 gene expression as a strategy to engineer a drought-tolerant potato. Plant Biotech. J. 2013, 11, 459–469. [Google Scholar] [CrossRef] [PubMed]
  314. Yue, E.; Cao, H.; Liu, B. OsmiR535, a potential genetic editing target for drought and salinity stress tolerance in Oryza sativa. Plants 2020, 9, 1377. [Google Scholar] [CrossRef] [PubMed]
  315. Quattro, C.; De Enrico Pe, M.; Bertolini, E. Long noncoding RNAs in the model species Brachypodium distachyon. Sci. Rep. 2017, 7, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Zhang, H.; Tao, Z.; Hong, H.; Chen, Z.; Wu, C.; Li, X.; Xiao, J.; Wang, S. Transposon-derived small RNA is responsible for modified function of WRKY45 locus. Nat. Plants 2016, 2, 3. [Google Scholar] [CrossRef]
  317. Nosaka, M.; Itoh, J.I.; Nagato, Y.; Ono, A.; Ishiwata, A.; Sato, Y. Role of transposon-derived small RNAs in the interplay between genomes and parasitic DNA in rice. PLoS Genet. 2012, 8, e1002953. [Google Scholar] [CrossRef]
  318. Sharma, N.; Kumar, S.; Sanan-Mishra, N. Osa-miR820 regulatory node primes rice plants to tolerate salt stress in an agronomically advantageous manner. bioRxiv 2021. [Google Scholar] [CrossRef]
  319. Creasey, K.M.; Zhai, J.; Borges, F.; Van Ex, F.; Regulski, M.; Meyers, B.C.; Martienssen, R.A. miRNAs trigger widespread epigenetically activated siRNAs from transposons in Arabidopsis. Nature 2014, 508, 411–415. [Google Scholar] [CrossRef] [Green Version]
  320. Li, Y.; Li, C.; Xia, J.; Jin, Y. Domestication of transposable elements into microRNA genes in plants. PLoS ONE 2011, 6, e19212. [Google Scholar] [CrossRef] [Green Version]
  321. Pant, B.D.; Buhtz, A.; Kehr, J.; Scheible, W.R. MicroRNA399 is a long-distance signal for the regulation of plant phosphate homeostasis. Plant J. 2008, 53, 731–738. [Google Scholar] [CrossRef] [Green Version]
  322. Huen, A.K.; Rodriguez-Medina, C.; Ho, A.Y.Y.; Atkins, C.A.; Smith, P.M.C. Long-distance movement of phosphate starvation-responsive microRNAs in Arabidopsis. Plant Biol. 2017, 19, 643–649. [Google Scholar] [CrossRef] [PubMed]
  323. Akdogan, G.; Tufekci, E.D.; Uranbey, S.; Unver, T. miRNA-based drought regulation in wheat. Funct. Integr. Genom. 2016, 16, 221–233. [Google Scholar] [CrossRef] [PubMed]
  324. Pagliarani, C.; Gambino, G. Small RNA mobility: Spread of RNA silencing effectors and its effect on developmental processes and stress adaptation in plants. Int. J. Mol. Sci. 2019, 20, 4306. [Google Scholar] [CrossRef] [Green Version]
  325. Pan, W.J.; Tao, J.J.; Cheng, T.; Bian, X.H.; Wei, W.; Zhang, W.K.; Ma, B.; Chen, S.Y.; Zhang, J.S. Soybean miR172a improves salt tolerance and can function as a long-distance signal. Mol. Plant 2016, 9, 1337–1340. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic representation of different drought stress-response mechanisms operative in plants. The combined actions of these processes govern plant survival or susceptibility. The survival of plants can be grouped under drought avoidance, drought tolerance and drought escape, based on the plant response. ABA = abscisic acid, CSI = chlorophyll stability index, MSI = membrane stability index, PPO = polyphenol oxidase, ROS = reactive oxygen species, RWC = relative water content.
Figure 1. Schematic representation of different drought stress-response mechanisms operative in plants. The combined actions of these processes govern plant survival or susceptibility. The survival of plants can be grouped under drought avoidance, drought tolerance and drought escape, based on the plant response. ABA = abscisic acid, CSI = chlorophyll stability index, MSI = membrane stability index, PPO = polyphenol oxidase, ROS = reactive oxygen species, RWC = relative water content.
Ijms 22 12519 g001
Figure 2. Generalized scheme to illustrate the various steps involved in plant miRNA biogenesis and activity. The miRNA genes are transcribed to primary miRNA transcripts (pri-miRNAs) of 100–120 nt long by RNA polymerase II that are then capped and polyadenylated. In the first maturation step of miRNA synthesis, pri-miRNA is cleaved by DCL1 in collaborative action of other enzymes to form precursor miRNA (pre-miRNA) of approximately 70–100 nt long. In the second maturation step, the hairpin structured pre-miRNA is processed by the same enzymes to mature miRNA duplex. The DCL1 cleavage results in the formation of a miRNA–miRNA* duplex of 21 to 24 nt containing two nucleotide 3′ overhangs and 5′ monophosphate regions. This duplex is transported out of the nucleus and into the cytoplasm, where it associates with an AGO (Argonaute) containing protein complex to form RNA induced silencing complex (RISC). The RISC is guided to the target sequence of single-stranded mRNA. Depending upon the nature of miRNA and AGO, the target mRNA is either cleaved or subjected to suppression of translation.
Figure 2. Generalized scheme to illustrate the various steps involved in plant miRNA biogenesis and activity. The miRNA genes are transcribed to primary miRNA transcripts (pri-miRNAs) of 100–120 nt long by RNA polymerase II that are then capped and polyadenylated. In the first maturation step of miRNA synthesis, pri-miRNA is cleaved by DCL1 in collaborative action of other enzymes to form precursor miRNA (pre-miRNA) of approximately 70–100 nt long. In the second maturation step, the hairpin structured pre-miRNA is processed by the same enzymes to mature miRNA duplex. The DCL1 cleavage results in the formation of a miRNA–miRNA* duplex of 21 to 24 nt containing two nucleotide 3′ overhangs and 5′ monophosphate regions. This duplex is transported out of the nucleus and into the cytoplasm, where it associates with an AGO (Argonaute) containing protein complex to form RNA induced silencing complex (RISC). The RISC is guided to the target sequence of single-stranded mRNA. Depending upon the nature of miRNA and AGO, the target mRNA is either cleaved or subjected to suppression of translation.
Ijms 22 12519 g002
Table 1. Functions of lncRNAs identified in various plants.
Table 1. Functions of lncRNAs identified in various plants.
PlantlncRNA NamePathwayFunctional AnnotationReferences
ArabidopsisIPS1 lncRNAPhosphate homeostasisTarget mimic for miR399, which regulates PHO2, a negative regulator of the phosphate transporters [112]
Hidden Treasure 1 (HID1)PhotomorphogenesisPromotes photomorphogenesis in continuous red light by directly repressing Phytochrome Interacting Factor 3 (PIF3)[100]
ASCO-lncRNA and NSRAlternate splicing moduleRecognizes alternatively spliced mRNA targets[111]
Drought Induced lncRNAs (DRIR)Drought response Positively regulates several drought responsive transcripts such as ABA-signal transducers (P5CS1, RD29A,B and ABI5); annexins (ANNAT7) and aquaporins (TIP4, NIP1) [113]
Brassica napus, (Q2 and Qinyou8)XLOC_042431 and XLOC_071559Hormone signalingTargets BnaC06g05090D gene to regulate ethylene metabolism, IAA, Cytokinins and ABA signaling [114]
XLOC_ 095305 and XLOC_100682Targets BnaA01g17750D genes to regulate alpha trehalose phosphate synthase
Cassava (TMS60444 and Ku50)lincRNA340Target mimicrymiR169 target mimicry, also targets Nuclear Factor Y (NF-Y)[115]
TCONS_00003360, TCONS_00015102, Signal transductionCalcium and ABA signaling [116]
TCONS_00149293Ethylene metabolism
TCONS_00097416 Hormone signaling and target mimicryTargets CSLD5, ERL1 and SPCH genes to modulate ethylene signaling;
TCONS_00069665Targets LAX2, HDG11 and SCR genes; and regulates expression by targeting miR156
TCONS_00060863 TCONS_00068353Targets CYP707A1 gene and regulates in ABA catabolism
TCONS_00040721MiRNA targetTargets GRF1, HB51 and DOX1; regulates gene expression by targeting miR156, miR164, miR169 and miR172
Cleistogenes songoricaMSTRG.25585.13Metabolic pathway Regulates sucrose metabolism[97]
MSTRG.42613.1Regulates starch metabolism
MSTRG.43964.1, MSTRG.4400.2Hormone signaling and target mimicryTargets ABA pathway and related genes, regulates miR164, miR166, miR393 and miR397a,b and act as endogenous target mimic
Panicum virgatum (Alamo) XLOC_033252Hormone signalingRegulates ABA synthesis and signaling by targeting Pavir.Eb01847 gene[117]
Populus trichocarpa (Nisqually 1) lincRNA20, lincRNA2752, lincRNA2962, lincRNA1039, lincRNA3241miRNA regulationControl drought stress by regulating ptc-miR476 and ptc-miR169[118]
Oryza sativa (DXWR)lncRNAMSTRG69391Transcription regulationRegulates biological process by targeting genes encoding calmodulin[96]
lncRNA MSTRG41712 and MSTRG68635Translation inhibition Targeting genes encoding heat shock protein and mitochondrial carrier proteins
lncRNA MSTRG65848, MSTRG27834 and MSTRG46301Differential regulation Upregulated and downregulated the lncRNAs themselves; response to drought stress and targets several genes
Oryza sativa cv (Ilmi)NATOs02g0250700–01 andOs02g0180800–01Regulate response to drought by targeting Os02g0250600-01 (encodes highly abundant protein during late embryogenesis) and Os02g0180700-01 (encodes Cinnamoyl-CoA reductase)[119]
Triticum aestivum(Kiziltan and TR39477)c70772_g2_i1 and c90557_g1_i1lncRNA-miRNA-mRNA networkTargets c69036_g1_i1 and c9653_g1_i2 genes to regulate drought stress [120]
Zea maysLi_TCONS_00080887, Zhang_TCONS_00012690, Zhang_TCONS_00012690 625-646, Boerner_Z27kG1_14953, Boerner_Z27kG1_09751, Boerner_Z27kG1_15115, Boerner_Z27kG1_08283, Boerner_Z27kG1_16361, Boerner_Z27kG1_23317, Boerner_Z27kG1_13892, Boerner_Z27kG1_01046, Boerner_Z27kG1_22106, Boerner_Z27kG1_03819, Boerner_Z27kG1_17085, Boerner_Z27kG1_06707, Boerner_Z27kG1_17308, Boerner_Z27kG1_01291, Boerner_Z27kG1_22188, Boerner_Z27kG1_15675, Boerner_Z27kG1_06005, Zhang_TCONS_00011169, etc.miRNA targets or decoysTargets or decoys of zma-miR156e-3p, zma-miR156h-3p, zma-miR159c,d-3p, zma-miR159e-5p, zma-miR159e-5p, zma-miR160b,g-3p, zma-miR160c-3p, zma-miR160c-3p, zma-miR162-5p, zma-miR164b-3p, zma-miR164d-3p, zma-miR164e-3p, zma-miR166h-5p, zma-miR166i-5p, zma-miR166i-5p, zma-miR166n-5p, zma-miR169c-3p, zma-miR169f-3p, zma-miR169l-3p, zma-miR169m-3p, respectively, etc.[59]
Miniature inverted-repeat transposable element (MITE-ZmNAC111)RNA-directed DNA methylation Represses ZmNAC111 expression and enhances drought tolerance[121]
lncRNAMSTRG6838.1Transcription regulationTargets V-ATPase- and VPP4-encoding genes and regulates transcription[122]
ZmPHO2, PILNCR1Phosphate homeostasisTargets of Zma-miR399 in response to low phosphate[123]
Table 3. Some drought-responsive lncRNAs reported in different plants.
Table 3. Some drought-responsive lncRNAs reported in different plants.
PlantNumber of Putative lncRNAs IdentifiedPlatform of IdentificationFunctional AnnotationReferences
Arabidopsis thaliana303qRT-PCRResponsive to heat, cold, drought and salt stress[94]
Hidden Treasure 1 (HID1)Northern blottingPromote photomorphogenesis in continuous red light by directly repressing PIF3[100]
13,230Transcriptome Analysis, published tiling array datasetsResponse to drought, cold, high-salt and/or ABA treatments[98]
Banana8471Transcriptome Analysis, HiSeqDrought stress-response[260]
Cassava682HiSeq 2500, qRT-PCR, CNCI, CPCHormone signal transduction, sucrose metabolism pathway, etc.[115]
124qRT-PCRMelatonin responsive, drought stress regulation, cellular metabolism, Calvin cycle, hormone regulation, etc.[116]
1379qRT-PCRDifferent roles[261]
56,840RNA-Seq Transcriptome AnalysisDifferential expression in cold or drought conditions[262]
Chickpea3457RT-qPCR, PLncPRODifferentially expressed under drought stress[257]
Cleistogenes songorica3397HiSeq2500, CPC, CNCI, CPATqRT-PCRRegulate drought stress response[97]
Dimocarpus longan Lour7643Real-time qPCREarly somatic embryogenesis[134]
Oryza sativa98HiSeq 2500, qRT- PCRRegulatory role in drought response[119]
3714RT-qPCR, PLncPRODifferentially expressed under drought stress[242]
Panicum virgatum L16,551HiSeq2500, qRT-PCRRegulate drought stress response[117]
Populus trichocarpa504HiSeq™ 2000, RT-qPCRDrought- stress response, putative targets and target mimics of miRNAs[118]
Pyrus betulifolia251HiSeq 4000, CNCI, CPC, qRT-PCRRegulate various metabolic processes [263]
Setaria italica19HiSeq 2000, qRT- PCRControl drought stress response[259]
Solanum lycopersicum521RT-qPCRVariety of biological processes via lncRNA-mRNA co-expression[264]
Triticum aestivum (Kiziltan, TR39477 and TTD-22 varieties)59,110, 57,944 and 40,858HiSeq 2000, qRT- PCRDifferential expression under drought stress response in cultivated and wild varieties[120]
Zea mays1724RT-qPCRRegulatory role in drought response[102]
637Ribosomal RNA depletion and ultra-deep total RNA sequencingRegulatory roles in response to N stress[265]
1535HiSeq 2500, qRT- PCROxidoreductase activity, water binding and electron carrier activity[122]
1199RiboMinus RNA-SeqControl drought and salt stress[266]
1769Strand-specific RNA sequencing,NATs in drought stress response[267]
Table 4. Drought-responsive miRNAs reported in different crop plants.
Table 4. Drought-responsive miRNAs reported in different crop plants.
Plant NamemiRNAsTargetTarget DescriptionReferences
ArabidopsismiR160 ARF[273]
miR165/166 HD-ZIPIII, CLP-1, RDD1, ABA signaling[274,275,276]
miR167 IAR3[277]
miR169 NFY-A, HAP2[278]
miR408 LAC[279]
BarleymiR397aMLOC_54246.3LAC-23[280]
miR399MLOC_52822.6Phosphatase 2
Novel-m0406-3pMLOC_70587.1PHD finger protein
LOC_50162.1Sucrose synthase 1
MLOC_67419.2PBS1, Ser/Thr-protein kinas,
MLOC_67450.11D27, beta-carotene isomerase
MLOC_73965.1Homocysteine S methyltransferase 3
Novel-m0598-3pMLOC_34795.2RNA polymerase (25-kDa subunit)
Novel-m0624-3pMLOC_55820.2Pectinesterase
Novel-m0793-3pMLOC_52822.6Phosphatase 2
Novel-m1587-5pMLOC_56261.3ABC transporter C family member 2-like
Novel-m1738-3pMLOC_3895.3Dro1 (coding for early auxin response protein)
Novel-m1900-5pMLOC_16998.3Glycine-rich RNA-binding protein 10
Novel-m2311-5pMLOC_61629.2Transcription elongation factor, SPT6
Novel-m2328-3pMLOC_6972.2DNA crosslink repair 1A protein
ChickpeamiR159 GA-MYB-like[281]
miR160 ARF 16 (Seed germination and post germination stages)
miR166 ATHB-15 (axillary meristem initiation, leaf and vascular development)
miR167 ABI 5 (Gynoecium and stamen development)
miR169 NFY-A (plant development and flowering timing; response to different biotic stresses)
miR171 NSP2 (response to abiotic stresses and floral development)
miR172 RAP2-7 (flowering time, floral organ identity and cold stress response)
miR393 AFB2 (susceptibility to virulent bacteria)
miR396 CP29 (leaf and cotyledon development)
miR408 Plantacyanin (regulation of DREB and other drought responsive gene)[282]
Creeping bentgrassmiR319 TCP[283]
Cucumis sativusmiR159T159MYB protein 306-like[284]
miR167T167ARF 8-like
miR170T170GRAS transcription factor
miR172T172Floral homeotic protein, APETALA 2-like
miR319T319Transcription factor, MYB75-like
b-miR-n-07TB7ATPase
b-miR-n10TB10GRAS transcription factor
b-miR-n24TB24DELLA protein GAI1-like
miR169T169NFY-A-1-like
miR395T395ATP sulfurylase 1
miR398T398Superoxide dismutase
csa-miR-n19TC19Pleiotropic drug resistance protein 2-like
miR168T168Argonaute 1A-like
miR396T396Endoribonuclease dicer homolog 1-like
b-miR-n02TB2Pre-mRNA-processing factor 17-like
b-miR-n20TB20Dicer-like protein 4-like
Euphrates poplarmiR30a,beugene3.00010640Electron carrier activity[285]
miR71*eugene3.00010640Electron carrier activity
grail3.0008024501Electron carrier activity
eugene3.105640001Electron carrier activity
fgenesh4_pg.C_scaffold_263000013Electron carrier activity
miR77eugene3.00002056Electron carrier activity
estExt_Genewise1_v1.C_LG_XIV3469Electron carrier activity
miR84*fgenesh4_pm.C_LG_XIII000061Electron carrier activity
miR101agw1.I.9350.1Transcription factor
miR131eugene3.00120942Electron carrier activity
fgenesh4_pg.C_LG_X001404DNA binding
estExt_Genewise1_v1.C_LG_XV2187Electron carrier activity
fgenesh4_pg.C_scaffold_9189000001Electron carrier activity
fgenesh4_pg.C_LG_II001303DNA binding
miR58 estExt_Genewise1_v1.C_LG_XV2187SBP-box Transcription factor
miR67*gw1.VIII.1137.1Function unknown
eugene3.00031501Vesicle transport v-SNARE
miR93agrail3.0010018301Function unknown
estExt_Genewise1_v1.C_LG_IV3721NADH-ubiquinone oxidoreductase
miR93bgrail3.0010018301Function unknown
miR106*estExt_fgenesh4_pg.C_17020003Cytochrome oxidase biogenesis protein
estExt_fgenesh4_pm.C_1230037Function unknown
miR115agw1.57.264.1Function unknown
miR123aestExt_fgenesh4_pg.C_LG_III1182Development and cell-death domain
MaizemiR156c Putative protein phosphatase 2C[286]
miR159a,b Serine/threonine protein phosphatase
miR159a-d GA-MYB transcription factor
miR160a-e S16, 40S ribosomal protein
miR160b,i ARR11, response regulator
miR166l,m Homeodomain–leucine zipper protein
miR167a-i ARF 12
miR167c ARF 17, Putative eIF3e
miR167f,g ARF 25
miR167d Phospholipase D
miR168a,b Serine/threonine-protein phosphatase
miR168b Receptor-like protein kinase
miR168a,b AGO1-1, mitogen-activated protein kinase 13
miR396a,bTC250636DEAD-box ATP-dependent RNA helicase 3,
TC251979Putative early responsive to dehydration stress protein,
TC274109GTPase,
TC259098Heat shock protein 90,
TC26999GA-MYB-binding protein
miR396d,e Putative serine/threonine protein kinase
miR398bTC248005Pyruvate, orthophosphate dikinase,
TC253981Putative protein serine/threonine kinase,
TC270251Putative selenium binding protein,
TC270802Fructose-bisphosphate aldolase
miR408 Leucine-rich repeat family protein
miR474bCF008935Putative CBL-interacting protein kinase,
TC263244Proline dehydrogenase family protein,
miR474c CF055555Putative transcription factor MYB,
CF632829WRKY transcription factor 31
miR528TC250873Cu/Zn SOD,
TC274952Peroxidase
MiR827 N/Pi metabolism
miR156a/b,c,d,e,g,h,k,lAC233751, GRMZM2G061734, GRMZM2G065451DNA-binding putative protein[159]
GRMZM2G040785),Unknown
GRMZM2G307588SPL 6
GRMZM2G414805SPL 11
GRMZM2G460544SPL 7
GRMZM2G067624Homoserine kinase
GRMZM2G465165Serine/threonine protein kinase
miR159a,b,f,cGRMZM2G167088 and GRMZM2G416652DNA-binding protein
GRMZM2G027100Unknown
AC217264MYB55
miR159a,b,f and miR319a,cGRMZM2G028054GA-MYB
miR159a,b,fGRMZM2G423833, GRMZM2G075064DNA-binding protein
miR166dAC187157MPPN domain
GRMZM2G003509Protein methyltransferase
GRMZM2G499154Metabolic process
miR167a,cGRMZM2G078274, GRMZM2G475882Hormone stimulus
miR395bGRMZM2G04217Secondary active sulfate transmembrane transporter (1)
GRMZM2G149952, GRMZM2G051270ATP sulfurylase
miR396fGRMZM2G178990Actin binding protein
miR1432GRMZM2G423139Calcium-binding allergen Ole e 8
miR1436GRMZM2G125531RNA binding protein
miR2097-5pGRMZM2G151955Serine/threonine protein kinase
mir319a-d-3pGRMZM2G089361TOlTCP family transcription factor[287]
GRMZM2G145112 T02, GRMZM2G100579 T02Putative uncharacterized protein
miR393ac-5pGRMZM2G135978 Tol, GRMZM5G848945_T02Transport inhibitor response 1-like protein
miR396cdGRMZM2G033612 T02Putative uncharacterized protein
GRMZM2G098594_ T06, GRMZM2G099862_ T04, GRMZM2G119359_T01, GRMZM5G893117 T01, GRMZM2G105335_ T02, GRMZM2G067743_T03GRF-transcription factor
GRMZM2G029323_T01AP2/EREBP transcription factor protein
miR398ab-3pGRMZM2G023847 Tol, GRMZM2G097851 TolPutative uncharacterized protein
GRMZM2G352678 T01Chemocyanin
GRMZM5G866053_T01Basic blue protein-like
GRMZM2G122302_T01, GRMZM2G082940_T01Blue copper protein
miR444abGRMZM2G492156_T01, GRMZM2G033093_T01MADS-box transcription factor
GRMZM2G005000 T02Putative uncharacterized protein
miR168a-3pGRMZM2G369839 To1Putative uncharacterized protein
miR168b-3pGRMZM2G136486 T02Putative uncharacterized protein
miR319a-d-3pGRMZM2G020805_T01TCP family transcription factor
miR390ab-3pGRMZM2Gl07498_T01Putative uncharacterized protein
miR827-3pGRMZM2G175406_T01Putative uncharacterized protein
miR399PHO2, UBC24Control Pi homeostasis[288]
miR529 SPB domain transcription factor
miR399PHO2, UBC24Control Pi homeostasis
miR529 SPB domain transcription factor
miR156SPLShoot development and delayed change in vegetative phase[288,289]
miR160 ARF (root development and auxin signals)
miR166 HD-ZIPIII (leaf development and polarity)
miR169HAP2Nitrogen homeostasis and stress response
miR395APS, ASTControl ATP Sulfurylase activity
miR171SCLRegulate root development[289]
miR172AP2Maintain nitrogen remobilization and floral development
miR167 CCAAT-binding factor, ARF
miR397 LAC (regulate copper homeostasis and reduces root growth)
miR159MYBRegulate flowering time; leaf shape and size[288]
miR162DCL1Negative feedback regulatory function[258]
miR164NAC1Control lateral root development[258,288]
miR168AGO1Nutrient homeostasis and feedback regulation[290]
miR2275gnl|GNOMON|55702013.mMitochondrial protein[254]
miR393gnl|GNOMON|39086093.mProtein transport inhibitor response 1-like
miR398CSDCopper homeostasis and oxidative stress[291]
miR156k ↓ in drought and submergence[292]
miR159ab ↑ in drought, ↓ in submergence
miR164e ↓ in drought and submergence
miR166b,d ↓ in drought and submergence
miR167c,d,e,g ↓ in drought and submergence
miR169c,r ↓ in drought and submergence
miR319b ↑ in drought, ↓ in submergence
miR396c,d ↓ in drought and submergence
miR398a,b ↓ in drought and submergence
miR398b ↓ in drought and submergence
miR408 ↓ in drought and submergence
miR408b ↓ in drought and submergence
miR528ab ↓ in drought and submergence
miR166c Constitutive expression
Medicago sativamiR156 SBP-like protein[293]
Medicago truncatulamiR164 NAC domain transcription factor (lateral root development) ↓[294]
miR169 CBF (response to drought, cold and salinity, nodule development) ↓
miR171 GRAS transcription factors (response to drought, cold and salinity, nodule
Morphogenesis and floral development) ↓
miR396 GRF (response to drought and salt; cell proliferation) ↓
miR398 Cu/Zn CSD1, CSD2
(response to oxidative stress) ↓
miR399PHO2ubiquitin conjugating enzyme
balance of phosphorus, ↑
miR2118 TIR–NBS–LRR domain protein encoding response to drought, cold, salinity and ABA, ↑
miR1510a PDC isozyme 1, concanavalin A-like lectin/glucanase 3. F-box protein, ↓
miR2089 NB–ARC domain protein, ↑
miR2111a-s,u-v Calcineurin-like phosphoesterase, membrane protein SAK, ↑
miR5274b DNA-damage-repair, toleration protein, ↑
miR5554a- c Polynucleotidyl transferase, ribonuclease
H fold, ↓
miR5558 Initiation factor eIF-4 gamma, homeodomain-related POX, ↑
Rice66 miRNAs Response to drought stress[119]
miR167, miR9774, miR398, miR162, miR319, miR156, miR408, miR166, miR531, miR827 and miR8175 ↓ expression profiling in response to drought stress[294]
miR6300, miR160, miR1861, miR440, miR9773, miR3982, miR171 and miR1876 ↑ expression profiling in response to drought stress
67 novel drought responsive miRNAs 27 novel miRNAs ↓ and 40 novel miRNAs ↑ in response to drought stress[295]
Osa-miR159f, Osa-miR1871, Osa-miR398b, Osa-miR408-3p, Osa-miR2878-5p, Osa-miR528-5p and Osa-miR397a ↑ in the flag-leaves of tolerant cultivar (N22 and Vandana, while ↓ in sensitive cultivar (PB1 and IR64) during drought[296]
miR398CSDRegulate copper homeostasis and oxidative stress[292]
SugarcaneMiR160, miR399 and miR528 ↑ in tolerant cultivar (RB867515)[297]
miR160, miR394, miR399 and miR1432 ↑ in sensitive cultivar (RB855536)
miR166, miR169, miR171, MiR172, miR393, miR396, miR399 and miR1432 ↓ in tolerant cultivar (RB867515)
miR166, miR171, miR396 ↓ in sensitive cultivar (RB855536)
SunflowermiR399a-2HannXRQ_chr02g0057111Environment adaptation; leaf ↑; root ↑[251]
Novel-mir40 4HannXRQ_chr03g0090941DNA repair protein XRCC; root ↑
Novel-mir3, Novel-mir42HannXRQ_chr04g0098561Putative toll/interleukin-1 receptor; root ↑
miR396b HannXRQ_chr04g0115781Serine/threonine protein kinase; leaf ↑
miR156a-5p,f,k,q, 157a-5pHannXRQ_chr05g0138971SBP transcription factor; leaf ↑
Novel-mir3HannXRQ_chr05g0149501P-loop containing nucleoside triphosphate hydrolase; leaf ↓
miR396a,b-5pHannXRQ_chr05g0150421Glutamyl tRNA reductase and chlorophyll metabolism; leaf ↓
miR156hHannXRQ_chr07g0196531Leaf ↓
miR396f-1HannXRQ_chr08g0211484Serine/threonine dual specificity protein kinase; root ↑
miR394a-3p-1 HannXRQ_chr08g0216701Related to Zn ion transport; leaf ↑
Novel-mir36HannXRQ_chr08g0219981Putative plant disease resistance response protein; root ↓
Novel-mir42HannXRQ_chr09g0239281Putative toll/interleukin-1 receptor homology (TIR) domain; root ↑
Novel-mir3HannXRQ_chr09g0239531P-loop containing nucleoside triphosphate hydrolase; root ↑
Novel-mir55HannXRQ_chr09g0252001C-terminal LisH motif-containing protein, Leaf ↑; root ↑
Novel-mir42HannXRQ_chr13g0396521P-loop containing nucleoside triphosphate hydrolase; root ↑
Novel-mir3HannXRQ_chr13g0396531Putative toll/interleukin-1 receptor; leaf ↓
Novel-mir65HannXRQ_chr14g0435381Root ↑
Novel-mir66HannXRQ_chr14g0435571Auxin-induced protein, leaf ↑
MiR172a-2HannXRQ_chr15g0491641Leaf ↑; root ↑
MiR156a-2HannXRQ_chr17g0534011(S)-urea glycine amidohydrolase; leaf ↑
Novel-mir17HannXRQ_chr17g0569261Probable response regulator 11; root ↓
Triticum aestivummiR156 SPL; leaf ↑; root ↑[298]
miR159 MYB transcription factor, leaf ↑; root ↓
miR160 ARF, leaf ↑; root ↑
miR162 GTPase activating protein-like; leaf ↑
miR164 NAC domain-containing protein; leaf ↑; root ↑
miR169 CCAAT-box-transcription factor; leaf ↓; root ↑
miR172 APETALA2 transcription factor; leaf ↓; root ↓
miR319 MYB transcription factor; leaf ↑; root ↓
miR396 Heat shock protein; leaf ↓; root ↓
miR398 Cu/Zn superoxide dismutase; leaf ↑; root ↓
miR482 TPGR; leaf ↑; root ↑
miR528 Glyceraldehyde-3-phosphate dehydrogenase; leaf ↑; root ↓
miR838 Small heat shock protein (Mds1); leaf ↓
miR1120 Glyceraldehyde-3-phosphate dehydrogenase; leaf ↑
miR1169 Small GTP-binding protein; root ↑
miR1436 Glutathione S-transferase; root ↑
miR1450 Manganese superoxide dismutase; leaf ↓
miR2102 Calmodulin-binding family protein; root ↑
miR4393 ARF; leaf ↑; root ↓
miR4993 SKP1/ASK1-like protein; root ↑
miR5048 RPG1, serine/threonine protein kinase; root ↓
miR5049 Wpk4 protein kinase, leaf ↑; root ↑
miR5059 Heat shock protein; root ↑
miR5075 Serine/threonine protein kinase 3; root ↑
miR5083 Hydroxymethylglutaryl-CoA synthase; leaf ↑
miR5174 NBS–LRR genes, leaf ↑; root ↑
miR5175 Methylene-tetrahydrofolate reductase; leaf ↑
miR5205 Malate dehydrogenase, CBS domain-containing protein; leaf ↑
miR5568 Pathogenesis-related protein, leaf ↑; root ↓
miR6108 Glycosyltransferase; leaf ↑
37 miRNAs including 5 novel miRNAs 27 ↑, 10 ↓[34]
Zanthoxylum bungeanummiR396a-5p Superoxide dismutase [Mn] 1, mitochondrial[14]
miR834 Superoxide dismutase [Fe], chloroplastic-like isoform X2
miR167a-3p Peroxiredoxin-2E, chloroplastic (POD)
miR169b-3p Catalase isozyme 1(CAT)
miR447a-3p L-ascorbate peroxidase 3
miR773b-3p Phospholipid hydroperoxide glutathione peroxidase 1, chloroplastic
miR397b Delta-1-pyrroline-5-carboxylate synthase, key enzyme for the synthesis of proline
miR397bJAR1Jasmonic acid-amido synthetase (participate in the synthesis of jasmonic acid)
miR859 ABSCISIC ACID–INSENSITIVE 5-like protein 5, (regulate a variety of ABA responses, such as stomatal closure, plasma membrane permeability and water permeability)
miR5632-5p Mitogen-activated protein kinase 1
miR1888a Protein disulfide-isomerase 5-2 isoform X1
miR5638a Respiratory burst oxidase homolog protein C (Citrus sinensis)
miR398a-3p Probable nucleoredoxin 1
miR3434-3p Translationally controlled tumor protein homolog; involved in the regulation of abscisic acid–mediated and calcium-mediated stomatal closure
AP2, Apetala 2; ARF, auxin response factor; CBF, CCAAT Binding Factor; CBS, cystathionine beta synthase; CLP-1, Cysteine Protease-1; CSD, copper/zinc superoxide dismutase; DCL, dicer-like protein; GTP, guanosine triphosphate; HAP2, heme activator protein 2; HD-ZIPIII, Homeodomain Leucine Zipper III; IAR3, Indole-3-Acetic Acid-Ala Resistant 3; LAC, Laccases; LMW, low molecular weight; LRR, leucine-rich repeats; MYB, myeloblastosis; NBS–LRR, nucleoside binding site–leucine-rich repeat; NFY-A, Nuclear Transcription Factor Y Subunit Alpha; NLA, nitrogen limitation adaptation; PHD, plant homeodomain; PDC, pyruvate decarboxylase; RDD1, Rice D of Daily Fluctuations 1; SBP or SPL, Squamosa promoter binding protein-like; SCL, scarecrow-like; SKP1, S-phase kinase-associated protein 1; SOD, superoxide dismutase; TCP, Teosinte Branched/Cycloidea/Proliferating Cell Factors (PCF); TPGR, transmembrane proton gradient regulation; UBC24, ubiquitin-conjugating enzyme; E2, phosphate 2.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gelaw, T.A.; Sanan-Mishra, N. Non-Coding RNAs in Response to Drought Stress. Int. J. Mol. Sci. 2021, 22, 12519. https://doi.org/10.3390/ijms222212519

AMA Style

Gelaw TA, Sanan-Mishra N. Non-Coding RNAs in Response to Drought Stress. International Journal of Molecular Sciences. 2021; 22(22):12519. https://doi.org/10.3390/ijms222212519

Chicago/Turabian Style

Gelaw, Temesgen Assefa, and Neeti Sanan-Mishra. 2021. "Non-Coding RNAs in Response to Drought Stress" International Journal of Molecular Sciences 22, no. 22: 12519. https://doi.org/10.3390/ijms222212519

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