Next Article in Journal
Extracellular Vesicles Derived from Endothelial Progenitor Cells Protect Human Glomerular Endothelial Cells and Podocytes from Complement- and Cytokine-Mediated Injury
Next Article in Special Issue
Identification of miRNA–mRNA Regulatory Modules Involved in Lipid Metabolism and Seed Development in a Woody Oil Tree (Camellia oleifera)
Previous Article in Journal
No Evidence for Seed Transmission of Tomato Yellow Leaf Curl Sardinia Virus in Tomato
Previous Article in Special Issue
Transcriptome and Degradome Profiling Reveals a Role of miR530 in the Circadian Regulation of Gene Expression in Kalanchoë marnieriana
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Non-Coding RNAs in Legumes: Their Emerging Roles in Regulating Biotic/Abiotic Stress Responses and Plant Growth and Development

1
ICAR—Indian Institute of Pulses Research (IIPR), Kanpur 208024, India
2
Department of Botany, Panjab University, Chandigarh 160014, India
3
School of Science, RMIT University, Melbourne 3083, Australia
4
The UWA Institute of Agriculture, The University of Western Australia, Perth 6001, Australia
*
Authors to whom correspondence should be addressed.
Cells 2021, 10(7), 1674; https://doi.org/10.3390/cells10071674
Submission received: 1 June 2021 / Revised: 24 June 2021 / Accepted: 28 June 2021 / Published: 2 July 2021
(This article belongs to the Special Issue Non-coding RNAs in the Growth and Development of Plants)

Abstract

:
Noncoding RNAs, including microRNAs (miRNAs), small interference RNAs (siRNAs), circular RNA (circRNA), and long noncoding RNAs (lncRNAs), control gene expression at the transcription, post-transcription, and translation levels. Apart from protein-coding genes, accumulating evidence supports ncRNAs playing a critical role in shaping plant growth and development and biotic and abiotic stress responses in various species, including legume crops. Noncoding RNAs (ncRNAs) interact with DNA, RNA, and proteins, modulating their target genes. However, the regulatory mechanisms controlling these cellular processes are not well understood. Here, we discuss the features of various ncRNAs, including their emerging role in contributing to biotic/abiotic stress response and plant growth and development, in addition to the molecular mechanisms involved, focusing on legume crops. Unravelling the underlying molecular mechanisms and functional implications of ncRNAs will enhance our understanding of the coordinated regulation of plant defences against various biotic and abiotic stresses and for key growth and development processes to better design various legume crops for global food security.

1. Introduction

Legumes are the third largest family of flowering plants, and grain legumes are essential components of the human food diet, supplying ‘plant-based dietary proteins’ and essential micronutrients and vitamins [1,2,3]. Thus, legume crops serve as an essential component for sustaining global food security. Their ability to fix atmospheric nitrogen through symbiotically active bacteria in root nodules enriches soil nitrogen content and minimizes the use of chemical-based nitrogenous fertilizers, thus protecting the environment from pollution [1]. In the past, elucidating the function of protein-coding genes controlling biotic and abiotic stresses and developmental processes in plants has involved conventional breeding and biochemical and molecular approaches [4]. However, rapid progress in functional genomics, especially transcriptome sequencing by RNA-seq, has given us the opportunity to investigate RNAs that do not code proteins, known as ncRNAs, which control diverse biological functions in the plant kingdom [5]. These ncRNAs are classified as small ncRNAs, comprising miRNAs (21–24 nt long) [6], small interfering RNAs (siRNAs) [7], Piwi-interacting RNAs (piRNAs) (generally found in animals) [8] and lncRNAs (>200 nt long) [9]. circRNA are another class of ncRNA generated from pre-mRNA splicing, featuring closed 3′ and 5′ ends covalently [10]. In addition to these ncRNAs, small nucleolar RNAs (snoRNAs), ribosomal RNAs (rRNAs), and transfer RNAs (tRNAs) known as housekeeping ncRNAs are also found in plant species [11]. The main classes of ncRNAs, illustrated in Figure 1, contribute to various plant development pathways and abiotic and biotic stresses by modulating the expression of associated genes [12,13,14,15,16,17,18]. In this review, we discuss the biogenesis of major plant ncRNAs and their interplay with corresponding target gene(s) controlling plant responses to biotic and abiotic stresses and with key developmental processes, including flowering, pod and seed development, nodulation, and nutrient acquisition in various legume crop species.

2. Types, Origin, and Function of Major Regulatory ncRNAs

Plant ncRNAs are ubiquitous and versatile repressors [6]. The major ncRNAs found in plants are broadly classified as small ncRNAs comprising miRNAs, siRNAs, long noncoding RNAs (lncRNAs), and circular noncoding RNAs (circRNAs) [6,7,19,20,21]. miRNAs are endogenous ncRNAs, about 20–24 nt long, abundant in both animal and plant kingdoms. They originated from miRNA genes through the transcription process by RNA pol II followed by processing of primary transcripts into mature miRNA catalysed by DICER-like (DCL) proteins [6,22,23]. Eventually, the mature miRNA is incorporated into the ARGONAUTE protein to assemble a miRNA-induced silencing complex (miRISC) [24]. Primarily, miRNAs function at the post-transcription level by base pairing with cognate mRNA, degrading or inhibiting mRNA translation [7,25,26]. Likewise, siRNAs (~22 nt long) originated from DCL-catalyzed processing of double-stranded RNA (dsRNA) precursors [7,24]. Primarily, siRNAs are classified as (1) trans-acting siRNAs (ta-siRNAs) generated from long noncoding single stranded RNAs, (2) natural antisense transcript-derived siRNAs (nat-siRNAs) derived from natural antisense RNAs, and (3) siRNAs belonging to repetitive DNA or transposons (see [27]). They play a central role in DNA methylation, chromatin modification, and repression of distinct mRNA targets by trans-acting siRNAs [28]. lncRNAs are > 200 nt noncoding RNAs found in animals and plants, located in the cytoplasm and nucleus [29,30]. The major classes of lncRNAs are long intergenic RNAs (lincRNAs), natural antitranscripts (NATs), and intronic ncRNAs (incRNAs) [31,32]. They are transcribed by RNA polymerase II or III and polymerase IV/V [33]. These lncRNAs can serve as precursors of miRNAs and siRNAs and act as endogenous target mimics (eTM) competing for various miRNAs [34]. Moreover, they participate in chromatin topology modification [35], alternative spicing [36], post-translational regulation [37], and protein relocalization [38]. Further detail on plant-lncRNA function is available elsewhere [39,40,41]. Circular RNA is a covalently closed, single-stranded RNA molecule generated by back-splicing events, categorized into exonic circRNAs, intronic circRNAs, intergenic circRNAs, and UTR circRNAs [10]. Our understanding on the role of circRNAs in plants is still limited [42].

3. Evolution, Conservation, Species Specificity, Tissue Specificity, and Genotype- and Stress-Dependent Expression of ncRNAs

Among the various plant ncRNAs, miRNAs are evolutionarily highly conserved in plant species ranging from nonvascular mosses to flowering monocots and dicots [43,44]. Researchers have found that individual plant species harbor conserved miRNAs and species-specific miRNAs [45]. Various conserved miRNAs have been reported, viz., miR156, miR159, miR165, and miR169 [44]. Likewise, species-specific miRNAs, viz., miR4414, miR5037, miR5208, miR5287, and miR5559, have been reported in Astragalus chrysochlorus [46] and Ammopiptanthus mongolicus [47] legume species and may be specifically expressed in legumes. De la Rosa (2020) [45] found that genes for miR398 are distributed throughout spermatophytes, but miR2119 was only found in legume species, indicating its recent emergence. The function of miR2119 in Phaseolus vulgaris and its presence in other legumes such as Glycine max, Medicago truncatula, and Arachis hypogaea have been reported [43,48]. Conserved miRNAs are involved in regulating common plant developmental processes, e.g., plant morphology; however, species-specific miRNAs may regulate special trait development, e.g., legume-specific cell processes and nodulation in legumes [49,50]. Expression patterns of conserved miRNAs vary greatly across plant species [51]. This has been supported by various research groups [52,53,54] by observing the abundance of miR398 expressed in the leaves but not the inflorescence of Arabidopsis. Conversely, M. truncatula had a high abundance of miR398 expressed in flowers but not in leaves [51]. Moreover, the expression of miRNAs varies from tissue to tissue, genotype to genotype, and stress to stress [44]. Under drought stress, Barrera-Figueroa et al. (2011) [55] noted 20 miRNAs differentially expressed in IT93K503-1 (drought-tolerant) and CB46 (drought-sensitive) cowpea genotypes. Among these, nine were only expressed in one genotype and not the other. Likewise, 11 miRNAs were expressed in one cowpea genotype but not in other genotypes under water stress, indicating genotype-dependent expression of miRNAs [55].
In groundnut, the leaves, flowers, and roots had higher expression of miR3 and miR7 than the seeds, and the stems’ leaves, roots, and stems had higher expression of miR156 than the flowers and seeds, suggesting tissue-specific expression of miRNAs in legumes [56]. Similarly, for lncRNAs, Das et al. (2019) [57] noted a higher expression of Cc_lncRNA-765 and its target mRNA, a carboxy peptidase-like mRNA, in seed tissue than pod tissue in pigeon pea. The reverse was true for Cc_lncRNA-2150 in pods compared with seeds at 30 days after podding. Similarly, Tridade et al. (2010) [58] reported upregulatory activity of miR398 and miR408 in response to drought stress in M. truncatula, but others reported downregulatory activity of miR398a, miR398b, and miR408 under salinity and alkalinity stress [59]. In the same way, miR399 was upregulated under phosphorus deficiency and downregulated under nitrogen deficiency in common bean [60]. Golicz et al. (2018) [61] witnessed sequence homology of four lncRNAs in various legume species, including soybean, chickpea, and M. truncatula. Several plant and legume plant-based ncRNA databases, viz., SoyKB [62], PNRD [63], PLNlncRbase [64], GreeNC [65], and PLncPRO [66], have been developed to discover and functionally annotate ncRNAs. The continual evolution of ncRNA databases and advances in computational and comparative analysis will improve our understanding of the conservation of ncRNA genes with their precise mode of function across various species in the plant kingdom [41].

4. ncRNAs Mediating Plant Immunity against Attacking Pathogens

Among the various biotic stresses, infections caused by fungi, bacteria, viruses, and nematodes significantly damage plants, resulting in substantial yield losses in various legumes [67,68,69]. Plants evoke a two-layer defence mechanism known as pathogen-associated molecular patterns (PAMPs)-triggered immunity (PTI) and effector-triggered immunity (ETI) against evading pathogens [70,71,72]. A series of protein-encoding gene(s), viz., pathogenesis-related genes, R genes, and other defense-related genes, are switched on and mediate conferring ETI and PTI in response to pathogen attack [72]. However, the emerging RNA-seq-based transcriptome sequencing approach underpinned a plethora of ncRNAs modulating various pathogenesis-related genes and R genes, thus regulating the plant immune response to various attacking pathogens [73]. ncRNAs play vital role in protecting plants from pathogen invasion by modulating ROS, the MAPK signalling cascade, and various TFs involved in switching on defence gene(s) [67,69]. Likewise, these, ncRNAs also participate in turning on downstream R genes and genes encoding pathogenesis-related proteins/phenolic compounds/phytoalexins and various phytohormone signalling in response to pathogen attack, thereby regulating plant disease resistance [67,69,74].
To establish the role of miRNAs regulating Ascochyta blight (AB) resistance in chickpea, Garg et al. (2019) [69] unveiled 651 miRNAs, including 173 novel miRNAs, in response to AB infection in contrasting parents. The authors noted both upregulation and downregulation of various miRNAs identified at various time points of AB infection. Functional analysis suggested the role of these miRNAs regulating AB resistance by evoking various TFs, phytohormones, and pathogenesis-related protein and R genes. Of the 12 miRNAs, 5 miRNAs, such as miR482b-3p, miR167c, miR171b, miR157a, and miR5232, were validated through degradome sequencing [69] (see Table 1). The predicted target genes of the above corresponding miRNAs were identified as Ca_08122 (encoding CC-NBS-LRR), Ca_19433 encoding (Dof zinc finger protein), Ca_00359 (encoding ERF), Ca_15107 (encoding senescence-associated protein), and Ca_12185 (encoding calcium-transporting ATPase). The study also explained the possible causative mechanism of AB infection in the susceptible genotype through the upregulation of miR482b-3p, miR159k-3p, nov_miR66, and miR171 miRNAs and the downregulation of the corresponding target genes encoding NBS-LRR, PR protein, a serine-threonine kinase, and PPR proteins, allowing AB infection [69] (see Table 2). Considering fusarium wilt (FW) in chickpea, Kohli et al. (2014) [68] reported 122 conserved and 59 novel miRNAs by sequencing small RNA from ICC4958, a FW-tolerant chickpea genotype. The authors noted the upregulation of FW-responsive miRNAs, viz., miR530 (targeting zinc knuckle proteins) and the microtubule-associated proteins miR156_1 miR156_10, car-miR2118, and car-miR5213 (targeting TIR-NBS-LRR). Deep sequencing of two soybean cultivars, Hairbin xiaoheidou (resistant to soybean cyst nematode) and Liaodou 10 (susceptible to soybean cyst nematode), unearthed 364 and 21 novel miRNAs [74]. Among the conserved miRNAs identified, MiR169 was upregulated in Liaodou 10 and downregulated in Hairbin xiaoheidou; however, MiR319 (targeting TCP gene) was upregulated in both cultivars.
Likewise, gma-miR390b was upregulated by soybean cyst nematode (SCN) in Hairbin xiaoheidou and downregulated in Liaodou 10. Of the 21 novel miRNAs identified, soy_1, soy_2, and soy_3 (targeting HD-ZIP transcript factor) and soy_9 (targeting calmodulin) were noted [74]. Likewise, 60 SCN-responsive miRNAs were identified in KS4607 (susceptible) and KS4313N (resistant) soybean genotypes using deep sequencing and miRDeep2 pipeline analysis [74]. Among the SCN-responsive miRNAs, various conserved miRNAs, viz., miR171, mir399, miR159, and miR398, and legume-specific miRNAs, viz., miR9750, miR2119, and miR1512, were identified. Of the DE miRNAs, 34 miRNAs were upregulated; notably, miR159b-3p, miR159f-3p, and miR972 were downregulated in the susceptible cultivar, while 14 miRNAs were upregulated and miR2119, miR398a, and miR398b were downregulated in the resistant cultivar [67]. In groundnut, small RNA transcriptome sequencing of pod rot infected groundnut using Illumina HiSeq 2000 elucidated 334 miRNAs, of which 97 were downregulated and 27 were upregulated [104]. Functional validation of selected miRNA, viz., ahy-miR396e-5, was downregulated, but its target gene, c39419_g1_i1, was upregulated after infection. Likewise, ahy-miR3509-5p, ahy-miR166f, and ahy-miR159b were downregulated after infection, but their corresponding target genes, c40055_g1_i3, c31393_g1_i1, and c41016_g4_i1, were upregulated [104]. However, a complete understanding of ncRNAs identified as regulating disease resistance in legumes remains elusive. Future identification of novel disease-responsive ncRNAs will provide novel insights into the interplay of ncRNAs and the plant immune response for developing disease-resistant legumes.

5. Deciphering the Molecular Mechanisms of ncRNAs Regulating the Response of Legumes to Water Stress

Drought stress is the most important abiotic stress globally, affecting all plant growth and developmental stages, and ultimately reducing crop yields [145]. Plants adapt to a water deficit environment by evoking various physiological, biochemical, metabolic, and molecular mechanisms [146]. Many QTL/genes contributing to drought tolerance have been investigated in various legumes [147]. Indeed, the participatory role of various regulatory ncRNAs and their corresponding target gene(s) controlling drought stress have been deciphered in various plant species, including legumes [55,95,113,125]. A plethora of novel drought-responsive miRNAs have been identified in legume crops—157 in cowpea [55], 143 and 128 in grass pea [121], and 284 in chickpea [92]—and 3457 high-confidence lncRNAs have been identified in chickpea [66]. ncRNAs confer drought tolerance by regulating gene(s) encoding various regulatory TFs and osmoregulatory/osmoprotective compounds by activating hormone signalling and antioxidants that minimize oxidative stress/reactive oxygen species (ROS) activity in plants under water stress [55,113,121].
Deep sequencing of two contrasting cowpea genotypes —CB46 (drought-sensitive) and IT93K503-1 (drought-tolerant)—grown under normal and drought stress conditions enabled in identifying 44 drought-responsive miRNAs (30 upregulated and 14 downregulated) [55] (see Table 1). Notably, miR156 (targeting SPB transcription factors) was upregulated and miR169 (targeting NFYA5) was downregulated in both genotypes under water stress. miR160a and miR160b (targeting Auxin Response Factors) and vun_cand015 (targeting bHLH transcription factor) were upregulated in the tolerant cultivar, and miR2111 (targeting Kelch repeat-containing F-box proteins) was upregulated in the drought-sensitive cultivar [55].
To predict the possible role of miRNAs in producing osmoprotective compounds to regulate the drought stress response, Shui et al. (2013) [113] elucidated and validated the active role of vun-miR5021, vun-miR156b-3p and vun-miR5021 (targeting CPRD86), vun-miR156b (targeting 1-pyrroline-5-carboxylate synthase P5CS involved in proline synthesis), and vun-miR156f (targeting multicystatin gene encoding cystatins) miRNAs in leaf and root tissue of two contrasting cowpea genotypes (Danlla and Tvu7778) under water stress.
A study on the participatory role of conserved miRNAs—miR398a/b and miR408—in regulating water stress in pea revealed the downregulation of these miRNAs in root and shoot tissue under water deficit conditions [125]. However, the copper superoxide dismutase, CSD1 target gene of miR398a/b, was upregulated, suggesting an inverse relationship between the target gene and the involved miRNA controlling water stress in pea. Similarly, De la Rosa et al. (2019) [48] supported the upregulatory role of CSD1 and ADH1 mRNAs targeted by miR398 and miR2119 in common bean adapting to drought stress.
Grasspea sequencing of small RNA identified numerous drought-responsive miRNAs [121]. Among the differentially expressed miRNAs, lsa-miR-169b, lsa-miR-319, lsa-miR-398, lsa-miR786, lsa-miR1361, and lsa-mir-156 were upregulated, and lsa-miR897, lsa-miR969, lsa-miR186, and lsa-mir-1520b were downregulated. lsa-miR-319 and lsa-miR-398 were predicted to target the TCP gene and cytosolic CSOD1 and chloroplastic CSOD2 genes, respectively [121]. In chickpea, small RNA sequencing of root tissues under water stress identified 284 miRNAs [95]. Functional validation of selected miRNAs, including miR397, miR398, miR164, and miR399 targeting LACCASE4, COPPER SUPEROXIDE DISMUTASE, NAC1, and the PHO2/UBC24 gene, respectively, showed an inverse relationship under drought stress [95]. Illustrating the role of abiotic stress responsive lncRNA, Singh et al. (2017) [66] identified a total of 3457 high-confidence lncRNAs responding to drought and salinity stress in chickpea. The drought sensitive genotype ICC1882 showed the least number of 126 differentially expressed lncRNAs at the early reproductive stage, while a large number of lncRNAs exhibited downregulation under drought stress in all the tested samples. In parallel, a large number of lncRNA showed differential expression at the early reproductive stage in the ICC4958 drought tolerant chickpea genotype [66].
Considering the role of circRNAs attributing drought tolerance, Dasmandal et al. (2020) [148] uncovered numerous drought responsive differentially expressed circRNAs in chickpea and soybean. The authors also predicted three eTMs those acted as sponge for miRNAs that target Glyma.18G065200.1 gene in soybean, and XM_004517122, XM_027336693 genes in chickpea. The functional role of these targeted genes was associated with hormone signalling and various transcription factors under drought stress [148]. Further mechanistic understanding of ncRNAs and the corresponding target gene(s) will enhance our understanding of ncRNAs regulating drought tolerance in legume crops.

6. Role of ncRNAs in Plant Adaptation to Salinity Stress

The rapid conversion of uncultivable land to cultivated land and the excessive use of irrigation water have increased salinity, which is a major challenge for crop growth, including legumes, and causes significant yield losses [149]. Plants orchestrate various biochemical and molecular mechanisms to survive the increasing salinity stress [149], including ncRNAs [15,82,83,102], which target genes related to photosynthesis, TFs regulating growth, genes related to salinity-responsive hormone signalling, genes that minimize the uptake of toxic ions, viz., Na+, and genes that limit ROS activity [15,83,95].
Paul et al. (2011) [114] investigated the role of miRNAs controlling salinity stress in cowpea and recovered 18 conserved miRNAs (e.g., miR160, miR156/157, miR159, miR169, miR172, miR408) from root tissue and identified 15 corresponding target gene(s) as TFs (e.g., ARF, SBP, AP2, TCP). Functional validation through quantitative real-time PCR (qRT-PCR) revealed the upregulation of seven miRNAs under salinity stress.
Transcriptome analysis of root apex treated with salinity stress using miRDeep2 identified 66 salt-responsive miRNAs in soybean, of which 14 were upregulated (notably, miR172f and miR390e) and 22 were downregulated (notably, miR399a/b, miR1512b, miR156g, and miR156j) under salinity stress [80]. The predicted putative target genes of miR399a/b were Glyma.14G188000, Glyma.15G074200, Glyma.08G359400, and Glyma18G177400 (encoding multicopper oxidases) and Glyma03G021900 (encoding a growth-regulating factor). Likewise, Glyma.02G281100 and Glyma.14G033500 encoding LRR receptor-like kinases were the target genes of miR390e [83]. Subsequently, strand-specific transcriptome sequencing identified 3030 lincRNAs and 275 lncNATs in soybean roots under salinity stress [82]. Importantly, 75% of these lncRNAs were upregulated under salinity stress. Genome-wide scanning of salinity-responsive miRNAs elucidated 876 miRNAs related to salinity and alkalinity stress in M. truncatula [59]. Thirty-five miRNAs (including mtr-miR156 family, mtr-miR159a, and mtr-miR171) were upregulated under salinity and alkalinity stress, and eight miRNAs (including mtr-miR171e-3p, mtr-miR2628, mtr-miR398a-3p, mtr-miR398a-5p, and four novel miRNAs) were downregulated under both stresses [59]. Functional validation of miR319 (targeting MTR_3g011610, MTR_1g102550, and MTR_1g052470) and miR408 (targeting BBLP and MTR_8g089110) indicated their participatory role in salinity and alkalinity stress tolerance [59]. In chickpea, small RNA sequencing of root tissues treated with salinity stress identified 284 miRNAs [95]. Inverse correlation patterns of miRNA397, Car-novmiR2, and Car-miR5507 targeting the LACCASE4, HAK5, and CIPK23 genes, respectively, were observed at the transcript level regulating salinity stress tolerance in chickpea [95].
A genome-wide survey of lncRNA through transcriptome analysis in groundnut identified 1442 lncRNAs [102]; notably, TCONS_ 00292946 lncRNA was downregulated in roots within 12 h of salinity stress but upregulated at 24 h. Expression of TCONS_00176941 was upregulated within 12 h in roots and downregulated within 12 h of salinity stress in leaves, while TCONS_00011551 was upregulated under salinity stress [102]. Wang et al. (2015) [15] investigated the role of lncRNAs involved in regulating the salinity stress response and conferring tolerance by alleviating ROS-related stress in Medicago truncatula. The authors identified the functional role of various lncRNAs attributing to salinity tolerance, including TCONS_00116877, which induced the Medtr7g094600 gene encoding glutathione peroxidase to minimize ROS-derived stress in roots (see Table 2).
Alzahrani et al. (2019) [116] uncovered 1220 salt-responsive miRNAs by small RNA sequencing of two contrasting faba bean (Vicia faba) genotypes for salinity stress response (ILB4347 tolerant and Hassawi-3 sensitive). The Hassawi-3 genotype had 284 upregulated and 243 downregulated miRNAs, while ILB4347 had 298 upregulated and 395 downregulated miRNAs in the control and under salinity stress. The target gene(s) were predicted to encode TFs, laccases, superoxide dismutase, plantacyanin, and F-box proteins in addition to genes involved in hormone signal transduction, phosphatidylinositol signalling, and the MAPK signalling pathway [116].

7. Contribution of ncRNAs Attributing Plant Adaptation under Metal Toxicity Stress

Metal toxicity is an abiotic stress increasingly faced by plants due to rapid industrialization, excessive use of inorganic fertilizers, and overuse of irrigation water contaminated with heavy metals, especially cadmium and mercury [150]. Among the various complex molecular mechanisms, identifying the role of ncRNAs, including miRNAs and lncRNAs, is a potential approach for minimizing metal toxicity in plants [77,151].
Deep sequencing and high-throughput degradome analysis of heavy metal mercury-treated and mercury-free M. truncatula seedlings identified 201 miRNAs [77]. Of these, 12 were specifically induced under mercury stress. Functional analysis of miR2681, miR2708, and miR2687 targeting the TIR-NBS-LRR (encoding disease resistance protein), TC114805 (encoding salinity tolerance protein), and XTH gene coding xyloglucan endotransglucosylase/hydrolase contributing to cell wall development, respectively, was deciphered (see Figure 2). Thus, these miRNAs and the putative target could be an important approach for regulating heavy metal stress tolerance in M. truncatula [77]. Earlier, Zhou et al. (2008) [152] reported the upregulatory role of miR171, miR319, miR393, and miR519 and the downregulatory role of miR166 and miR398 in response to Al3+ treatment in M. truncatula. Subsequently, Chen et al. (2012) [78] elucidated 326 known miRNAs and 21 new miRNAs responsive to aluminium toxicity using small RNA sequencing of Al3+-treated and Al3+-untreated M. truncatula. Functional characterization of selected miRNAs, viz., pmiR-003 and pmiR-008 (targeting genes encoding TIR-NBS-LRR resistance protein), revealed their possible role in mediating aluminium toxicity tolerance [78]. Twenty-eight miRNAs responsive to aluminium toxicity were recovered from roots and nodules in common bean using the miRNA-macroarray hybridization technique [99]. Functional validation of selected miRNAs revealed upregulation of miR164 targeting (NAC1 TF), miR170 targeting (SCL TF), and miR393 targeting TIR1, and downregulation of miR157 targeting (SPL) and miR398 targeting (CSD1) under aluminium stress in common bean nodules [99]. Eleven miRNAs, viz., miR157, miR156, miR170, miR172, and miR319, exhibiting strong upregulation in root nodules, and 11 miRNAs, viz., miR160, miR397, miR399, miR408, pvu-miR1509, and pvu-miR1514a, exhibiting strong downregulation in leaves or roots, were discovered under manganese toxicity in common bean [60] (see Table 2). Few toxic metal-responsive miRNAs have been reported in legumes. Therefore, further study is needed to gain insight into toxic metal-responsive miRNAs and their target genes and precise function.

8. Molecular Mechanisms of ncRNAs Regulating Nutrient Acquisition and Homeostasis in Legumes

Plants acquire essential nutrients by recruiting various physiological and molecular mechanisms via roots and soil for proper growth and development [153,154]. Of these mechanisms, the critical role of ncRNAs regulating the uptake of various macro- and micronutrients has been recognized [155,156].
Nitrogen (N)—serving as the source of various essential amino acids and acting as an important element for entire nitrogen metabolism—is a critical determinant for plant growth and development [157]. Emerging functional genomics approaches, viz., RNA-seq, can underpin the plethora of nitrate transporter QTLs, gene(s), and ncRNAs controlling N use efficiency (NUE) in plants [158]. However, the complete molecular mechanism of NUE/N homeostasis remains unclear in plants, including legumes.
Evidence for the miRNAs controlling the nitrogen response in plants has been reported [86,159]. The upregulation of pri-miR156 and pri-miR447c and downregulation of pri-miR169 and pri-miR398a were reported in Arabidopsis under nitrogen-limited conditions [160]. Several nitrogen-responsive miRNAs, viz., miR164, miR169, miR172, and miR397 in maize shoots and miR160, miR167, miR168, and miR169 in maize roots, under nitrogen deficiency conditions have been reported [159]. Likewise, several nitrogen-responsive miRNAs have been reported in legume crops [86]. Wang et al. (2013) [86] recovered 168 nitrogen-responsive miRNAs from small RNA sequencing of a low N tolerant (No.116 genotype) and low N sensitive (No.84-70 genotype) soybean genotype. The study revealed downregulation of gma-miR2606a/b-3p in the roots of variety No.116 and upregulation of gma-miR1512a-5p in the roots of variety No.84-70 under short-term low N. However, gma-miR396b/c/d/f/g-5p was downregulated in the shoots of No.116 and upregulated in the shoots of No.84-70 under short-term low N stress [86]. Moreover, some of the predicted miRNA targeting genes were predicted to play a role in protein degradation, viz., gma-miR156b/6f-5p (targeting Glyma07g31580) and gma-miR396bg-5p (targeting Glyma05g20930 and Glyma06g18790), encoding E3 ubiquitin ligase and Cathepsin L1 (see Table 2).
Phosphorus (P) is the second most essential macronutrient required for basic biochemical and metabolic processes in plants, including legumes [161]. Plants usually uptake P in the form of inorganic phosphate (Pi). Thus, P deficiency limits overall plant growth and development. The involvement of several P-responsive ncRNAs has been elucidated in various plant species [14,17,160,162,163,164]. Likewise, previously P-responsive miRNAs have been reported in common bean [60,144], white lupin [119], soybean [165], M. truncatula [132], alfalfa [166,167], and lupin (Lupinus albus) [119]. Several conserved regulatory miRNAs, such as miR399 [162,168,169,170] and miR156, miR169, and miR2111 [160] regulating Pi homeostasis have been reported in Arabidopsis. Li et al. (2018) [13] confirmed the inductive role of miR399 (targeting phosphate transporter genes) and miR398 (targeting Copper chaperone for SOD) under low Pi stress in roots of Medicago sativa. However, the authors noted downregulation of miR156 (targeting SPL TF), miR159 (targeting MYB TF), miR160 (targeting auxin response factor TF), miR171 (targeting GRAS TF), and miR2643 (targeting MATE). The molecular mechanism involving IPS1 lncRNA serving as eTM for miRNA399 targeting PHO2 gene expression and controlling Pi homeostasis has been established in Arabidopsis [34]. Under low Pi conditions, upregulation of the PHR1 gene and miRNA399 inhibiting the PHO2 gene (encoding transcript causing Pi transporter degradation) enabled high Pi acquisition in Medicago sativa [17]. Downregulation of PDIL2 and PDIL3 lncRNAs enhanced transcript expression of Medtr1g074930, a Pi transporter gene, enabling high Pi uptake under low Pi conditions. However, PDIL1 lncRNA serves as a target mimicry for miR399, inhibiting the degradation of MtPHO2 transcripts that could downregulate the Pi transport gene and Pi uptake [17] (see Figure 2). To gain insight into the role of P-responsive circRNAs, Lv et al. (2020) [14] uncovered 120 differentially expressed cicrRNAs by transcriptome sequencing of two contrasting P-responsive soybean genotypes at different P levels. Gene ontology (GO) enrichment analysis predicted that the putative role of the differentially expressed circRNAs is related to nucleoside binding, organic substance catabolic processes, and oxidoreductase activity [14]. Low P-responsive circRNAs could be targeted for improving phosphorus use efficiency in soybean. Thus, a complex network of ncRNAs and their corresponding target gene(s) play a central role in regulating Pi homeostasis in plants.

9. Regulatory Role of ncRNAs for Shaping Developmental Processes in Legume Species

Apart from various biotic and abiotic stresses, ncRNAs, including miRNAs (conserved and nonconserved) and lncRNAs, play a pivotal role in regulating plant growth and development and in various metabolic pathways, which has been investigated in various legume species [61,91,92,96,103,120,171,172]. Small, deep RNA sequencing analysis of seven chickpea tissues was used to investigate a comprehensive set of 440 known and conserved and 178 novel miRNAs targeting various TFs and gene(s) that control various developmental processes, including leaf, flower, pod, and root development and various metabolic processes in chickpea [92] (see Table 1). Subsequently, small RNA sequencing of chickpea leaves and flowers discovered 157 conserved and novel miRNAs that regulate various developmental processes and stress responses [12]. Of the identified miRNAs, miR156, miR159, miR160, miR162, miR164, miR172, miR408, and miR393 targeting SBP, MYB, ARF, DCL1, HD-zip, AP2, F-box protein, and plantacyanin encoding genes, respectively, contribute to various plant development processes [12] (see Table 2). The authors also disclosed the role of TAS3-derived tasiRNA targeting ARF2, ARF3, and ARF4 transcription factors controlling auxin response, and thus contributing to development pathways in chickpea. In this context, Jagadeeswaran et al. (2009) [51] identified and characterized Tas3-siRNAs from M.trucatula and also functionally validated three ARF genes targeted by these Tas3-siRNAs.
Considering ta-siRNA participating in regulating compound leaf and flower development in L. japonicus, Yan et al. (2010) [173] established the role of Reduced leaflet1 (REL1) and Reduced leaflet3 (REL3) genes encoding homologs of Arabidopsis (Arabidopsis thaliana) ‘SUPPRESSOR OF GENE SILENCING3′ and ‘ARGONAUTE7/ZIPPY’, respectively, key components required for ta-siRNA biogenesis. Positional cloning analysis of REL1 and REL3 genes revealed that the ta-siRNA pathway critically plays significant role in controlling compound leaf and flower development in L. japonicus [173]. Likewise, elucidating the role of trans-acting siRNA3 (TAS3) involved in leaf margin indentation and organ separation, Zhou et al. (2013) [174] examined that Mt-AGO7/LOBED LEAFLET1 is required for the biogenesis of ta-siRNA to suppress the expression of Auxin Response Factors. Evidence of lobed leaf margin and widely spaced lateral organ phenotype demonstrated in the ago7 mutant suggested that TAS3 plays a negative role in leaf margin and lateral organ development in M. truncatula [174]. Examining the functional role of lncRNA associated with flower development, Khemka et al. (2016) [91] discovered a total of 2248 long intergenic noncoding RNA obtained from the results of RNA-seq data of eight flower development tissues. Further, qRT-PCR result showed specific expression of Ca_linc_0051 and Ca_linc_0139 lncRNA in the flower bud and shoot apical meristem stage, confirming their possible role in flower development in chickpea [91].
Glazińska et al. (2019) [120] reported several miRNAs regulating floral development, viz., Ll-miR280, Ll-miR281, and Ll-miR285 (possibly targeting ARF6 and ARF8); Ll-miR445 and Ll-miR130 (targeting TCP4 and MYB33); and Ll-miR329/miR160-5p, Ll-miR332/miR160-5p, and Ll-miR333/miR160-5p miRNAs regulating flower abscission in yellow lupin (Lupinus luteus L.). Among the siRNAs identified from this study, Ll-siR173, Ll-siR4, and Ll-siR13 exhibited upregulation and downregulation of Ll-siR208, suggesting the active role of siRNA functioning in lupin pedicel [120]. Das et al. (2019) [57] explored a plethora of lncRNAs and target miRNAs forming an endogenous target mimicry leading to pod and seed development using transcriptome analysis of tissue collected during anthesis and pod development in pigeon pea. Functional validation revealed that sequestering Cc-miR160h by Cc_lncRNA-2830 enabled the transcription of XM_020377020 (encoding auxin response factor 18-like protein) during pod development at 10 and 20 days after anthesis (DAS). However, expression of Cc_lncRNA-2830 at 30 DAS decreased, which upregulated Cc-miR160h and degraded the XM_020377020 transcript [57] (see Figure 2).
To better understand the role of miRNAs regulating embryo and pod development in groundnut, small RNA profiling and degradome sequencing identified 70 known and 24 novel miRNA families [105]. Functional validation of selected miRNA, viz., miR164, miR167, miR172, miR390, miR7502, and miR9666, using qRT-PCR revealed upregulatory activity; however, miR156, miR396, miR894, miR1088, miR4414, and miRn8 were significantly downregulated during early embryo and pod development [105]. In groundnut [Chen et al. (2019) [109], 29 known and 132 novel miRNAs were identified when exploring the participatory role of miRNAs in embryo development under calcium deficiency. Transcriptome analysis identified 52 differentially expressed genes targeted by 20 miRNAs. Functional validation of selected miRNAs, viz., ahy_novel_miRn129 and ahy_novel_miRn130 (targeting transcription factor “LONE- SOME HIGHWAY” (LHW) encoding bHLH transcription factor), exhibited upregulation under calcium deficiency [109]. The same study showed upregulation of ahy_novel_miRn112 and downregulation of target gene (NAM/CUC), while ahy_novel_miRn23 (targeting CYP707A1 and CYP707A3 encoding ABA 8′-hydroxylase) was significantly upregulated, and ahy_novel_ miRn30, ahy_novel_miRn29, and ahy_novel_miRn38 with their corresponding targets TEOSINTE BRANCHED1, CYCLOIDEA, and PROLIFERATING CELL FACTORS 4 (TCP4) involved in jasmonic acid biosynthesis were downregulated [109]. Thus, these miRNAs with their target gene(s) modulate embryo development in groundnut.
As the entire underlying molecular mechanism for seed development, from embryogenesis and filling to maturation, remains elusive [98], several investigations have reported the involvement of various ncRNAs regulating seed development in grain legumes [92,93,98,102]. To investigate the contributory role of ncRNA involved in the seed development process, transcriptome sequencing of seed samples using an Illumina Genome Analyzer IIx uncovered 72 known and 39 new miRNAs involved in seed development, particularly embryogenesis, dormancy, and maturation, in common bean [98]. The notable miRNAs and the target genes involved in regulating seed development were MIR156 repressing SPL; MIR169 repressing NF-YA1 and NF-YA9; MIR399 inhibiting SUT1 related to sucrose transport; MIR399 inhibiting PHO2 contributing in phosphorus allocation; MIR160 repressing ARF10, ARF16, and ARF17; MIR167 inhibiting NCED1 associated with ABA synthesis; and MIR395 repressing SULTR2;1, APS contributing to sulphate assimilation and allocation during seed filling [98]. Likewise, genome-wide profiling of miRNAs using small RNA sequencing of seeds of two contrasting chickpea genotypes—Himchana1 (low seed weight) and JGK3 (high seed weight)—unfolded 113 known and 243 novel miRNAs controlling seed development in chickpea [93] (see Table 1). The target genes of identified miRNAs contributing to seed development were predicted to be SPL, GRF, MYB, ARF, HAIKU1, SHB1, KLUH/CYP78A5, and E2Fb. Low expression of Car-miR319 and Car-miR166 and upregulation of their corresponding target genes, bZIP and homeobox-REVOLUTA TFs, in JGK3 indicated their important role in seed size determination in chickpea [93]. The authors also located 19 miRNAs and 41 target genes in previously identified QTLs contributing to seed size.
The role of various conserved miRNAs, viz., miR167, miR390, miR164, miR399, miR156/157, miR1511, and mir319, and seven novel miRNAs, viz., NovmiR13, NovmiR12, and NovmiR04, regulating seed development in narrow-leafed lupin was confirmed in studies by DeBoer et al. (2019) [118]. Differential expression analysis revealed upregulation of Lan-miR-156a-2, Lan-miR-164-3, Lan-miR-167a/c, Lan-miR-319, Lan-miR-399b/c, NovmiR12, and Nov-miR13 in seeds, indicating their role in regulating seed development in lupin [118]. The role of miRNAs controlling genes related to sugar metabolism during seed development is worth mentioning [87,175]. In soybean, deep sequencing and degradome sequencing of developing soybean seed revealed several miRNAs targeting genes that contribute to seed development [87]. Among the identified miRNAs, functional validation of gma-miR1530 revealed its role in inhibiting the target transketolase gene that contributes to switching carbon assimilation to energy metabolism during seed development. Likewise, the pentatricopeptide repeat protein-encoding gene was targeted by Soy_3 and Soy_16, while Soy_25 (targeting Glyma05g33260 homolog of Arabidopsis “SUPPRESSOR OF GENE SILENCING 3”) contributing to seed development was identified [87] (see Figure 2). A total of 484 miRNAs were recovered from small RNA sequencing of four contrasting soybean lines with high protein/high oil, high protein/low oil, high oil/low protein, and low protein/low oil [175]. Functional validation of selected miRNAs, including Glyma.13G035200 and Glyma.14G156400 (encoding alcohol dehydrogenase 1) targeted by Gma-miR2119, Glyma.04G178400 (encoding ADP-glucose pyrophosphorylase family protein) targeted by Gma-miR1521a, and Glyma.19G094000 (related to sugar synthesis and metabolism) targeted by miR156, using RT-qPCR indicated their significant role in controlling storage genes during seed development in soybean [175].
Computational analysis identified 347 candidate circRNAs in groundnut [110]; the differential expression of 29 circRNAs was upregulated in seeds collected from RIL 8107′ at 35 days after flowering (DAF) and RIL 8106′ at 35 DAF, confirming their contributory role in seed development [110]. Likewise, Ma et al. (2020) [111] detected 9388 known and 4037 novel lncRNAs in groundnut, of which 1437 lncRNAs were differentially expressed. Functional validation of selected lncRNAs confirmed their role in seed development. The participatory role of miR156, miR159, miR171, and miR14 (targeting genes related to cellular amino acid metabolism, fatty acid metabolism, and lipid metabolism) in groundnut is noteworthy [56].
To establish the role of the DCL2-dependent 22-nucleotide siRNA (derived from long inverted repeats) regulating chalcone synthase (CHS) genes attributing seed coat colour in soybean, a study conducted by Jia et al. (2020) [176] revealed that CRISPR/Cas9-driven loss-of-function mutants of DCL2 (GmDCL2a and GmDCL2b) caused changes in seed coat colour from yellow to brown in Gmdcl2a/2b mutants in soybean. Thus, this study confirmed that DCL2 controls soybean seed coat colour via generating siRNA from long inverted repeats [176].
Further identification of ncRNAs related to the development process, especially pod and seed development, and their precise function will provide better new avenues for improving pod and seed size and thus grain yield in legumes.

10. ncRNAs Orchestrating Nodulation, Symbiosis, and Root Development Processes

Legumes are unique due to their inherent ability of forming root nodules in association with active soil rhizobacteria that assist in fixing atmospheric nitrogen [1]. The underlying molecular mechanism and around 200 genes involved in fixing atmospheric nitrogen in soil through nodulation and symbiosis have been deciphered [177,178]. Likewise, evidence of small RNAs, including miRNAs involved in nodule development and root symbiosis, has been reported in various model legumes, viz., M. truncatula, L. japonicus, and soybean [49,51,76,133,179,180,181,182,183]. The greater abundance of miR172 in root nodules than leaf tissue in Medicago truncatula [76], Lotus japonicus [138], common bean [60], and soybean [134] suggests its active role in nodulation. The role of MIR166 (targeting HD-ZIP III TF genes contributing to root nodule development) in Medicago truncatula was revealed by its overexpression, which downregulated HD-ZIP III, inhibiting symbiotic nodules and lateral root development [132]. Similarly, in soybean, miR166 and miR396 (targeting HD-ZIP III TF and cysteine protease gene, respectively) depicted downregulation during nodulation in soybean [49].
Considering the potential role of miRNAs involved in signalling pathways related to nodule infection and N2 fixation, De Luis et al. (2012) [138] demonstrated that the induction of miR171c in root nodules targeting NSP2 TF is correlated with bacterial nodule infection. While the induction of miR397 is noted strictly in rhizobial bacteria-infected active N2 fixing nodules, it participates in contributing to nitrogen fixation-related copper homeostasis and also targets the laccase copper protein family gene in Lotus japonicus [138]. Subsequently, the negative role of gma-miR171o and gma-miR171q miRNAs regulating soybean nodulation was functionally validated [184]. The authors demonstrated that the regulatory expression of two TF genes, GmSCL-6 and GmNSP2 (target genes of gma-miR171o and gma-miR171q miRNAs), plays an active role in the expression of NIN, ENOD40, and ERN genes involved in the nodulation process in soybean. Among the other miRNAs attributed to the nodulation process, the regulatory circuit of nodule development controlled by miRNA172-targeting AP2 and miRNA156-regulating miRNA172 expression in soybean has been investigated [49,134].
Various research groups [140,185,186] have suggested that the negative regulation of miR171h targeting MtNSP2 is needed for nodule formation and the mycorrhizal signalling pathway in Medicago truncatula. Overexpression of miR396b in roots of Medicago truncatula impaired root growth and diminished mycorrhizal colonization by targeting six growth-regulating factor genes (MtGRF) and two bHLH79-like genes, indicating the significant role of miR396b in root growth and mycorrhizal colonization [139] (see Table 2). Further insights into the underlying complete molecular mechanism of miR172c controlling rhizobial infection and precise nodulation regulation were elucidated in soybean [135]. The authors postulated that the absence of rhizobia Nodule Number Control1 (NNC1) suppresses the transcription of ENOD40 genes in soybean. However, in the presence of rhizobia, nod factor receptors induced a signal cascade that evokes the upregulation of miR172c targeting the NNC1 gene. Thus, the inhibition of NNC1 allows transcription of ENOD40 genes leading to nodule organogenesis in soybean (Figure 3).
Likewise, considering the underlying role of miR172a in rhizobial infection during symbiosis, Holt et al. (2015) [81] supported that the inductive activity of miR172a in L. japonicus roots requires the presence of both active rhizobial bacteria and bacterial Nod factor signalling during the early stage of symbiotic infection. Possible targets of miR172a were predicted to be the RAP2-7-like1, AP2-like1, and AP2-like2 genes during bacterial symbiosis. Subsequently, Yan et al. (2015) [84] functionally demonstrated that the overexpression of miR393j-3p miRNA targeting a nodulin gene Early Nodulin 93 (ENOD93) significantly inhibited nodule formation in soybean. Turner et al. (2012) [85] monitored the high expression of Glyma10g10240 and Glyma17g05920 (targets of miRNA169), which encode HAP proteins that contribute to nodule development.
The role of miR169 in regulating nodule development (transition from meristematic to differentiated cells) in M. truncatula by targeting the MtHAP2-1 novel symbiosis-specific TF gene has been established [133] (Figure 3). Li et al. (2010) [129] supported the role of miR482, miR1512, and miR1515 with enhanced nodule numbers at the transgenic level, thus suggesting their role in nodule development in soybean. However, Wang et al. (2015) [136] demonstrated that overexpression of miR156 in transgenic plants caused inhibited nodule development in Lotus japonicus. Similarly, in common bean, overexpression of miR319 the target TCP10 TF gene mRNA, which positively induces the action of the LOX2 gene involved in jasmonic acid synthesis [141], stimulated the nodule development but decreased the rhizobial infection process [141].
Furthermore, to gain deeper insight into the role of miRNAs regulating nodulation and the symbiosis process, Sós-Hegedűs et al. (2020) [142] established and functionally validated the regulatory mechanism of the nodulation and symbiosis process through silencing of target NB-LRR genes by miR2118, miR2109, and miR1507 miRNAs in Medicago truncatula. During nodulation and symbiotic nitrogen fixation, the symbiotic bacteria upregulate miR2118, miR2109, and miR1507 miRNAs, at the cost of downregulating NB-LRR genes; consequently, the plant’s innate immunity is compromised during symbiosis in nodules [142] (see Figure 2). Recently, Tsikou et al. (2018) [187] and Gautrat et al. (2020) [131] suggested that miR2111 targeting TOO MUCH LOVE (encoding F-box/kelch-repeat protein), a nodulation suppressor, could enhance nodulation. However, the prevalence of rhizobial inoculation/infection and nitrate treatment reduced the level of miR2111s in leaves and roots, depending on the shoot-acting hypernodulation and aberrant root 1 (HAR1) receptor. Moreover, describing the fine-tuning and autoregulation mechanism of nodulation, Gautrat et al. (2020) [131] postulated that the Clavata3/Embryo surrounding region 12 (CLE12) and the CLE13 signalling peptides synthesized in roots act through HAR1/super numeric nodule (SUNN) receptors to negatively regulate the action of miR2111 [130]. This miR2111 otherwise favours root symbiotic nodulation under nitrogen-starved conditions by C-terminally encoded peptide (CEP) produced in root and acts in shoot through the compact root architecture 2 (CRA2) receptor. Likewise, Okuma et al. (2020) [130] confirmed the regulatory role of HAR1-dependent miR2111s produced from the MIR2111-5 locus in shoots controlling root nodulation in Lotus japonicus using functional analysis.
Apart from these model legumes, three A. hypogaea-specific miRNAs, ahy-miR3508 (targeting gene encoding pectinesterase), ahy-miR3509, and ahy-miR3516, were identified; however, it is not known whether they participate in the nodulation process [108]. In common bean, genome-wide transcriptome analysis using Genome Analyzer IIx and degradome analysis identified 185 mature miRNAs and 181 targets for these identified miRNAs [100]. Functional characterization of selected miRNAs, viz., miRNov153 targeting uridine kinase (Phvul.003 g180800), miR319 targeting TCP TF family member (Phvul.011 g156900), and miR-Nov494 targeting aldehyde dehydrogenase (Phvul.004G162200.1), were upregulated, but their corresponding target genes were downregulated, indicating their significant involvement in controlling nodule development in common bean [100].
Furthermore, these miRNAs, an abundance of 21-nucleotide phased siRNAs derived from PHAS loci corresponding to protein coding genes NB-LRRs, were noted in soybean nodule [90] and in common bean nodule [100]. Likewise, evidence of circRNAs involved in nodule development and rhizobial symbiosis has been reported in common bean [188]. The authors suggested their role of acting as eTM and regulating the transmembrane transport and positive regulation of kinase activity during nodule development and the nitrogen fixation process. Recently, Tiwari et al. (2021) [189] and Hoang et al. (2020) [190] comprehensively discussed the interplay of various miRNAs impacting hormone signalling and regulating various regulatory genes during rhizobial infection, nodule organogenesis, and nitrogen fixation. A thorough understanding of various gene networks and their interplay with regulatory ncRNAs and precise function in controlling nodulation and related processes during the symbiosis process will further illuminate our insights into legume symbiosis at the molecular level involving ncRNAs.

11. Conclusions and Future Perspectives

The discovery of ncRNAs and their functional annotation have received considerable interest for investigating the underlying molecular mechanisms controlling various biological phenomena in legumes and opened a new avenue for improving traits of interest. As ncRNAs are dynamic, they are rapidly being discovered and functionally characterized in various plant species, including legumes [19]. However, the complete characterization of discovered ncRNAs at the functional level and their target gene(s) is limited to model legumes, viz., M. truncatula, L. japonicus, and soybean. Other legumes also need attention for the investigation of novel ncRNAs and their functions. Emerging approaches including powerful deep transcriptome sequencing technologies and advances in computational biology will facilitate the discovery of more ncRNAs and annotate their function. Moreover, emerging approaches of genome editing technology, viz., CRISPR/Cas9, will enable the functional characterization of novel ncRNAs (through loss-of-function/gain-of-function analysis) or manipulation of miRNAs causing the reprogramming of gene expression that controlling various traits of breeding importance with high precision [21,130,191]. Thus, the artificial manipulation of ncRNAs controlling various breeding traits could help us develop designer crops for sustaining global food security under predicted climate change scenarios.

Author Contributions

Conceptualization, U.C.J., K.H.M.S., and N.M.; writing—original draft preparation, U.C.J. and N.M.; writing—review and editing, U.C.J.; K.H.M.S., N.M., and H.N. All authors have read and agreed to the published version of the manuscript.

Funding

No funding is required for writing this review.

Institutional Review Board Statement

Ethical review and approval were waived for this study due to it being an extensive literature search study.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

U.C.J. thanks ICAR, New Delhi, India.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Graham, P.H.; Vance, C.P. Legumes: Importance and constraints to greater use. Plant Physiol. 2003, 131, 872–877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Edwards, T.J. Legumes of the World. S. Afr. J. Bot. 2007, 73, 272–273. [Google Scholar] [CrossRef] [Green Version]
  3. Jukanti, A.K.; Gaur, P.M.; Gowda, C.L.; Chibbar, R.N. Nutritional quality and health benefits of chickpea (Cicer arietinum L.): A review. Br. J. Nutr. 2012, 108, S11–S26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Koornneef, M.; Alonso-Blanco, C.; Peeters, A.J.M. Genetic approaches in plant physiology. New Phytol. 1997, 137, 1–8. [Google Scholar] [CrossRef]
  5. Song, X.; Li, Y.; Cao, X.; Qi, Y. MicroRNAs and their regulatory roles in plant-environment interactions. Annu. Rev. Plant Biol. 2019, 70, 489–525. [Google Scholar] [CrossRef]
  6. Kim, V.N. MicroRNA biogenesis: Coordinated cropping and dicing. Nat. Rev. Mol. Cell Biol. 2005, 6, 376–385. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, X. Small RNAs and their roles in plant development. Annu. Rev. Cell Dev. 2009, 25, 21–44. [Google Scholar] [CrossRef] [Green Version]
  8. Ghildiyal, M.; Zamore, P.D. Small silencing RNAs: An expanding universe. Nat. Rev. Genet. 2009, 10, 94–108. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, K.C.; Chang, H.Y. Molecular mechanisms of long non-coding RNAs. Mol. Cell. 2011, 43, 904–914. [Google Scholar] [CrossRef] [Green Version]
  10. 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] [Green Version]
  11. Santosh, B.; Varshney, A.; Yadava, P.K. Non-coding RNAs: Biological functions and applications. Cell Biochem. Funct. 2014, 33, 14–22. [Google Scholar] [CrossRef] [PubMed]
  12. Srivastava, S.; Zhen, Y.; Kudapa, H.; Jagadeeswaran, G.; Hivrale, V.; Varshney, R.K.; Sunkar, R. High throughput sequencing of small RNA component of leaves and inflorescence revealed conserved and novel miRNAs as well as phasiRNA loci in chickpea. Plant Sci. 2015, 235, 46–57. [Google Scholar] [CrossRef]
  13. Li, Z.; Xu, H.; Li, Y.; Wan, X.; Ma, Z.; Cao, J.; Li, Z.; He, F.; Wang, Y.; Wan, L.; et al. Analysis of physiological and miRNA responses to Pi deficiency in alfalfa (Medicago sativa L.). Plant Mol. Biol. 2018, 96, 473–492. [Google Scholar] [CrossRef] [PubMed]
  14. Lv, L.; Yu, K.; Lü, H.; Zhang, X.; Liu, X.; Sun, C.; Xu, H.; Zhang, J.; He, X.; Zhang, D. Transcriptome-wide identification of novel circular RNAs in soybean in response to low-phosphorus stress. PLoS ONE 2020, 15, e0227243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Wang, T.Z.; Liu, M.; Zhao, M.G.; Chen, R.; Zhang, W.H. Identification and characterization of long non-coding RNAs involved in osmotic and salt stress in Medicago truncatula using genome-wide high-throughput sequencing. BMC Plant Biol. 2015, 15, 131. [Google Scholar] [CrossRef] [Green Version]
  16. Zhao, W.; Cheng, Y.; Zhang, C.; You, Q.; Shen, X.; Guo, W.; Jiao, Y. Genome-wide identification and characterization of circular RNAs by high throughput sequencing in soybean. Sci. Rep. 2017, 7, 5636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. 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. Expt. Bot. 2017, 68, 5937–5948. [Google Scholar] [CrossRef] [Green Version]
  18. 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] [PubMed] [Green Version]
  19. 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, 367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Ye, C.Y.; Chen, L.; Liu, C.; Zhu, Q.H.; Fan, L. Widespread noncoding circular RNAs in plants. New Phytol. 2015, 208, 88–95. [Google Scholar] [CrossRef]
  21. Liu, D.; Mewalal, R.; Hu, R.; Tuskan, G.A.; Yang, X. New technologies accelerate the exploration of non-coding RNAs in horticultural plants. Hortic. Res. 2017, 4, 17031. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Jones-Rhoades, M.W.; Bartel, D.P.; Bartel, B. MicroRNAs and their regulatory roles in plants. Annu. Rev. Plant Biol. 2006, 57, 19–53. [Google Scholar] [CrossRef]
  23. Reinhart, B.J.; Weinstein, E.G.; Rhoades, M.W.; Bartel, B.; Bartel, D.P. MicroRNAs in plants. Genes Dev. 2002, 16, 1616–1626. [Google Scholar] [CrossRef] [Green Version]
  24. Ha, M.; Kim, V.N. Regulation of microRNA biogenesis. Nat. Rev. Mol. Cell Biol. 2014, 15, 509–524. [Google Scholar] [CrossRef]
  25. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [Green Version]
  26. Voinnet, O. Origin, biogenesis, and activity of plant microRNAs. Cell 2009, 136, 669–687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Mallory, A.C.; Vaucheret, H. Functions of microRNAs and related small RNAs in plants. Nat. Genet. 2006, 38, S31–S36. [Google Scholar] [CrossRef]
  28. Fei, Q.; Xia, R.; Meyers, B.C. Phased, secondary, small interfering RNAs in posttranscriptional regulatory networks. Plant Cell 2013, 25, 2400–2415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Budak, H.; Kaya, S.B.; Cagirici, H.B. Long non-coding RNA in plants in the era of reference sequences. Front. Plant Sci. 2020, 11, 276. [Google Scholar] [CrossRef] [Green Version]
  30. 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] [PubMed]
  31. Laurent, G.S.; Wahlestedt, C.; Kapranov, P. The landscape of long noncoding RNA classification. Trends Genet. 2015, 31, 239–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Quinn, J.J.; Chang, H.Y. Unique features of long non-coding RNA biogenesis and function. Nat. Rev. Genet. 2016, 17, 47–62. [Google Scholar] [CrossRef]
  33. Herr, A.J.; Jensen, M.B.; Dalmay, T.; Baulcombe, D.C. RNA polymerase IV directs silencing of endogenous DNA. Science 2005, 308, 118–120. [Google Scholar] [CrossRef]
  34. 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]
  35. Heo, J.B.; Sung, S. Vernalization-mediated epigenetic silencing by a long intronic noncoding RNA. Science 2011, 331, 76–79. [Google Scholar] [CrossRef] [Green Version]
  36. Bardou, F.; Ariel, F.; Simpson, C.G.; Romero-Barrios, N.; Laporte, P.; Balzergue, S.; Brown, J.W.; Crespi, M. Long noncoding RNA modulates alternative splicing regulators in Arabidopsis. Dev. Cell 2014, 30, 166–176. [Google Scholar] [CrossRef] [Green Version]
  37. Jabnoune, M.; Secco, D.; Lecampion, C.; Robaglia, C.; Shu, Q.Y.; 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]
  38. Sousa, C.; Johansson, C.; Charon, C.; Manyani, H.; Sautter, C.; Kondorosi, A.; Crespi, M. Translational and structural requirements of the early nodulin gene enod40, a short-open reading frame-containing RNA, for elicitation of a cell-specific growth response in the alfalfa root cortex. Mol. Cell Biol. 2001, 21, 354–366. [Google Scholar] [CrossRef] [Green Version]
  39. Nejat, N.; Mantri, N. Emerging roles of long non-coding RNAs in plant response to biotic and abiotic stresses. Crit. Rev. Biotechnol. 2018, 38, 93–105. [Google Scholar] [CrossRef] [PubMed]
  40. 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]
  41. Jha, U.C.; Nayyar, H.; Jha, R.; Khurshid, M.; Zhou, M.; Mantri, N.; Siddique, K.H. Long non-coding RNAs: Emerging players regulating plant abiotic stress response and adaptation. BMC Plant Biol. 2020, 20, 466. [Google Scholar] [CrossRef]
  42. Zhang, P.; Li, S.; Chen, M. Characterization and function of circular RNAs in plants. Front. Mol. Biosci. 2020, 7, 91. [Google Scholar] [CrossRef] [PubMed]
  43. Arenas-Huertero, C.; Pérez, B.; Rabanal, F.; Blanco-Melo, D.; De la Rosa, C.; Estrada-Navarrete, G.; Sanchez, F.; Covarrubias, A.A.; Reyes, J.L. Conserved and novel miRNAs in the legume Phaseolus vulgaris in response to stress. Plant Mol. Biol. 2009, 70, 385–401. [Google Scholar] [CrossRef]
  44. Zhang, B. MicroRNA: A new target for improving plant tolerance to abiotic stress. J. Exp. Bot. 2015, 66, 1749–1761. [Google Scholar] [CrossRef]
  45. De la Rosa, C.; Lozano, L.; Castillo-Ramírez, S.; Covarrubias, A.A.; Reyes, J.L. Origin and evolutionary dynamics of the miR2119 and ADH1 regulatory module in Legumes. Genome Biol. Evol. 2020, 12, 2355–2369. [Google Scholar] [CrossRef]
  46. Cakir, O.; Candar-Cakir, B.; Zhang, B. Small RNA and degradome sequencing reveals important micro RNA function in Astragalus chrysochlorus response to selenium stimuli. Plant Biotech. J. 2016, 14, 543–556. [Google Scholar] [CrossRef] [PubMed]
  47. Gao, F.; Wang, N.; Li, H.; Liu, J.; Fu, C.; Xiao, Z.; Wei, C.; Lu, X.; Feng, J.; Zhou, Y. Identification of drought-responsive microRNAs and their targets in Ammopiptanthus mongolicus by using high-throughput sequencing. Sci. Rep. 2016, 6, 34601. [Google Scholar] [CrossRef] [Green Version]
  48. De la Rosa, C.; Covarrubias, A.A.; Reyes, J.L. A dicistronic precursor encoding miR398 and the legume-specific miR2119 coregulates CSD1 and ADH1 mRNAs in response to water deficit. Plant Cell Environ. 2019, 42, 133–144. [Google Scholar] [CrossRef]
  49. Subramanian, S.; Fu, Y.; Sunkar, R.; Barbazuk, W.B.; Zhu, J.K.; Yu, O. Novel and nodulation-regulated microRNAs in soybean roots. BMC Genom. 2008, 9, 160. [Google Scholar] [CrossRef] [Green Version]
  50. Szittya, G.; Moxon, S.; Santos, D.M.; Jing, R.; Fevereiro, M.P.; Moulton, V.; Dalmay, T. High-throughput sequencing of Medicago truncatula short RNAs identifies eight new miRNA families. BMC Genom. 2008, 9, 593. [Google Scholar] [CrossRef] [Green Version]
  51. Jagadeeswaran, G.; Zheng, Y.; Li, Y.F.; Shukla, L.I.; Matts, J.; Hoyt, P.; Macmil, S.L.; Wiley, G.B.; Roe, B.A.; Zhang, W.; et al. Cloning and characterization of small RNAs from Medicago truncatula reveals four novel legume-specific microRNA families. New Phytol. 2009, 184, 85–98. [Google Scholar] [CrossRef]
  52. Jones-Rhoades, M.W.; Bartel, D.P. Computational identification of plant microRNAs and their targets, including a stress-induced miRNA. Mol. Cell 2004, 14, 787–799. [Google Scholar] [CrossRef] [PubMed]
  53. 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]
  54. Sunkar, R.; Kapoor, A.; Zhu, J.K. Posttranscriptional induction of two Cu/Zn superoxide dismutase genes in Arabidopsis is mediated by downregulation of miR398 and important for oxidative stress tolerance. Plant Cell 2006, 18, 2415. [Google Scholar] [CrossRef] [Green Version]
  55. Barrera-Figueroa, B.E.; Gao, L.; Diop, N.N.; Wu, Z.; Ehlers, J.D.; Roberts, P.A.; Close, T.J.; Zhu, J.K.; Liu, R. Identification and comparative analysis of drought-associated microRNAs in two cowpea genotypes. BMC Plant Biol. 2011, 11, 127. [Google Scholar] [CrossRef] [Green Version]
  56. Chi, X.; Yang, Q.; Chen, X.; Wang, J.; Pan, L.; Chen, M.; Yang, Z.; He, Y.; Liang, X.; Yu, S. Identification and characterization of microRNAs from peanut (Arachis hypogaea L.) by high-throughput sequencing. PLoS ONE 2011, 6, e27530. [Google Scholar] [CrossRef]
  57. Das, A.; Nigam, D.; Junaid, A.; Tribhuvan, K.U.; Kumar, K.; Durgesh, K.; Singh, N.K.; Gaikwad, K. Expressivity of the key genes associated with seed and pod development is highly regulated via lncRNAs and miRNAs in pigeonpea. Sci. Rep. 2019, 9, 18191. [Google Scholar] [CrossRef] [Green Version]
  58. Trindade, I.; Capitão, C.; Dalmay, T.; Fevereiro, M.P.; Dos 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]
  59. Cao, C.; Long, R.; Zhang, T.; Kang, J.; Wang, Z.; Wang, P.; Sun, H.; Yu, J.; Yang, Q. Genome-wide identification of microRNAs in response to salt/alkali stress in Medicago truncatula through high-throughput sequencing. Int. J. Mol. Sci. 2018, 19, 4076. [Google Scholar] [CrossRef] [Green Version]
  60. Valdés-López, O.; Yang, S.S.; Aparicio-Fabre, R.; Graham, P.H.; Reyes, J.L.; Vance, C.P.; Hernández, G. MicroRNA expression profile in common bean (Phaseolus vulgaris) under nutrient deficiency stresses and manganese toxicity. New Phytol. 2010, 187, 805–818. [Google Scholar] [CrossRef]
  61. Golicz, A.A.; Singh, M.B.; Bhalla, P.L. The long intergenic noncoding RNA (LincRNA) landscape of the soybean genome. Plant Physiol. 2018, 176, 2133–2147. [Google Scholar] [CrossRef] [Green Version]
  62. Joshi, T.; Patil, K.; Fitzpatrick, M.R.; Franklin, L.D.; Yao, Q.; Cook, J.R.; Wang, Z.; Libault, M.; Brechenmacher, L.; Valliyodan, B.; et al. Soybean Knowledge Base (SoyKB): A web resource for soybean translational genomics. BMC Genom. 2012, 13, S15. [Google Scholar] [CrossRef] [Green Version]
  63. Yi, X.; Zhang, Z.; Ling, Y.; Xu, W.; Su, Z. PNRD: A plant non-coding RNA database. Nucleic Acids Res. 2015, 43, D982–D989. [Google Scholar] [CrossRef] [Green Version]
  64. Xuan, H.; Zhang, L.; Liu, X.; Han, G.; Li, J.; Li, X.; Liu, A.; Liao, M.; Zhang, S. PLNlncRbase: A resource for experimentally identified lncRNAs in plants. Gene 2015, 573, 328–332. [Google Scholar] [CrossRef] [PubMed]
  65. Gallart, A.P.; Pulido, A.H.; de Lagrán, I.A.M.; Sanseverino, W.; Cigliano, R.A. GREENC: A wiki-based database of plant lncRNAs. Nucleic Acid Res. 2016, 44, D1161–D1166. [Google Scholar] [CrossRef] [PubMed]
  66. 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, e183. [Google Scholar] [CrossRef]
  67. Tian, B.; Wang, S.; Todd, T.C.; Johnson, C.D.; Tang, G.; Trick, H.N. Genome-wide identification of soybean microRNA responsive to soybean cyst nematodes infection by deep sequencing. BMC Genom. 2017, 18, 572. [Google Scholar] [CrossRef]
  68. Kohli, D.; Joshi, G.; Deokar, A.A.; Bhardwaj, A.R.; Agarwal, M.; Katiyar-Agarwal, S.; Srinivasan, R.; Jain, P.K. Identification and characterization of wilt and salt stress-responsive microRNAs in chickpea through high-throughput sequencing. PLoS ONE 2014, 9, e108851. [Google Scholar] [CrossRef]
  69. Garg, V.; Khan, A.W.; Kudapa, H.; Kale, S.M.; Chitikineni, A.; Qiwei, S.; Sharma, M.; Li, C.; Zhang, B.; Xin, L.; et al. Integrated transcriptome, small RNA and degradome sequencing approaches provide insights into Ascochyta blight resistance in chickpea. Plant Biotech. J. 2019, 17, 914–931. [Google Scholar] [CrossRef] [Green Version]
  70. Yuan, M.; Ngou, B.P.M.; Ding, P.; Xin, X.F. 2021. PTI-ETI crosstalk: An integrative view of plant immunity. Curr. Opin. Plant Biol. 2021, 62, 102030. [Google Scholar] [CrossRef]
  71. Zipfel, C.; Robatzek, S. Pathogen-associated molecular pattern triggered immunity: Veni, vidi? Plant. Physiol. 2010, 154, 551–554. [Google Scholar] [CrossRef] [Green Version]
  72. Li, W.; Deng, Y.W.; Ning, Y.S.; Hu, Z.H.; Wang, G.L. Exploiting broad-spectrum disease resistance in crops: From molecular dissection to breeding. Annu. Rev. Plant. Biol. 2020, 71, 575–603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Song, L.; Fang, Y.; Chen, L.; Wang, J.; Chen, X. Role of non-coding RNAs in plant immunity. Plant. Commun. 2021, 20, 100180. [Google Scholar] [CrossRef]
  74. Li, X.; Wang, X.; Zhang, S.; Liu, D.; Duan, Y.; Dong, W. Identification of soybean microRNAs involved in soybean cyst nematode infection by deep sequencing. PLoS ONE 2012, 7, e39650. [Google Scholar] [CrossRef] [Green Version]
  75. Formey, D.; Sallet, E.; Lelandais-Brière, C.; Ben, C.; Bustos-Sanmamed, P.; Niebel, A.; Frugier, F.; Combier, J.P.; Debellé, F.; Hartmann, C.; et al. The small RNA diversity from Medicago truncatula roots under biotic interactions evidences the environmental plasticity of the miRNAome. Genome Biol. 2014, 15, 457. [Google Scholar] [CrossRef] [PubMed]
  76. Lelandais-Brière, C.; Naya, L.; Sallet, E.; Calenge, F.; Frugier, F.; Hartmann, C.; Gouzy, J.; Crespi, M. Genome-wide Medicago truncatula small RNA analysis revealed novel microRNAs and isoforms differentially regulated in roots and nodules. Plant. Cell 2009, 21, 2780–2796. [Google Scholar] [CrossRef] [Green Version]
  77. Zhou, Z.S.; Zeng, H.Q.; Liu, Z.P.; Yang, Z.M. Genome-wide identification of Medicago truncatula microRNAs and their targets reveals their differential regulation by heavy metal. Plant Cell Environ. 2012, 35, 86–99. [Google Scholar] [CrossRef]
  78. Chen, L.; Wang, T.; Zhao, M.; Tian, Q.; Zhang, W.H. Identification of aluminum-responsive microRNAs in Medicago truncatula by genome-wide high-throughput sequencing. Planta 2012, 235, 375–386. [Google Scholar] [CrossRef]
  79. Chen, L.; Wang, T.; Zhao, M.; Zhang, W. Ethylene-responsive miRNAs in roots of Medicago truncatula identified by high-throughput sequencing at the whole genome level. Plant. Sci. 2012, 184, 14–19. [Google Scholar] [CrossRef] [PubMed]
  80. Long, R.C.; Li, M.N.; Kang, J.M.; Zhang, T.J.; Sun, Y.; Yang, Q.C. Small RNA deep sequencing identifies novel and salt-stress-regulated microRNAs from roots of Medicago sativa and Medicago truncatula. Physiol. Plant. 2015, 154, 13–27. [Google Scholar] [CrossRef]
  81. Holt, D.B.; Gupta, V.; Meyer, D.; Abel, N.B.; Andersen, S.U.; Stougaard, J.; Markmann, K. micro RNA 172 (miR172) signals epidermal infection and is expressed in cells primed for bacterial invasion in Lotus japonicus roots and nodules. New Phytol. 2015, 208, 241–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. 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, 730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Sun, Z.; Wang, Y.; Mou, F.; Tian, Y.; Chen, L.; Zhang, S.; Jiang, Q.; Li, X. Genome-wide small RNA analysis of soybean reveals auxin-responsive microRNAs that are differentially expressed in response to salt stress in root apex. Front. Plant. Sci. 2016, 6, 1273. [Google Scholar] [CrossRef] [Green Version]
  84. Yan, Z.; Hossain, M.S.; Arikit, S.; Valdés-López, O.; Zhai, J.; Wang, J.; Libault, M.; Ji, T.; Qiu, L.; Meyers, B.C.; et al. Identification of micro RNA s and their mRNA targets during soybean nodule development: Functional analysis of the role of miR393j-3p in soybean nodulation. New Phytol. 2015, 207, 748–759. [Google Scholar] [CrossRef]
  85. Turner, M.; Yu, O.; Subramanian, S. Genome organization and characteristics of soybean microRNAs. BMC Genom. 2012, 13, 169. [Google Scholar] [CrossRef] [Green Version]
  86. Wang, Y.; Zhang, C.; Hao, Q.; Sha, A.; Zhou, R.; Zhou, X.; Yuan, L. Elucidation of miRNAs-mediated responses to low nitrogen stress by deep sequencing of two soybean genotypes. PLoS ONE 2013, 8, e67423. [Google Scholar] [CrossRef] [Green Version]
  87. Song, Q.X.; Liu, Y.F.; Hu, X.Y.; Zhang, W.K.; Ma, B.; Chen, S.Y.; Zhang, J.S. Identification of miRNAs and their target genes in developing soybean seeds by deep sequencing. BMC Plant. Biol. 2011, 11, 5. [Google Scholar] [CrossRef] [Green Version]
  88. Ma, B.; Zhang, A.; Zhao, Q.; Li, Z.; Lamboro, A.; He, H.; Li, Y.; Jiao, S.; Guan, S.; Liu, S.; et al. Genome-wide identification and analysis of long non-coding RNAs involved in fatty acid biosynthesis in young soybean pods. Sci. Rep. 2021, 11, 7603. [Google Scholar] [CrossRef]
  89. Ding, X.; Li, J.; Zhang, H.; He, T.; Han, S.; Li, Y.; Yang, S.; Gai, J. Identification of miRNAs and their targets by high-throughput sequencing and degradome analysis in cytoplasmic male-sterile line NJCMS1A and its maintainer NJCMS1B of soybean. BMC Genom. 2016, 17, 24. [Google Scholar] [CrossRef] [Green Version]
  90. Arikit, S.; Xia, R.; Kakrana, A.; Huang, K.; Zhai, J.; Yan, Z.; Valdés-López, O.; Prince, S.; Musket, T.A.; Nguyen, H.T.; et al. An atlas of soybean small RNAs identifies phased siRNAs from hundreds of coding genes. Plant. Cell 2014, 26, 4584–4601. [Google Scholar] [CrossRef] [Green Version]
  91. Khemka, N.; Singh, V.K.; Garg, R.; Jain, M. Genome-wide analysis of long intergenic non-coding RNAs in chickpea and their potential role in flower development. Sci. Rep. 2016, 6, 33297. [Google Scholar] [CrossRef] [Green Version]
  92. Jain, M.; Chevala, V.N.; Garg, R. Genome-wide discovery and differential regulation of conserved and novel microRNAs in chickpea via deep sequencing. J. Exp. Bot. 2014, 65, 5945–5958. [Google Scholar] [CrossRef] [Green Version]
  93. Khemka, N.; Singh Rajkumar, M.; Garg, R.; Jain, M. Genome-wide profiling of miRNAs during seed development reveals their functional relevance in seed size/weight determination in chickpea. Plant. Direct 2021, 5, e00299. [Google Scholar] [CrossRef]
  94. Pradhan, S.; Verma, S.; Chakraborty, A.; Bhatia, S. Identification and molecular characterization of miRNAs and their target genes associated with seed development through small RNA sequencing in chickpea. Func. Integr. Genom. 2021, 21, 283–298. [Google Scholar] [CrossRef]
  95. Khandal, H.; Parween, S.; Roy, R.; Meena, M.K.; Chattopadhyay, D. MicroRNA profiling provides insights into post-transcriptional regulation of gene expression in chickpea root apex under salinity and water deficiency. Sci. Rep. 2017, 7, 4632. [Google Scholar] [CrossRef] [Green Version]
  96. Peláez, P.; Trejo, M.S.; Iñiguez, L.P.; Estrada-Navarrete, G.; Covarrubias, A.A.; Reyes, J.L.; Sánchez, F. Identification and characterization of microRNAs in Phaseolus vulgaris by high-throughput sequencing. BMC Genom. 2012, 13, 83. [Google Scholar] [CrossRef] [Green Version]
  97. Patwa, N.; Nithin, C.; Bahadur, R.P.; Basak, J. Identification and characterization of differentially expressed Phaseolus vulgaris miRNAs and their targets during mungbean yellow mosaic India virus infection reveals new insight into Phaseolus-MYMIV interaction. Genomics 2019, 111, 1333–1342. [Google Scholar] [CrossRef]
  98. Parreira, J.R.; Cappuccio, M.; Balestrazzi, A.; Fevereiro, P.; de Sousa Araújo, S. MicroRNAs expression dynamics reveal post-transcriptional mechanisms regulating seed development in Phaseolus vulgaris L. Hortic. Res. 2021, 8, 18. [Google Scholar] [CrossRef]
  99. Mendoza-Soto, A.B.; Naya, L.; Leija, A.; Hernández, G. Responses of symbiotic nitrogen-fixing common bean to aluminum toxicity and delineation of nodule responsive microRNAs. Front. Plant. Sci. 2015, 6, 587. [Google Scholar] [CrossRef]
  100. Formey, D.; Iñiguez, L.P.; Peláez, P.; Li, Y.F.; Sunkar, R.; Sánchez, F.; Reyes, J.L.; Hernández, G. Genome-wide identification of the Phaseolus vulgaris sRNAome using small RNA and degradome sequencing. BMC Genom. 2015, 16, 423. [Google Scholar] [CrossRef] [Green Version]
  101. Vlasova, A.; Capella-Gutiérrez, S.; Rendón-Anaya, M.; Hernández-Oñate, M.; Minoche, A.E.; Erb, I.; Câmara, F.; Prieto-Barja, P.; Corvelo, A.; Sanseverino, W.; et al. Genome and transcriptome analysis of the Mesoamerican common bean and the role of gene duplications in establishing tissue and temporal specialization of genes. Genome Biol. 2016, 17, 32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Tian, H.; Guo, F.; Zhang, Z.; Ding, H.; Meng, J.; Li, X.; Peng, Z.; Wan, S. Discovery, identification, and functional characterization of long noncoding RNAs in Arachis hypogaea L. BMC Plant. Biol. 2020, 20, 308. [Google Scholar] [CrossRef]
  103. Zhao, X.; Gan, L.; Yan, C.; Li, C.; Sun, Q.; Wang, J.; Yuan, C.; Zhang, H.; Shan, S.; Liu, J.N. Genome-wide identification and characterization of Long non-coding RNAs in Peanut. Genes 2019, 10, 536. [Google Scholar] [CrossRef] [Green Version]
  104. Wang, M.; Zhang, C.X.; Pan, L.J.; Wang, T.; Chen, N.; Chen, M.N.; Yang, Z.; Guo, X.G.; Yu, S.L.; Chi, X.Y. Small RNA profiling reveal regulation of microRNAs in field peanut pod rot pathogen infection. Biologia 2020, 75, 1779–1788. [Google Scholar] [CrossRef]
  105. Gao, C.; Wang, P.; Zhao, S.; Zhao, C.; Xia, H.; Hou, L.; Ju, Z.; Zhang, Y.; Li, C.; Wang, X. Small RNA profiling and degradome analysis reveal regulation of microRNA in peanut embryogenesis and early pod development. BMC Genom. 2017, 18, 220. [Google Scholar] [CrossRef] [Green Version]
  106. Ram, M.K.; Mukherjee, K.; Pandey, D.M. Identification of miRNA, their targets and miPEPs in peanut (Arachis hypogaea L.). Comput. Biol. Chem. 2019, 83, 107100. [Google Scholar] [CrossRef]
  107. Ma, X.; Zhang, X.; Zhao, K.; Li, F.; Li, K.; Ning, L.; He, J.; Xin, Z.; Yin, D. Small RNA and degradome deep sequencing reveals the roles of microRNAs in seed expansion in peanut (Arachis hypogaea L.). Front. Plant Sci. 2018, 9, 349. [Google Scholar] [CrossRef]
  108. Figueredo, M.S.; Formey, D.; Rodríguez, J.; Ibáñez, F.; Hernández, G.; Fabra, A. Identification of miRNAs linked to peanut nodule functional processes. J. Biosci. 2020, 45, 62. [Google Scholar] [CrossRef]
  109. Chen, H.; Yang, Q.; Chen, K.; Zhao, S.; Zhang, C.; Pan, R.; Cai, T.; Deng, Y.; Wang, X.; Chen, Y.; et al. Integrated microRNA and transcriptome profiling reveals a miRNA-mediated regulatory network of embryo abortion under calcium deficiency in peanut (Arachis hypogaea L.). BMC Genom. 2019, 20, 392. [Google Scholar] [CrossRef] [Green Version]
  110. Zhang, X.; Ma, X.; Ning, L.; Li, Z.; Zhao, K.; Li, K.; He, J.; Yin, D. Genome-wide identification of circular RNAs in peanut (Arachis hypogaea L.). BMC Genom. 2019, 20, 653. [Google Scholar] [CrossRef]
  111. Ma, X.; Zhang, X.; Traore, S.M.; Xin, Z.; Ning, L.; Li, K.; Zhao, K.; Li, Z.; He, G.; Yin, D. Genome-wide identification and analysis of long noncoding RNAs (lncRNAs) during seed development in peanut (Arachis hypogaea L.). BMC Plant Biol. 2020, 20, 192. [Google Scholar] [CrossRef]
  112. Martins, T.F.; Souza, P.F.; Alves, M.S.; Silva, F.D.A.; Arantes, M.R.; Vasconcelos, I.M.; Oliveira, J.T. Identification, characterization, and expression analysis of cowpea (Vigna unguiculata [L.] Walp.) miRNAs in response to cowpea severe mosaic virus (CPSMV) challenge. Plant Cell Rep. 2020, 39, 1061–1078. [Google Scholar] [CrossRef]
  113. Shui, X.R.; Chen, Z.W.; Li, J.X. MicroRNA prediction and its function in regulating drought-related genes in cowpea. Plant Sci. 2013, 210, 25–35. [Google Scholar] [CrossRef] [PubMed]
  114. Paul, S.; Kundu, A.; Pal, A. Identification and validation of conserved microRNAs along with their differential expression in roots of Vigna unguiculata grown under salt stress. Plant Cell Tissue Organ Cult. 2011, 105, 233–242. [Google Scholar] [CrossRef]
  115. Nithin, C.; Thomas, A.; Basak, J.; Bahadur, R.P. Genome-wide identification of miRNAs and lncRNAs in Cajanus cajan. BMC Genom. 2017, 18, 878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Alzahrani, S.M.; Alaraidh, I.A.; Khan, M.A.; Migdadi, H.M.; Alghamdi, S.S.; Alsahli, A.A. Identification and characterization of salt-responsive microRNAs in Vicia faba by high-throughput sequencing. Genes 2019, 10, 303. [Google Scholar] [CrossRef] [Green Version]
  117. Paul, S.; Kundu, A.; Pal, A. Identification and expression profiling of Vigna mungo microRNAs from leaf small RNA transcriptome by deep sequencing. J. Integr. Plant Biol. 2014, 56, 15–23. [Google Scholar] [CrossRef] [PubMed]
  118. DeBoer, K.; Melser, S.; Sperschneider, J.; Kamphuis, L.G.; Garg, G.; Gao, L.L.; Frick, K.; Singh, K.B. Identification and profiling of narrow-leafed lupin (Lupinus angustifolius) microRNAs during seed development. BMC Genom. 2019, 20, 135. [Google Scholar] [CrossRef]
  119. Zhu, Y.Y.; Zeng, H.Q.; Dong, C.X.; Yin, X.M.; Shen, Q.R.; Yang, Z.M. microRNA expression profiles associated with phosphorus deficiency in white lupin (Lupinus albus L.). Plant Sci. 2010, 178, 23–29. [Google Scholar] [CrossRef]
  120. Glazińska, P.; Kulasek, M.; Glinkowski, W.; Wojciechowski, W.; Kosiński, J. Integrated analysis of small RNA, transcriptome and degradome sequencing provides new insights into floral development and abscission in yellow lupine (Lupinus luteus l.). Int. J. Mol. Sci. 2019, 20, 5122. [Google Scholar] [CrossRef] [Green Version]
  121. Bhat, K.V.; Mondal, T.K.; Gaikwad, A.B.; Kole, P.R.; Chandel, G.; Mohapatra, T. Genome-wide identification of drought-responsive miRNAs in grass pea (Lathyrus sativus L.). Plant Gene 2020, 21, 100210. [Google Scholar] [CrossRef]
  122. 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]
  123. Wu, J.; Wang, L.; Wang, S. MicroRNAs associated with drought response in the pulse crop common bean (Phaseolus vulgaris L.). Gene 2017, 628, 78–86. [Google Scholar] [CrossRef]
  124. Zheng, Y.; Hivrale, V.; Zhang, X.; Valliyodan, B.; Lelandais-Brière, C.; Farmer, A.D.; May, G.D.; Crespi, M.; Nguyen, H.T.; Sunkar, R. Small RNA profiles in soybean primary root tips under water deficit. BMC Syst. Biol. 2016, 10, 126. [Google Scholar] [CrossRef] [Green Version]
  125. Jovanović, Ž.; Stanisavljević, N.; Mikić, A.; Radović, S.; Maksimović, V. Water deficit down-regulates miR398 and miR408 in pea (Pisum sativum L.). Plant Physiol. Biochem. 2014, 83, 26–31. [Google Scholar] [CrossRef] [PubMed]
  126. 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] [PubMed] [Green Version]
  127. 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, 20, 99. [Google Scholar] [CrossRef] [Green Version]
  128. Yan, Z.; Hossain, M.S.; Valdés-López, O.; Hoang, N.T.; Zhai, J.; Wang, J.; Libault, M.; Brechenmacher, L.; Findley, S.; Joshi, T.; et al. Identification and functional characterization of soybean root hair micro RNA s expressed in response to Bradyrhizobium japonicum infection. Plant Biotech. J. 2016, 14, 332–341. [Google Scholar] [CrossRef]
  129. Li, H.; Deng, Y.; Wu, T.; Subramanian, S.; Yu, O. Misexpression of miR482, miR1512, and miR1515 increases soybean nodulation. Plant Physiol. 2010, 153, 1759–1770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Okuma, N.; Soyano, T.; Suzaki, T.; Kawaguchi, M. MIR2111-5 locus and shoot-accumulated mature miR2111 systemically enhance nodulation depending on HAR1 in Lotus japonicus. Nat. Commun. 2020, 11, 5192. [Google Scholar] [CrossRef]
  131. Gautrat, P.; Laffont, C.; Frugier, F. Compact root architecture 2 promotes root competence for nodulation through the miR2111 systemic effector. Curr. Biol. 2020, 30, 1339–1345. [Google Scholar] [CrossRef] [PubMed]
  132. Boualem, A.; Laporte, P.; Jovanović, M.; Laffont, C.; Plet, J.; Combier, J.P.; Niebel, A.; Crespi, M.; Frugier, F. MicroRNA166 controls root and nodule development in Medicago truncatula. Plant J. 2008, 54, 876–887. [Google Scholar] [CrossRef]
  133. Combier, J.P.; Frugier, F.; de Billy, F.; Boualem, A.; El-Yahyaoui, F.; Moreau, S.; Vernié, T.; Ott, T.; Gamas, P.; Crespi, M.; et al. MtHAP2-1 is a key transcriptional regulator of symbiotic nodule development regulated by microRNA169 in Medicago truncatula. Genes Dev. 2006, 20, 3084–3088. [Google Scholar] [CrossRef] [Green Version]
  134. Yan, Z.; Hossain, M.S.; Wang, J.; Valdés-López, O.; Liang, Y.; Libault, M.; Qiu, L.; Stacey, G. miR172 regulates soybean nodulation. Mol. Plant-Microbe Int. 2013, 26, 1371–1377. [Google Scholar] [CrossRef] [Green Version]
  135. Wang, Y.; Wang, L.; Zou, Y.; Chen, L.; Cai, Z.; Zhang, S.; Zhao, F.; Tian, Y.; Jiang, Q.; Ferguson, B.J.; et al. Soybean miR172c targets the repressive AP2 transcription factor NNC1 to activate ENOD40 expression and regulate nodule initiation. Plant Cell 2014, 26, 4782–4801. [Google Scholar] [CrossRef] [Green Version]
  136. Wang, Y.; Wang, Z.; Amyot, L.; Tian, L.; Xu, Z.; Gruber, M.Y.; Hannoufa, A. Ectopic expression of miR156 represses nodulation and causes morphological and developmental changes in Lotus japonicus. Mol. Genet. Genom. 2015, 290, 471–484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Yang, W.C.; Katinakis, P.; Hendriks, P.; Smolders, A.; de Vries, F.; Spee, J.; van Kammen, A.; Bisseling, T.; Franssen, H. Characterization of GmENOD40, a gene showing novel patterns of cell-specific expression during soybean nodule development. Plant J. 1993, 3, 573–585. [Google Scholar] [CrossRef]
  138. De Luis, A.; Markmann, K.; Cognat, V.; Holt, D.B.; Charpentier, M.; Parniske, M.; Stougaard, J.; Voinnet, O. Two microRNAs linked to nodule infection and nitrogen-fixing ability in the legume Lotus japonicus. Plant Physiol. 2012, 160, 2137–2154. [Google Scholar] [CrossRef] [Green Version]
  139. Bazin, J.; Khan, G.A.; Combier, J.P.; Bustos-Sanmamed, P.; Debernardi, J.M.; Rodríguez, R.; Sorin, C.; Palatnik, J.; Hartmann, C.; Crespi, M.; et al. miR396 affects mycorrhization and root meristem activity in the legume Medicago truncatula. Plant J. 2013, 74, 920–934. [Google Scholar] [CrossRef] [PubMed]
  140. Lauressergues, D.; Delaux, P.M.; Formey, D.; Lelandais-Brière, C.; Fort, S.; Cottaz, S.; Bécard, G.; Niebel, A.; Roux, C.; Combier, J.P. The microRNA miR171h modulates arbuscular mycorrhizal colonization of Medicago truncatula by targeting NSP2. Plant J. 2012, 72, 512–522. [Google Scholar] [CrossRef]
  141. Martín-Rodríguez, J.Á.; Leija, A.; Formey, D.; Hernández, G. The MicroRNA319d/TCP10 node regulates the common bean–rhizobia nitrogen-fixing symbiosis. Front. Plant Sci. 2018, 9, 1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Sós-Hegedűs, A.; Domonkos, Á.; Tóth, T.; Gyula, P.; Kaló, P.; Szittya, G. Suppression of NB-LRR genes by miRNAs promotes nitrogen-fixing nodule development in Medicago truncatula. Plant Cell Environ. 2020, 43, 1117–1129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Kundu, A.; Paul, S.; Dey, A.; Pal, A. High throughput sequencing reveals modulation of microRNAs in Vigna mungo upon Mungbean Yellow Mosaic India Virus inoculation highlighting stress regulation. Plant Sci. 2017, 257, 96–105. [Google Scholar] [CrossRef]
  144. Liu, J.Q.; Allan, D.L.; Vance, C.P. Systemic signaling and local sensing of phosphate in common bean: Cross-talk between photosynthate and microRNA399. Mol. Plant 2010, 3, 428–437. [Google Scholar] [CrossRef] [Green Version]
  145. Farooq, M.; Wahid, A.; Kobayashi, N.S.M.A.; Fujita, D.B.S.M.A.; Basra, S.M.A. Plant drought stress: Effects, mechanisms and management. Agron. Sustain. Dev. 2009, 29, 153–188. [Google Scholar] [CrossRef] [Green Version]
  146. Fang, Y.; Xiong, L. General mechanisms of drought response and their application in drought resistance improvement in plants. Cell. Mol. Life Sci. 2015, 72, 673–689. [Google Scholar] [CrossRef]
  147. Jha, U.C.; Bohra, A.; Nayyar, H. Advances in “omics” approaches to tackle drought stress in grain legumes. Plant Breed. 2020, 139, 1–27. [Google Scholar] [CrossRef]
  148. Dasmandal, T.; Rao, A.R.; Sahu, S. Identification and characterization of circular RNAs regulating genes responsible for drought stress tolerance in chickpea and soybean. Indian J. Genet. 2020, 80, 1–8. [Google Scholar] [CrossRef]
  149. Jha, U.C.; Bohra, A.; Jha, R.; Parida, S.K. Salinity stress response and ‘omics’ approaches for improving salinity stress tolerance in major grain legumes. Plant Cell Rep. 2019, 38, 255–277. [Google Scholar] [CrossRef]
  150. Jha, U.C.; Bohra, A. Genomics enabled breeding approaches for improving cadmium stress tolerance in plants. Euphytica 2016, 208, 1–31. [Google Scholar] [CrossRef]
  151. Feng, S.J.; Zhang, X.D.; Liu, X.S.; Tan, S.K.; Chu, S.S.; Meng, J.G.; Zhao, K.X.; Zheng, J.F.; Yang, Z.M. Characterization of long non-coding RNAs involved in cadmium toxic response in Brassica napus. RSC Adv. 2016, 6, 82157. [Google Scholar] [CrossRef]
  152. Zhou, Z.S.; Huang, S.Q.; Yang, Z.M. Bioinformatic identification and expression analysis of new microRNAs from Medicago truncatula. Biochem. Biophys. Res. Commun. 2008, 374, 538–542. [Google Scholar] [CrossRef]
  153. Schachtman, D.P.; Reid, R.J.; Ayling, S.M. Phosphorus uptake by plants: From soil to cell. Plant Physiol. 1998, 116, 447–453. [Google Scholar] [CrossRef] [Green Version]
  154. Masclaux-Daubresse, C.; Daniel-Vedele, F.; Dechorgnat, J.; Chardon, F.; Gaufichon, L.; Suzuki, A. Nitrogen uptake, assimilation and remobilization in plants: Challenges for sustainable and productive agriculture. Ann. Bot. 2010, 105, 1141–1157. [Google Scholar] [CrossRef] [Green Version]
  155. Smith, F.W.; Mudge, S.R.; Rae, A.L.; Glassop, D. Phosphate transport in plants. Plant Soil 2003, 248, 71–83. [Google Scholar] [CrossRef]
  156. Fan, X.; Naz, M.; Fan, X.; Xuan, W.; Miller, A.J.; Xu, G. Plant nitrate transporters: From gene function to application. J. Exp. Bot. 2017, 68, 2463–2475. [Google Scholar] [CrossRef] [PubMed]
  157. Uchida, R. Essential nutrients for plant growth: Nutrient functions and deficiency symptoms. Plant Nutr. Manag. Hawaiis Soils 2000, 4, 31–55. [Google Scholar]
  158. Simons, M.; Saha, R.; Guillard, L.; Clément, G.; Armengaud, P.; Cañas, R.; Maranas, C.D.; Lea, P.J.; Hirel, B. Nitrogen-use efficiency in maize (Zea mays L.): From ‘omics’ studies to metabolic modelling. J. Exp. Bot. 2014, 65, 5657–5671. [Google Scholar] [CrossRef] [Green Version]
  159. Xu, Z.; Zhong, S.; Li, X.; Li, W.; Rothstein, S.J.; Zhang, S.; Bi, Y.; Xie, C. Genome-wide identification of microRNAs in response to low nitrate availability in maize leaves and roots. PLoS ONE 2011, 6, e28009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Hsieh, L.C.; Lin, S.I.; Shih, A.C.C.; Chen, J.W.; Lin, W.Y.; Tseng, C.Y.; Li, W.H.; Chiou, T.J. Uncovering small RNA-mediated responses to phosphate deficiency in Arabidopsis by deep sequencing. Plant Physiol. 2009, 151, 2120–2132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Rausch, C.; Bucher, M. Molecular mechanisms of phosphate transport in plants. Planta 2002, 216, 23–37. [Google Scholar] [CrossRef]
  162. Chiou, T.J.; Aung, K.; Lin, S.I.; Wu, C.C.; Chiang, S.F.; Su, C.L. Regulation of phosphate homeostasis by microRNA in Arabidopsis. Plant Cell 2006, 18, 412–421. [Google Scholar] [CrossRef] [Green Version]
  163. 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] [PubMed] [Green Version]
  164. 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, 655. [Google Scholar] [CrossRef] [Green Version]
  165. Zeng, H.Q.; Zhu, Y.Y.; Huang, S.Q.; Yang, Z.M. Analysis of phosphorus-deficient responsive miRNAs and cis-elements from soybean (Glycine max L.). J. Plant Physiol. 2010, 167, 1289–1297. [Google Scholar] [CrossRef] [PubMed]
  166. Aung, B.; Gruber, M.Y.; Amyot, L.; Omari, K.; Bertrand, A.; Hannoufa, A. Ectopic expression of LjmiR156 delays flowering, enhances shoot branching, and improves forage quality in alfalfa. Plant Biotech. Rep. 2015, 9, 379–393. [Google Scholar] [CrossRef]
  167. Gao, R.; Austin, R.S.; Amyot, L.; Hannoufa, A. Comparative transcriptome investigation of global gene expression changes caused by miR156 overexpression in Medicago sativa. BMC Genom. 2016, 17, 658. [Google Scholar] [CrossRef] [Green Version]
  168. Fujii, H.; Chiou, T.J.; Lin, S.I.; Aung, K.; Zhu, J.K. A miRNA involved in phosphate-starvation response in Arabidopsis. Curr. Biol. 2005, 15, 2038–2043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. 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]
  170. Bari, R.; Datt Pant, B.; Stitt, M.; Scheible, W.R. PHO2, microRNA399, and PHR1 define a phosphate-signaling pathway in plants. Plant Physiol. 2006, 141, 988–999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Wang, R.; Yang, Z.; Fei, Y.; Feng, J.; Zhu, H.; Huang, F.; Zhang, H.; Huang, J. Construction and analysis of degradome-dependent microRNA regulatory networks in soybean. BMC Genom. 2019, 20, 534. [Google Scholar] [CrossRef] [Green Version]
  172. Shin, S.Y.; Shin, C. Regulatory non-coding RNAs in plants: Potential gene resources for the improvement of agricultural traits. Plant Biotechnol. Rep. 2016, 10, 35–47. [Google Scholar] [CrossRef]
  173. Yan, J.; Cai, X.; Luo, J.; Sato, S.; Jiang, Q.; Yang, J.; Cao, X.; Hu, X.; Tabata, S.; Gresshoff, P.M.; et al. The REDUCED LEAFLET genes encode key components of the trans-acting small interfering RNA pathway and regulate compound leaf and flower development in Lotus japonicus. Plant Physiol. 2010, 152, 797–807. [Google Scholar] [CrossRef] [Green Version]
  174. Zhou, C.; Han, L.; Fu, C.; Wen, J.; Cheng, X.; Nakashima, J.; Ma, J.; Tang, Y.; Tan, Y.; Tadege, M.; et al. The trans-acting short interfering RNA3 pathway and no apical meristem antagonistically regulate leaf margin development and lateral organ separation, as revealed by analysis of an argonaute7/lobed leaflet1 mutant in Medicago truncatula. Plant Cell 2013, 25, 4845–4862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Yu, J.Y.; Zhang, Z.G.; Huang, S.Y.; Han, X.; Wang, X.Y.; Pan, W.J.; Qin, H.T.; Qi, H.D.; Yin, Z.G.; Qu, K.X.; et al. Analysis of miRNAs targeted storage regulatory genes during soybean seed development based on transcriptome sequencing. Genes 2019, 10, 408. [Google Scholar] [CrossRef] [Green Version]
  176. Jia, J.; Ji, R.; Li, Z.; Yu, Y.; Nakano, M.; Long, Y.; Feng, L.; Qin, C.; Lu, D.; Zhan, J.; et al. Soybean DICER-LIKE2 Regulates Seed Coat Color via Production of Primary 22-Nucleotide Small Interfering RNAs from Long Inverted Repeats. Plant Cell 2020, 32, 3662–3673. [Google Scholar] [CrossRef]
  177. Mus, F.; Crook, M.B.; Garcia, K.; Costas, A.G.; Geddes, B.A.; Kouri, E.D.; Paramasivan, P.; Ryu, M.H.; Oldroyd, G.E.; Poole, P.S.; et al. Symbiotic nitrogen fixation and the challenges to its extension to nonlegumes. Appl. Environ. Microbiol. 2016, 82, 3698–3710. [Google Scholar] [CrossRef] [Green Version]
  178. Roy, S.; Liu, W.; Nandety, R.S.; Crook, A.; Mysore, K.S.; Pislariu, C.I.; Frugoli, J.; Dickstein, R.; Udvardi, M.K. Celebrating 20 years of genetic discoveries in legume nodulation and symbiotic nitrogen fixation. Plant Cell 2020, 32, 15–41. [Google Scholar] [CrossRef] [Green Version]
  179. Wang, Y.; Li, P.; Cao, X.; Wang, X.; Zhang, A.; Li, X. Identification and expression analysis of miRNAs from nitrogen-fixing soybean nodules. Biochem. Biophys. Res. Commun. 2009, 378, 799–803. [Google Scholar] [CrossRef] [PubMed]
  180. Joshi, T.; Yan, Z.; Libault, M.; Jeong, D.H.; Park, S.; Green, P.J.; Sherrier, D.J.; Farmer, A.; May, G.; Meyers, B.C.; et al. Prediction of novel miRNAs and associated target genes in Glycine max. BMC Bioinform. 2010, 11, S14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Kulcheski, F.R.; de Oliveira, L.F.; Molina, L.G.; Almerão, M.P.; Rodrigues, F.A.; Marcolino, J.; Barbosa, J.F.; Stolf-Moreira, R.; Nepomuceno, A.L.; Marcelino-Guimarães, F.C.; et al. Identification of novel soybean microRNAs involved in abiotic and biotic stresses. BMC Genom. 2011, 12, 307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Molina, L.G.; Cordenonsi da Fonseca, G.; Morais, G.L.D.; de Oliveira, L.F.V.; Carvalho, J.B.D.; Kulcheski, F.R.; Margis, R. Metatranscriptomic analysis of small RNAs present in soybean deep sequencing libraries. Genet. Mol. Biol. 2012, 35, 292–303. [Google Scholar] [CrossRef]
  183. Zabala, G.; Campos, E.; Varala, K.K.; Bloomfield, S.; Jones, S.I.; Win, H.; Tuteja, J.H.; Calla, B.; Clough, S.J.; Hudson, M.; et al. Divergent patterns of endogenous small RNA populations from seed and vegetative tissues of Glycine max. BMC Plant Biol. 2012, 12, 177. [Google Scholar] [CrossRef] [Green Version]
  184. Hossain, M.S.; Hoang, N.T.; Yan, Z.; Tóth, K.; Meyers, B.C.; Stacey, G. Characterization of the spatial and temporal expression of two soybean miRNAs identifies SCL6 as a novel regulator of soybean nodulation. Front. Plant Sci. 2019, 10, 475. [Google Scholar] [CrossRef]
  185. Branscheid, A.; Devers, E.A.; May, P.; Krajinski, F. Distribution pattern of small RNA and degradome reads provides information on miRNA gene structure and regulation. Plant Signal. Bhav. 2011, 6, 1609–1611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Devers, E.A.; Branscheid, A.; May, P.; Krajinski, F. Stars and symbiosis: microRNA-and microRNA*-mediated transcript cleavage involved in arbuscular mycorrhizal symbiosis. Plant Physiol. 2011, 156, 1990–2010. [Google Scholar] [CrossRef] [Green Version]
  187. Tsikou, D.; Yan, Z.; Holt, D.B.; Abel, N.B.; Reid, D.E.; Madsen, L.H.; Bhasin, H.; Sexauer, M.; Stougaard, J.; Markmann, K. Systemic control of legume susceptibility to rhizobial infection by a mobile microRNA. Science 2018, 362, 233–236. [Google Scholar] [CrossRef] [PubMed]
  188. Wu, Z.; Huang, W.; Qin, E.; Liu, S.; Liu, H.; Grennan, A.K.; Liu, H.; Qin, R. Comprehensive Identification and Expression Profiling of Circular RNAs During Nodule Development in Phaseolus vulgaris. Front. Plant Sci. 2020, 11, 1638. [Google Scholar] [CrossRef] [PubMed]
  189. Tiwari, M.; Pandey, V.; Singh, B.; Bhatia, S. Dynamics of miRNA mediated regulation of legume symbiosis. Plant Cell Environ. 2021, 44, 1279–1291. [Google Scholar] [CrossRef]
  190. Hoang, N.T.; Tóth, K.; Stacey, G. The role of microRNAs in the legume–Rhizobium nitrogen-fixing symbiosis. J. Exp. Bot. 2020, 71, 1668–1680. [Google Scholar] [CrossRef]
  191. Sunkar, R.; Chinnusamy, V.; Zhu, J.; Zhu, J.K. Small RNAs as big players in plant abiotic stress responses and nutrient deprivation. Trends Plant Sci. 2007, 12, 301–309. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Example of major classes of ncRNA regulation for growth and development processes and stress tolerance in legume plants [12,13,14,15,16,17].
Figure 1. Example of major classes of ncRNA regulation for growth and development processes and stress tolerance in legume plants [12,13,14,15,16,17].
Cells 10 01674 g001
Figure 2. ncRNA module controlling various abiotic and biotic responses and developmental pathways in legume plants. Increased expression of Cc_lncRNA-2830 sequesters miR160h, resulting in upregulation of Auxin responsive factor-18 allowing proper pod development [57]. The role of Soy_25 miRNA targeting Glyma05g33260 gene attributing seed development is noteworthy [87] in soybean. In response to aschochyta blight attack, downregulation of miR482b-3p and miR159k-3p enhance expression of NBS-LRR and PR, respectively, inhibiting pathogen attack [69]. Under water stress, the downregulatory activity of miR398 and miR2119 increases the expression of CSD1 and ADH1 genes contributing to drought tolerance [48]. Under excess salinity stress, induction of miR172a cleaves mRNA transcripts of salt-suppressed AP2 domain-containing genes, allowing high expression of thiamine biosynthesis gene THI1 that ultimately enables transcription of the salinity tolerance regulator in soybean [126]. For nutrient deficiency stress, such as phosphate, downregulation of PDIL2 and PDIL3 lncRNAs increases the expression of Medtr1g074930 and phosphate uptake [17]. The repressive action of gma-miR396b/c/d/f/g-5p upregulates Glyma05g20930 and Glyma06g18790 genes, increasing N uptake [86]. During mercury metal stress, induction of miR2681, miR2708, and miR2687 enhances expression of the TIR-NBS-LRR/(XTH) gene imparting resistance against mercury [77]. During nodulation and symbiosis, miR2111 inhibits expression of the TOO MUCH LOVE gene, upregulating the nodule development process [130], while upregulation of miR2118, miR2109, and miR1507 enables nodulation by repressing NB-LRR genes [142].
Figure 2. ncRNA module controlling various abiotic and biotic responses and developmental pathways in legume plants. Increased expression of Cc_lncRNA-2830 sequesters miR160h, resulting in upregulation of Auxin responsive factor-18 allowing proper pod development [57]. The role of Soy_25 miRNA targeting Glyma05g33260 gene attributing seed development is noteworthy [87] in soybean. In response to aschochyta blight attack, downregulation of miR482b-3p and miR159k-3p enhance expression of NBS-LRR and PR, respectively, inhibiting pathogen attack [69]. Under water stress, the downregulatory activity of miR398 and miR2119 increases the expression of CSD1 and ADH1 genes contributing to drought tolerance [48]. Under excess salinity stress, induction of miR172a cleaves mRNA transcripts of salt-suppressed AP2 domain-containing genes, allowing high expression of thiamine biosynthesis gene THI1 that ultimately enables transcription of the salinity tolerance regulator in soybean [126]. For nutrient deficiency stress, such as phosphate, downregulation of PDIL2 and PDIL3 lncRNAs increases the expression of Medtr1g074930 and phosphate uptake [17]. The repressive action of gma-miR396b/c/d/f/g-5p upregulates Glyma05g20930 and Glyma06g18790 genes, increasing N uptake [86]. During mercury metal stress, induction of miR2681, miR2708, and miR2687 enhances expression of the TIR-NBS-LRR/(XTH) gene imparting resistance against mercury [77]. During nodulation and symbiosis, miR2111 inhibits expression of the TOO MUCH LOVE gene, upregulating the nodule development process [130], while upregulation of miR2118, miR2109, and miR1507 enables nodulation by repressing NB-LRR genes [142].
Cells 10 01674 g002
Figure 3. Role of selected miRNAs regulating nodulation process in legume plant [81,84,131,133,135,138,178].
Figure 3. Role of selected miRNAs regulating nodulation process in legume plant [81,84,131,133,135,138,178].
Cells 10 01674 g003
Table 1. List of published ncRNAs in legume plants regulating growth and development and biotic and abiotic stress responses.
Table 1. List of published ncRNAs in legume plants regulating growth and development and biotic and abiotic stress responses.
Number of ncRNACropGenotypeTraitTissueReferences
416 miRNAsM. truncatulaJemalong A17Symbiosis and pathogenic interactionsRoots[75]
100 novel candidate miRNAsM. truncatula Root and nodule development[76]
201 individual miRNAsM. truncatulaJemalongHeavy metalSeedlings[77]
326 known miRNAs and 21 new miRNAsM. truncatulaJemalong A17Aluminium toxicityRoot apices[78]
301 known miRNAs and identified 3 new miRNAsM. truncatulaEthylene responseRoots[79]
26 novel miRNAsM. truncatulaJemalongLeaves[50]
385 conserved miRNAs and 68 novel miRNAsM. truncatula Medicago sativaJemalong A17, Zhongmu-1Salinity stressRoots[80]
876 miRNAsM. truncatulaR108SalinitySeedlings[59]
100 novel candidate miRNAsM. truncatulaJemalong A17Root and nodule developmentRoots[76]
8 miRNAsM. truncatulaJemalongRoots, shoots[51]
219 novel L. japonicus micro RNAsLotus japonicusGifu wild-typeEpidermal and cortical signalling events[81]
3030 long intergenic noncoding RNAs (lincRNAs), 275 natural antisense transcripts (lncNATs)SoybeanWilliams 82SalinityRoots[82]
55 families of miRNAsSoybeanWilliams82NodulationRoots[49]
5372 circRNAsSoybeanDevelopmental processStems, roots, mature leaves[16]
537 known and 70 putative novel miRNAsSoybeanKS4607, KS4313NSoybean cyst nematodeRoots[67]
71 miRNAsSoybeanWilliams 82SalinityRoots[83]
364 + 21SoybeanHairbin xiaoheidou, Liaodou 10Soybean cyst nematodeRoots[74]
284 miRNAsSoybeanWilliams 82NodulationRoots[84]
120 miRNA genesSoybeanWilliams82Root, nodule, organ developmentRoots, stems, young leaves[85]
362 known miRNAsSoybeanNo.116, No.84-70Nitrogen stressRoots, shoots[86]
38+8 miRNAsSoybeanHeinong44Seed developmentSeeds[87]
6018 lincRNAsSoybeanVarious agronomic traitFlower buds, unopened flowers, florescence, pods, seeds[61]
46 lncRNAsSoybeanMT72 and JN18Fatty acid synthesisPods[88]
158 novel miRNAs and 160 high-confidence soybean miRNAsSoybeanNJCMS1A, NJCMS1BMale sterilityFlower buds[89]
500 loci generating phasiRNAs from PHAS lociSoybeanWilliams 82Reproductive developmentAnther and ovary tissues[90]
2248 lincRNAsChickpea Flower developmentVegetative tissues, shoot apical meristem, young leaves[91]
59 novel miRNAsChickpeaICC4958Fusarium wilt, salinityRoots[68]
157 miRNA lociChickpeaICC4958Stress responseLeaves, inflorescence[12]
440 conserved miRNAs + 178 novel miRNAsChickpeaICC4958Diverse cellular processes and metabolismLeaves, stems, flower buds, young pods[92]
651 miRNAsChickpeaC 214, Pb 7, ILC 3279, ICCV 05530, BC3F6Aschochyta blightSeedlings[69]
113 +243 miRNAsChickpeaJGK3 and Himchana1Seed size and developmentSeeds[93]
74 known and 26 novel miRNAsChickpeaSeed developmentSeeds[94]
3457 high-confidence lncRNAsChickpeaICC4958, ICC1882, ICCV2, JG62Drought and salinity[66]
284 unique miRNAsChickpeaBGD72Drought and salinityRoots[95]
114 miRNAsCommon bean Leaves, flower, roots[96]
422 miRNAsCommon bean MYMIVLeaves[97]
68 miRNAsCommon bean Nutrient deficiency and manganese toxicity stressLeaves, roots, nodules[60]
72 known and 39 new miRNAsCommon beanSER16Seed developmentSeeds[98]
28 miRNAsCommon beanNegro Jamapa 81Aluminium toxicityNodules[99]
185 mature miRNAsCommon beanNegro Jamapa, Pinto VillaN2-fixing symbiotic nodulesFlowers, leaves, roots, seedlings[100]
197 lncRNAsCommon beanBAT93Fruit developmentFlowers, pods, seeds, leaves, roots, stems[101]
16 conserved miRNAsCommon beanNegro Jamapa, Pinto VillaDifferent stress[43]
1442+ 189 lncRNAsGroundnutFenghua-1Development, growth and stress toleranceRoots, leaves, seeds[102]
50,873 lncRNAsGroundnut Growth and development15 different tissues[103]
334 peanut miRNAsGroundnutHuayu 20Pod rot [104]
70 known and 24 novel miRNAsGroundnutLuhua-14Pod developmentGynophores[105]
126 known miRNAs + 25 novel peanutGroundnut DevelopmentLeaves, stems, roots, seeds[56]
18 miRNAsGroundnut Disease resistant proteins, auxin responsive proteins[106]
1,082 miRNAsGroundnut8106, 8107Seed expansionSeeds[107]
32 miRNAsGroundnut Nodule developmentNodules[108]
29 known and 132 potential novel miRNAsGroundnutBaisha1016Ca deficiency[109]
347 circRNAsGroundnutRIL 8106, RIL 8107Seed development and size[110]
9388 known and 4037 novel lncRNAsGroundnutHuayou 7, Huayou 4Seed developmentSeeds[111]
617 mature microRNAsCowpea Cowpea severe mosaic virusLeaves[112]
17 new miRNAsCowpeaDan lla, Tvu7778DroughtLeaves, roots[113]
157 miRNA genesCowpeaCB46, IT93K503-1DroughtLeaves[55]
18 miRNAsCowpea Salinity stressRoots[114]
616 mature miRNAs + 3919 lncRNAsPigeonpea[115]
3919 lncRNAsPigeonpea [115]
3019 lncRNAs and 227 miRNAsPigeonpeaAshaSeed and pod developmentSeeds, pods[57]
298 upregulated and 395 downregulated
284 upregulated and 243 downregulated
Faba beanHassawi-3
ILB4347
SalinityLeaves[116]
66 miRNAsUrd bean Leaves, stems, roots[117]
56miRNAsNarrow-leafed lupinTanjilSeed developmentStems, leaves, seeds[118]
167 miRNAsWhite lupin Phosphate deficiencyRoots, stems, leaves[119]
394 known and 28 novel miRNAs and 316 phased siRNAsYellow lupineTaperFloral development and abscissionFlowers[120]
143 and 128LathyrusIC-143067Drought[121]
47 and 44 miRNAsAlfalfa Phosphorus deficiencyRoots, shoots[13]
371 circRNAsSoybeanBogao, Nannong 94156Phosphorus deficiencyRoots[14]
Table 2. Role of ncRNAs controlling abiotic and biotic stresses and other growth and development in legume plants with possible molecular mechanisms involved.
Table 2. Role of ncRNAs controlling abiotic and biotic stresses and other growth and development in legume plants with possible molecular mechanisms involved.
Name of ncRNACropTrait/StressTarget Gene(s)/Protein Coding Gene(s)FunctionReferences
miR408ChickpeaDroughtDREBOverexpression represses plantacyanin encoding genes and controls DREB regulation under water stress[122]
16 drought-responsive miRNAsCommon beanDroughtTFs and protein kinasesControl drought stress by targeting various TFs and protein kinases[123]
6 downregulated and 6 upregulated miRNAsSoybeanDroughtAuxin signalling, plantacyanin, Cu/Zn superoxide dismutasesControl drought stress by targeting auxin signalling, plantacyanin and Cu/Zn superoxide dismutases encoding genes[124]
44 drought-responsive miRNAsCowpeaDroughtZinc finger family protein, serine/threonine protein kinaseInvolved in development and stress response[55]
vun-miR5021, vun-miR156b-3p, vun-miR5021, vun-miR156b, vun-miR156fCowpeaDroughtKelch repeat-containing F-box protein, CPRD86, P5CS, multicystatin gene, and glutathione reductaseInduce genes PLD (phospholipase D), APX (ascorbate peroxidase) and P5CS (delta 1-pyrroline-5-carboxylate synthase) under stress[113]
miR162, miR164, miR319, miR403, miR828, miR160a, miR160b, miR171e, vun_cand015, vun_cand033, vun_cand048, miR171b, miR171d, miR2111b, miR390b, and miR393, vun_cand001, vun_cand010, vun_cand041, vun_cand057CowpeaDroughtARF10, ARF8, zinc finger protein, basic-helix-loop-helix (bHLH), TF leucine-rich repeat transmembrane protein kinase, pentatricopeptide repeat-containing proteinInvolved in development and stress response[55]
miR398a/b, miR408PeaDroughtCopper superoxide dismutase, CSD1Reduce oxidative stress[125]
lsa-miR169b, lsa-miR1508a, lsa-miR319a, lsa-miR156a, lsa-miR398b, lsa-miR396d, lsa-miR166b, lsa-miR390a, lsa-miR167b, lsa-miR186, lsa-miR786, lsa-miR897, lsa- miR969 and lsa-miR1361, miR397, miR398, miR164, miR399LathyrusDroughtF-box, U-Box or protein coding genes involved in proline, betain, and osmolyte biosynthesis pathwayInduce osmo-protective compounds under stress[121]
ChickpeaDrought and salinityLACCASE4, COPPER SUPEROXIDE DISMUTASE (Cu-SOD), NAC1 and PHO2/UBC24Increase lateral root formation and improves uptake of K+ under salinity stress[95]
MIR2119 and MIR398aCommon beanDroughtALCOHOL DEHYDROGENASE 1 (ADH1) and COPPER-ZINC SUPEROXIDE DISMUTASE 1 (CSD1)By reducing oxidative stress[45,48]
pvu-miR2118Common beanDroughtControls drought stress[43]
miR169, miR398a/b and miR408M. truncatulaDrought stressCopper proteins COX5b, copper superoxide dismutase, and plantacyanin [58]
miR172aSoybeanSalinityGlyma.10G116600, Glyma.02G087400, Glyma.13G329700, Glyma.12G073300, Glyma.15G044400, Glyma.11G053800, AP2/EREBP-type TF gene SSAC1, thiamine biosynthesis gene THI1Induction cleaves mRNA transcripts of salt-suppressed AP2 domain-containing genes increasing expression of thiamine biosynthesis gene THI1 and resulting salinity tolerance[126]
18 conserved miRNAsCowpeaSalinity15 target genesControl plant development and root growth under stress conditions by targeting various TF genes viz., SBP, ARF, SPL, TCP, NFY, and AP2[114]
miR156_1, miR156_10, car-miR008, car-miR011, car-miR015ChickpeaSalinitySquamosa promoter-binding proteinTarget protein-encoding gene to control salinity stress[68]
lncRNA TCONS_ 00097188, TCONS_00046739, TCONS_00100258, TCONS_ 00118328, TCONS_00047650, lncRNA TCONS_ 00020253, TCONS_00116877Medicago truncatulaSalinityMedtr6g006990, cytochrome P450, Medtr3g069280, Medtr1g081900, Medtr7g094600Upregulate various gene expression contributing to salinity stress adaptation[15]
TCONS_ 00292946, TCONS_00176941, TCONS_00011551GroundnutSalinityControl salinity stress tolerance[102]
pvu-miR159.2Common beanSalinity[43]
miR160, miR156/157, miR159, miR169, miR172, miR408CowpeaSalinity stressAuxin response factor (ARF), squamosa promoter-binding protein (SBP), TCP family transcription factor, CCAAT-binding transcription factor (CBF), PHAP2B protein, APETALA2 protein (AP2), Basic blue copper protein/PlantacyaninTarget TFs and control salinity stress[114]
lncRNA MtCIR1Medicago truncatulaCold stressMtCBF genesControls cold tolerance[127]
soy_25SoybeanSeed developmentGlyma05g33260Controls seed development[87]
gma-miR168SoybeanGlyma16g34300
miR167, miR399, miR156, miR319, miR164, miR166, miR1507 and miR396Narrow leaf lupinSeed developmentGROWTH-REGULATING FACTOR (GRF) TF, SBP-box transcription factors, MYB transcription factors, Zinc finger domain proteins, molybdate transporter 1, calcium-transporting ATPase 8, TMV resistance protein N, lysine-specific demethylase JMJ16, nudix hydrolase proteinTarget TF (Class III HD-Zip, NAC) related to seed development process[118]
ahy_novel_miRn1 to ahy_novel_miRn132, miR3509, miR3511, and miR3512, miR159 and miR167, miR3514, miR3518GroundnutCa deficiency driven embryo abortionTCP3, AP2, EMB2750, GRFs, HsfB4, DIVARICATA, CYP707A1, CYP707A3Regulate embryo abnormality under Ca deficiency by modulating the target genes[109]
miR_18, miR_6, miR_11, miR_29, miR_6, miR_38, miR_6, pvu-miR399a, miR_18, miR_33, miR_16, pvu-miR156iCommon beanSeed developmentDEHYDRIN FAMILY PROTEIN (RAB18), DEAD BOX RNA HELICASE (PRH75) CESA3, LEUCINE-RICH PROTEIN KINASE FAMILY PROTEIN, PRH75, MEE9, EM1, PHO2, RAB18, PROTEIN KINASE SUPERFAMILY PROTEIN, DUF827, and SPL2Regulate these genes during various stages of seed development, viz., seed filling, maturation, and dormancy[98]
XR_001593099.1, MSTRG.18462.1, MSTRG.34915.1, MSTRG.41848.1, MSTRG.22884.1, MSTRG.12404.1, MSTRG.26719.1, MSTRG.35761.1, MSTRG.20033.1, MSTRG.13500.1, MSTRG.9304.1GroundnutSeed developmentXM_016114848.1, XM_ 016087708.1, XM_016309191.1, XM_ 016324297.1, XM_016327810.1, XM_016116309.1, XM_ 016335443.1, XM_ 016310265.1, XM_ 016091385.1Regulate groundnut seed development by modulating the target genes encoding MADS-box transcription factor 23-like, protein transport protein sec31-like, squamosa promoter-binding-like protein 14[111]
Ca_linc_0051 and Ca_linc_0139ChickpeaFlower development [91]
miR156/157, miR164, miR167, miR1088, miR172, miR396GroundnutPod developmentSPL, NAC, PPRP, AP2, GRFControl pod development[105]
Cc_lncRNA-2830PigeonpeaPod developmentmiR160h- Auxin responsive factor-18Upregulates Cc_lncRNA-2830, sequesters miR160h promoting expression of auxin responsive factor-18 and helps in pod formation[57]
gma-miR156b and gma-miR156f, gma-miR162a, gma-miR162b, gma- miR162c, gma-miR399d, gma-miR399e, gma- miR399f gma-miR399gSoybeanMale sterilityMADS-box transcription factor, NADP-dependent isocitrate dehydrogenase, 6-phosphogluconate dehydrogenase, NADH-ubiquinone oxidoreductaseTarget these genes and cause programmed cell death, ROS toxicity and energy deficiency[89]
lncRNA MSTRG.45502.1, lncRNAs MSTRG.40968.1SoybeanLipid metabolic processesXM_003538388.3,XM_006588497.2 00,061 [88]
miR393j-3pSoybeanNodule developmentEarly Nodulin 93 (ENOD93)Targets Early Nodulin 93 (ENOD93) gene and regulates nodule formation[84]
gma-miR2606b, gma-miR4416SoybeanNodule developmentMannosyl- oligosaccharide 1, 2-alpha-mannosidase, Rhizobium-induced peroxidase 1 (RIP1)-like peroxidase geneTarget these genes to positively and negatively regulate the nodulation process[128]
miR482, miR1512, miR1515SoybeanNodule developmentGm12g28730, Gm17g04060, Gm04g05920, Glyma09g27690Regulates nodulation process[129]
miR2111Lotus japonicusNodulationTOO MUCH LOVE, a nodulation suppressorLow expression after rhizobial infection relying on shoot-acting HYPERNODULATION ABERRANT ROOT FORMATION1 (HAR1) receptor[130]
miR2111M. truncatulaNodulation and symbiosisToo Much Love 1, Too Much Love 2Positively controls root symbiotic nodulation, which is systemic from shoots and depends on the CRA2 receptor[131]
MIR166M. truncatulaRoot and nodule developmentClass-III HD-ZIP genesOverexpression reduced the number of symbiotic nodules and lateral roots[132]
microRNA169M. truncatulaNodule developmentMtHAP2-1Regulates MtHAP2-1 gene controlling symbiotic nodule formation[133]
ahy-mi399, ahy-miR159, ahy-miR3508GroundnutNodule infectionPectinesterase geneRegulate nodulation development process[108]
miRNA 172SoybeanNodulationAP2 transcription factorControls miR172 expression and regulates AP2 TF activity[134]
miRNA 172cSoybeanNodulationNodule Number Control1Controls nodule formation by repressing its target gene[135]
miRNA156Lotus japonicusNodulationENOD genes, SymPK, POLLUX, CYCLOPS, Cerberus, and Nsp1, SPLsRepresses downstream target SPLs and other nodulation genes[136]
MtENOD40M. truncatulaNodule developmentRegulates re-localization of proteins[38]
GmENOD40SoybeanNodule developmentRegulates re-localization of proteins[137]
miR156e, miR156g, miR167bM. truncatulaSymbiosis signalsInduced by Myc-LCO and repressed by Nod signals [75]
miR172aLotus japonicusEpidermal infection during symbiosisAPETALA2-type (AP2) transcription factorsTargets AP2 TF and regulates bacterial symbiosis[81]
miR171 isoform, miR397Lotus japonicusN2 fixationLaccase copper protein family, Nodulation Signalling Pathway2Respond to symbiotic infection and nodule function[138]
miR396M. truncatulaRoot growth and mycorrhizal associationsGrowth-regulating factor genes (MtGRF) and two bHLH79-like target genesRegulates root growth and mycorrhizal associations[139]
miR171hM. truncatulaMycorrhizal colonizationNSP2Targets NSP2 and modulates mycorrhizal colonization[140]
miR1507, miR2118, miR2119, miR2199M. truncatulaPathogen infectionTIR-NBS-LRR proteins targeted by miR2118 auxin response factor (ARF)miRNA-mediated plant defence response[51]
miR319dCommon beanRhizobium N2 fixationTCP10 (Phvul.005G067950) [141]
miR1507, miR2109, miR2118M. truncatulaNodulation and symbiosisNB-LRR genesSuppress activity of NB-LRR genes and allow nodulation process[142]
ENOD40SoybeanNodule development[137]
ENOD40and M. truncatulaNodule development-[38]
617 mature microRNAsCowpeaCowpea severe mosaic virusKat-p80, DEAD-Box, GST, and SPB9Involved in defence response to CSMV[112]
vun-miR156a, vun-miR156b, vun-miR156b-3p, vun-miR156b-5p, vun-miR156f, vun-miR156 g, vun-miR157d, vun-miR2610a, vun-miR2673b, vun-miR5021 Ted2 protein gene, Glutathione reductase, R3H domain protein gene, P5CS, Phosphoribosylpyrophosphate amidotransferase, 5-aminoimidazole ribonucleotide carboxylase, R3H domain protein gene, Ted2 protein, 5-aminoimidazole ribonucleotide carboxylase, Vigna unguiculata extensine-like protein 3, Aspartic proteinase, CPRD86
miR156, miR159, miR160, miR166, miR398, miR1511, miR1514, miR2118, and novel vmu-miRn7, vmu-miRn8, vmu-miRn13, vmu-miRn14UrdbeanMYMIVNB-LRR, NAC, MYB, Zinc finger, CCAAT-box transcription factor, fructose 2-6 bisphosphate, HDZIP proteinParticipate in defence/immune response to MYMIV[143]
miR530ChickpeaFusarium wilt infectionZinc knuckle- and microtubule-associated proteinsRegulates plant defence against pathogen attack[68]
miR166ChickpeaHD-ZIPIII transcription factor[68]
car-miRNA008ChickpeaNatural defenceChalcone synthase (CHS) geneRegulates plant defence against pathogen attack[68]
car-miR2118, car-miR5213ChickpeaDefence responseTIR-NBS-LRRRegulate plant defence against pathogen attack[68]
miR156, miR159, miR160, miR162, miR164, miR168, miR172, miR393, miR408ChickpeaStress response and development processesSPB factor, MYB transcription factor, ARFs, DCL1, HD-Zip, Arg onaute 1, AP2, F-box protein, plantacyaninTarget superoxide dismutases, plantacyanin, laccases and F-box proteins genes during stress[12]
ahy-miR396e-5, ahy-miR3509-5p, ahy-miR166f, ahy-miR159bGroundnutPod rotc39419_g1_i1, c40055_g1_i3, c31393_g1_i1, c41016_g4_i1Related to defence response[104]
miR482b-3p, miR159k-3p, nov_miR66, miR171, miR162, miR167c, miR171bChickpeaAscochyta blight resistanceNBS-LRR, PR protein, serine-threonine kinase, PPR protein, Dicer-like gene (Ca_01367), Dof zinc finger (Ca_19433), ERF (Ca_00359) geneProduce pathogenesis-related protein, ROS activity, cell wall synthesis, hormone synthesis, R gene activation[69]
miR171, miR159, miR399, miR398, miR408, miR9750, miR2119, miR1512SoybeanRootknot nematodeATPase, Glycosyl hydrolases, multicopper oxidase, SOD, peroxidase, Glucose-6-phosphate dehydrogenase encoding genesRegulate PR genes, oxidative stress and defence response[67]
miR156/157, miR164, miR167 and miR1088, miR172, miR396GroundnutSPL, NAC, PPRP, AP2 GRFControl seed development[105]
miR157, miR156, miR170, miR172, miR319, miR398, pvu-miR159.2, pvu-miR2118, gma-miR1508, gma-miR1526, gma-miR1532, miR160, miR397, miR399, miR408, pvu-miR1509, pvu-miR1514aCommon beanManganese toxicityUpregulated miR157, miR156, miR170, miR172 and downregulated pvu-miR2118, gma-miR1508, gma-miR1526, and gma-miR1532, etc.[60]
miR2681, miR2708, miR2687M. truncatulaMercury toleranceTIR-NBS-LRR, TC114805, xyloglucan endotransglucosylase/ hydrolase (XTH)XTH helps in cell wall development under heavy metal stress[77]
Gm03circRNA1785Soybean gma-miR167c and GmARF6 and GmARF8 [16]
PDIL1, PDIL2, PDIL3M. truncatulaPhosphate starvationMtPHO2, Medtr1g074930Regulate phosphate uptake[17]
miR399Common beanPhosphorus deficiencyPvHAD1 [144]
gma-miR156b/6f-5p, gma-miR396b∼g-5p, gma-miR5372-5p, gma-miR159d-3p, gma-miR396b∼g-5pSoybeanNitrogen deficiencyGlyma07g31580, Glyma05g20930, Glyma06g18790, Glyma09g02600, Glyma05g23280, Glyma07g05550, Glyma16g02090, Glyma17g16750, Glyma19g44930, Glyma15g08010, Glyma19g01200Play role in protein degradation[86]
miR399, miR398, miR156, miR159, miR164, miR168, miR172, miR393, miR408AlfalfaPhosphate starvationPhosphate transporter, Copper chaperone for SOD, Squamosa promoter-binding-like (SPL), MYB TF, auxin response factor (ARF), GRAS, MATERegulate phosphate uptake[13]
circ_000232SoybeanPhosphorus deficiencyGlyma.13G117700Regulates P use efficiency[14]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chand Jha, U.; Nayyar, H.; Mantri, N.; Siddique, K.H.M. Non-Coding RNAs in Legumes: Their Emerging Roles in Regulating Biotic/Abiotic Stress Responses and Plant Growth and Development. Cells 2021, 10, 1674. https://doi.org/10.3390/cells10071674

AMA Style

Chand Jha U, Nayyar H, Mantri N, Siddique KHM. Non-Coding RNAs in Legumes: Their Emerging Roles in Regulating Biotic/Abiotic Stress Responses and Plant Growth and Development. Cells. 2021; 10(7):1674. https://doi.org/10.3390/cells10071674

Chicago/Turabian Style

Chand Jha, Uday, Harsh Nayyar, Nitin Mantri, and Kadambot H. M. Siddique. 2021. "Non-Coding RNAs in Legumes: Their Emerging Roles in Regulating Biotic/Abiotic Stress Responses and Plant Growth and Development" Cells 10, no. 7: 1674. https://doi.org/10.3390/cells10071674

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