Next Article in Journal
Role of Serum Cholesterol and Statin Use in the Risk of Prostate Cancer Detection and Tumor Aggressiveness
Next Article in Special Issue
Symbiotic Nitrogen Fixation in Legume Nodules: Metabolism and Regulatory Mechanisms
Previous Article in Journal
Age Dependent Changes in Cartilage Matrix, Subchondral Bone Mass, and Estradiol Levels in Blood Serum, in Naturally Occurring Osteoarthritis in Guinea Pigs
Previous Article in Special Issue
The Role of Symbiotic Nitrogen Fixation in Sustainable Production of Biofuels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification of Cold-Responsive miRNAs and Their Target Genes in Nitrogen-Fixing Nodules of Soybean

1
The State Key Laboratory of Plant Cell & Chromosome Engineering, Center of Agricultural Research Resources, Institute of Genetics and Developmental Biology, Chinese Academy of Sciences, 286 Huaizhong Road, Shijiazhuang, Hebei 050021, China
2
College of Life Sciences, University of Chinese Academy of Sciences, 19A Yuquan Road, Beijing 100049, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2014, 15(8), 13596-13614; https://doi.org/10.3390/ijms150813596
Submission received: 29 May 2014 / Revised: 14 July 2014 / Accepted: 17 July 2014 / Published: 5 August 2014

Abstract

:
As a warm climate species, soybean is highly sensitive to chilling temperatures. Exposure to chilling temperatures causes a significant reduction in the nitrogen fixation rate in soybean plants and subsequent yield loss. However, the molecular basis for the sensitivity of soybean to chilling is poorly understood. In this study, we identified cold-responsive miRNAs in nitrogen-fixing nodules of soybean. Upon chilling, the expression of gma-miR397a, gma-miR166u and gma-miR171p was greatly upregulated, whereas the expression of gma-miR169c, gma-miR159b, gma-miR319a/b and gma-miR5559 was significantly decreased. The target genes of these miRNAs were predicted and validated using 5' complementary DNA ends (5'-RACE) experiments, and qPCR analysis identified putative genes targeted by the cold-responsive miRNAs in response to chilling temperatures. Taken together, our results reveal that miRNAs may be involved in the protective mechanism against chilling injury in mature nodules of soybean.

1. Introduction

Cold stress, including chilling (<20 °C) and freezing (<0 °C), influences plant development and crop yields. The genetic and physiological mechanisms underlying chilling injury and tolerance are well understood in Arabidopsis. In order to adapt to cold stress, plants alter their metabolism and growth through reprogramming of gene expression [1,2]. The perception of a cold signal in plants leads to the activation of C-repeat/drought-responsive element binding factor (CBF)-dependent and CBF-independent transcriptional pathways that modulate the response to cold stress [3,4].
Numerous studies conducted in a wide range of species have shown that microRNA (miRNA)-based gene regulation is essential for coordinating plant responses to cold stress [5,6,7]. In Arabidopsis, the expression of miR165/166, miR319c and miR169 is induced by cold stress [8,9], while in Brachypodium, the expression of both miR172 and miR397 is up-regulated by cold treatment [10]. The reported targets of miR166 are PHB (At2g34710) and ATHB-8 (At4g32880), which are down-regulated by cold exposure [8]. Recently, an analysis of deep sequencing data from a wheat thermosensitive genic male sterile line showed that 78 unique miRNAs may play a regulatory role in male sterility during cold stress [11]. miR319 has also been shown to regulate cold responses in both sugarcane [12] and Populus [13].
Soybean (Glycine max) is a globally important crop and a key source of proteins for human and animal consumption. Soybean is also used as a natural nitrogen source in agriculture, because nitrogen-fixing bacteria in specialized organs, called “root nodules”, can effectively convert atmospheric N2 into ammonium [14]. It has been reported that symbiotic nitrogen fixation (SNF) is highly sensitive to abiotic stresses, including defoliation, drought, extended dark treatment and chilling [15,16,17]. Notably, soybean is a warm climate crop species; thus, its growth and productivity are sensitive to chilling temperatures. For example, when subjected to even a single low temperature below 15 °C (also known as dark chilling), nitrogen fixation rates can drop up to 45% and pod formation is completely inhibited, leading to a significant loss of plant productivity [16,18]. In the past two decades, we have made great progress in elucidating the physiological mechanisms involved in plant responses to chilling, including photosynthesis, carbon metabolism and nitrogen fixation [16,19,20]. However, our understanding of the molecular mechanisms that modulate plant responses to chilling remains limited. In particular, the molecular mechanisms by which nitrogen-fixing nodules respond to chilling and modulate SNF in soybean are elusive.
Recently, a genome-wide analysis of miRNAs revealed key regulatory roles of miRNAs in nodulation, including Bradyrhizobium japonicum infection, nodule formation and nitrogen-fixation [21,22,23,24,25,26]. However, there is no report on the roles of miRNAs and their target genes in SNF in response to low temperatures in the nitrogen-fixing nodules of soybean. In this study, we performed a deep sequencing analysis of miRNAs in nitrogen-fixing soybean nodules to determine the expression profile of miRNAs in response to cold. Eleven miRNAs that were highly responsive to cold treatment were identified. The putative target genes of the miRNAs were predicted, and the predicted cleavage sites were verified in vivo for a subset of these targets by rapid amplification of cDNA ends (5'-RACE, 5' complementary DNA ends) analysis. Based on these results, we found that in cold-treated soybean plants, miR166u, miR171p, miR2111f and miR169c may regulate different targets in mature nodules through mRNA degradation. Our results have important implications for miRNA regulation in mature nodules in response to low temperatures and the modulation of SNF.

2. Results and Discussion

2.1. Results

2.1.1. The Nitrogen Use Efficiency of Nitrogen-Fixing Nodules Was Decreased in Response to Low Temperature Treatment

To investigate the effect of low temperatures on the nitrogen-fixing ability of mature nodules, we first measured the leghemoglobin (Lb) content and nitrogenase activity (measured as acetylene (C2H2) reduction activity, or ARA) at 28 days after rhizobium inoculation in nitrogen-fixing nodules treated at 4 °C for 24 h. Lb is bright red in color, and Lb accumulation or nodule coloration is considered to be an indicator of the maturity and functionality of root nodules [27]. It has been shown that the Lb content of plant nodules is adversely affected by various abiotic stresses, including drought, cadmium, sodium and nitrite, and by senescence in many legumes, such as Lupinus albus, Pisum sativum, Medicago truncatula, Lotus japonicus and soybean [28,29,30,31,32,33]. Lowering of the Lb level is mainly due to the degradation and/or heme nitration of Lb, leading to a color change from pink to green [34,35]. As shown in Figure 1A,B, there was no apparent difference in nodule color between the cold-treated and untreated nodules. The Lb content of the cold-treated nodules was also not significantly changed compared with that of the untreated control nodules (Figure 1C).
Next, the ARA of nitrogenase was analyzed using gas chromatography as described by Boyd [36]. Intriguingly, we found that nitrogenase activity in the functional nodules was markedly decreased by approximately 60% 24 h after chilling treatment (Figure 1D). Nitrogenases are used by rhizobia to fix atmospheric N2, and their enzymatic activity is directly correlated with nitrogen fixation efficiency [37]. Our results therefore suggest that the nitrogen-fixing ability of mature nodules is influenced by chilling treatment. Previous studies have shown that nitrogenase activity is sensitive to various environmental stresses, including drought, heat and nitrite [37,38,39,40,41]. Our results indicate that the reduced nitrogen fixing efficiency of the nodules caused by chilling was mainly due to decreased nitrogenase activity, rather than the Lb content.

2.1.2. Identification and Validation of Low Temperature-Responsive miRNAs in Functional Nodules

To explore whether there was any connection between miRNA expression and the decrease in nitrogenase activity, we used Solexa sequencing technology to analyze the miRNAs in cold-treated (CH) and untreated (CK) mature nodules (Figure S1 and Tables S1 and S2). Equal amounts of total RNA from CK and CH were pooled and used to construct two small regulatory RNA (sRNA) libraries to augment the chance of finding as many miRNAs as possible in a single experiment. Overall, more than 33 million raw reads were obtained from the two libraries. Clean reads (excluding reads smaller than 18 nucleotides (nt) and adaptors) accounted for around 89% of the total (Table S2). For unique sRNAs (i.e., sRNAs with a unique sequence), more than 42% of the reads from the CK library and more than 35% of the reads from the CH library could be mapped to the soybean genome, respectively (Table S2).
Figure 1. The nitrogen use efficiency of nitrogen-fixing nodules in soybean was affected by low temperature treatment. Seeds of Williams 82 were sown into a mixture of vermiculite and perlite and irrigated with a low nitrogen solution. Twenty-eight days after inoculation with Bradyrhizobium japonicum, the plants were dipped and washed with phosphate-buffered saline and then transferred to a low nitrogen solution without (CK) or with cold treatment (CH) at 4 °C for 24 h. The nodules were then collected and used for a phenotypic analysis and measurements of the leghemoglobin (Lb) content and acetylene reduction activity (ARA). Cross-sections of untreated and cold-treated nodules are shown in (A,B), respectively. Bars: 2 mm; The measured Lb contents and ARA levels are shown in (C,D), respectively. Student’s t-tests were performed; statistically significant results are marked with *** (p < 0.001).
Figure 1. The nitrogen use efficiency of nitrogen-fixing nodules in soybean was affected by low temperature treatment. Seeds of Williams 82 were sown into a mixture of vermiculite and perlite and irrigated with a low nitrogen solution. Twenty-eight days after inoculation with Bradyrhizobium japonicum, the plants were dipped and washed with phosphate-buffered saline and then transferred to a low nitrogen solution without (CK) or with cold treatment (CH) at 4 °C for 24 h. The nodules were then collected and used for a phenotypic analysis and measurements of the leghemoglobin (Lb) content and acetylene reduction activity (ARA). Cross-sections of untreated and cold-treated nodules are shown in (A,B), respectively. Bars: 2 mm; The measured Lb contents and ARA levels are shown in (C,D), respectively. Student’s t-tests were performed; statistically significant results are marked with *** (p < 0.001).
Ijms 15 13596 g001
We found that 61 miRNAs belonging to known miRNA families were responsive to low-temperature treatment. To validate our sequencing results, we performed quantitative reverse transcription PCR (RT-qPCR) to analyze the expression of the miRNAs in nitrogen-fixing nodules with or without low temperature treatment. We found that the expression level of eleven miRNAs was highly affected by low temperature treatment. These miRNAs were used for further study. As shown in Figure 2, six miRNAs were clearly upregulated by low-temperature treatment in mature nodules (Figure 2A). Among them, the expression levels of gma-miR397a, gma-miR166u and gma-miR171p were increased more than three-fold, whereas the expression of gma-miR167c, gma-miR399 and gma-miR2111f showed a fractional increase (Figure 2A). The rest of the miRNAs were downregulated by different degrees in low temperature-treated nitrogen-fixing nodules (Figure 2B). The expression levels of gma-miR169c, gma-miR159b and gma-miR319a/b were reduced by approximately 50%. These results suggest that miRNAs and miRNA-mediated molecular pathways modulate nitrogen fixation in low temperature-treated nodules.
To obtain additional information about the potential roles of miRNAs in nitrogen fixation at low temperatures, we analyzed the promoters (2 kb in length) of the miRNAs, gma-miR166u, gma-miR171p and gma-miR2111f, using plantPAN [42]. Interestingly, in addition to cis elements for binding auxin response factors, we found a binding site (MYCCONSENSUSAT) for inducer of CBF expression 1 (ICE1) and cis elements for binding cold-responsive transcription factors, such as RAV1. Taken together, our results reveal novel roles for these miRNAs in regulating the response of nitrogen-fixing nodules to low temperatures (Figures S2 and S3).
Figure 2. Quantitative reverse transcription PCR (RT-qPCR) analysis of miRNA expression in low temperature-treated mature nodules. Small RNAs were isolated from functional soybean nodules 28 days after inoculation with B. japonicum strain USDA110 in the cold (4 °C for 24 h); untreated nodules were used as a control. (A) The miRNAs that were up-regulated by low-temperature treatment are shown; (B) The miRNAs that were down-regulated by low-temperature treatment are shown. 5.8S rRNA was used as an internal control. The results shown are the averages ± SE of three biological replicates. Error bars indicate the standard deviation. Student’s t-tests were performed; statistically significant results are marked with * (p < 0.05); ** (p < 0.01); *** (p < 0.001); ns: not significant (p > 0.05).
Figure 2. Quantitative reverse transcription PCR (RT-qPCR) analysis of miRNA expression in low temperature-treated mature nodules. Small RNAs were isolated from functional soybean nodules 28 days after inoculation with B. japonicum strain USDA110 in the cold (4 °C for 24 h); untreated nodules were used as a control. (A) The miRNAs that were up-regulated by low-temperature treatment are shown; (B) The miRNAs that were down-regulated by low-temperature treatment are shown. 5.8S rRNA was used as an internal control. The results shown are the averages ± SE of three biological replicates. Error bars indicate the standard deviation. Student’s t-tests were performed; statistically significant results are marked with * (p < 0.05); ** (p < 0.01); *** (p < 0.001); ns: not significant (p > 0.05).
Ijms 15 13596 g002

2.1.3. Target Prediction and Validation for the Identified Low Temperature-Responsive miRNAs

miRNAs regulate gene expression post-transcriptionally by promoting mRNA decay or by inhibiting translation [24]. To identify candidate target genes of our low temperature-responsive miRNAs, we used two prediction algorithms: psRNAtarget [43] and the target search tool in WMD3 [44]. As shown in Table 1, 48 genes were predicted to be targets of the 11 low temperature-responsive miRNAs (Table 1). These putative target genes encode a variety of proteins, including a multicopper oxidase, putative lysophospholipase, transporters and transcription factors, indicating that the miRNAs play multiple regulatory roles in the cold response of soybean.
To validate whether these genes are directly targeted by the corresponding miRNAs, we performed a 5'-RACE assay to confirm the cleavage of the predicted target genes by miR166u, miR2111f, miR171p and miR169c (Figure 4). We found that all of the predicted target genes encoding basic leucine zipper (bZIP) or GAI-RGA-SCR (GRAS) family transcription factors (glyma11g17490 and glyma01g18040 for gma-miR171p and glyma07g01940 and glyma08g21620 for of gma-miR166u) were cleaved at a high frequency (Figure 3). In addition, two genes, glyma16g06160 and glyma19g25770, encoding Kelch-related proteins, were validated as targets of miR2111f. Glyma02g35190, which encodes a HAP2-like transcription factor, was also directly cleaved by gma-miR169c. All of the target genes were cleaved by only one miRNA at a single cleavage site. While the target genes of gma-miR166u and gma-miR2111f were cleaved at the expected sites (between the 10th and 11th nucleotides (nt) from the complementary 5' end of the miRNAs), we found that the cleavage sites of the target genes for gma-miR171p and gma-miR169c were located between the 13th/14th and 11th/12th nt, respectively.
We also analyzed the promoters of the validated target genes and found multiple cis elements for binding cold-responsive transcription factors (e.g., ICE1) and auxin response factors (Figure S2). Our findings indicate that these genes are functional targets of the miRNAs in cold-treated soybean.
Table 1. Putative target genes of known cold-responsive miRNAs and their function annotations.
Table 1. Putative target genes of known cold-responsive miRNAs and their function annotations.
miRNATarget GeneTarget Annotation
gma-miR397aGlyma01g26750 a,bMulticopper oxidase, oxidation/reduction
Glyma18g07240 bMulticopper oxidase, oxidation/reduction
Glyma03g15800 a,bMulticopper oxidase, oxidation/reduction
Glyma03g14450 bMulticopper oxidase, oxidation/reduction
Glyma06g43700 bMulticopper oxidase, oxidation/reduction
Glyma12g14230 bMulticopper oxidase, oxidation/reduction
gma-miR166uGlyma08g21620 a,cbZIP transcription factor
Glyma07g01940 a,cHomeobox-leucine zipper protein
gma-miR171pGlyma01g18040 a,b,cGRAS family transcription factor
Glyma11g17490 a,cGRAS family transcription factor
gma-miR2111fGlyma19g01430 a,bEamA-like transporter family
Glyma15g34831 a,bEamA-like transporter family
Glyma16g06160 b,cKelch motif; Protin Binding; Kelch-related protein
Glyma19g25770 b,cKelch motif; Protin Binding; Kelch-related protein
gma-miR169cGlyma02g35190 a,b,cHAP2 like transcription factor
Glyma19g38800 a,bHAP2 like transcription factor
Glyma13g16770 a,bHAP2 like transcription factor
Glyma07g04050 a,bHAP2 like transcription factor
Glyma13g27230 a,bHAP2 like transcription factor
gma-miR5037aGlyma17g13680 a,bGRAS family transcription factor
Glyma05g03020 a,bGRAS family transcription factor
gma-miR167cGlyma11g31940 a,bAuxin response factor
Glyma18g05330 a,bAuxin response factor
Glyma02g40650 a,bAuxin response factor
Glyma14g03650 a,bAuxin response factor
Glyma14g38940 a,bAuxin response factor
Glyma08g10550 a,bAuxin response factor
gma-miR319a/bGlyma13g04030 a,bMYB family transcription factor
Glyma13g25720 a,bMYB family transcription factor
Glyma15g35860 a,bMYB family transcription factor
Glyma20g11040 a,bMYB family transcription factor
Glyma06g43720 a,bTCP family transcription factor
Glyma12g14200 a,bTCP family transcription factor
gma-miR399iGlyma10g04230 a,bInorganic phosphate and sugar transporter
Glyma14g36650 bInorganic phosphate and sugar transporter
Glyma19g34710 aInorganic phosphate and sugar transporter
Glyma20g34620 aInorganic phosphate and sugar transporter
gma-miR1523aGlyma13g25040 bPutative lysophospholipase
Glyma13g25050 bPutative lysophospholipase
gma-miR5559Glyma20g34670 bMediator complex subunit 28
Glyma02g19340 bNo functional annotation
Glyma14g02030 bProtein binding
Glyma02g46640 bProtein binding
a predicted by psRNAtarget; b predicted by WMD3; c miRNA targets validated by RLM-5'-RACE.
Figure 3. Validation of the miRNA-directed cleavage of the putative target genes using rapid amplification of 5' complementary DNA ends (5'-RACE). The predicted targets of gma-miR171p, gma-miR166u, gma-miR2111f and gma-miR169c were experimentally validated. The mRNA cleavage sites were determined by modified 5' RNA ligase-mediated (5'-RLM)-RACE. The mRNA sequence of each complementary site and its 5' and 3' flanking sequences (5 nucleotides (nt) from 5' to 3') and cloned miRNA sequence (from 3' to 5') are shown. Vertical arrows indicate the 5' termini of the miRNA-guided cleavage products, as identified by 5'-RACE, with the frequency of clones shown.
Figure 3. Validation of the miRNA-directed cleavage of the putative target genes using rapid amplification of 5' complementary DNA ends (5'-RACE). The predicted targets of gma-miR171p, gma-miR166u, gma-miR2111f and gma-miR169c were experimentally validated. The mRNA cleavage sites were determined by modified 5' RNA ligase-mediated (5'-RLM)-RACE. The mRNA sequence of each complementary site and its 5' and 3' flanking sequences (5 nucleotides (nt) from 5' to 3') and cloned miRNA sequence (from 3' to 5') are shown. Vertical arrows indicate the 5' termini of the miRNA-guided cleavage products, as identified by 5'-RACE, with the frequency of clones shown.
Ijms 15 13596 g003

2.1.4. Correlation between the Expression of miRNAs and Their Targets in Response to Low Temperature Stress in Mature Nodules

If a gene acts as a functional target of a miRNA in regulating the cold response of nitrogen-fixing nodules, it should show an opposite pattern of expression to the miRNA in response to low temperatures. To this end, we analyzed the expression patterns of our validated miRNA targets in low temperature-treated nitrogen-fixing nodules using RT-qPCR. As shown in Figure 4, the expression levels of the target genes were negatively correlated with the corresponding miRNAs. Importantly, the expression of glyma07g01940, glyma06g21620, glyma16g06160 and glyma19g25770 was greatly repressed by cold treatment (Figure 4), in contrast to the corresponding miRNAs (Figure 2). The expression levels of gma-miR169c and glyma02g35190 were also negatively correlated (Figure 4). Interestingly, an exception was noted for the validated target genes, glyma11g17490 and glyma01g18040 (Figure 3). Of these, glyma11g17490 showed an opposite expression pattern to gma-miR171p, whereas glyma01g18040 was highly induced by cold treatment. These data suggest that glyma11g17490 is the direct target of gma-miR171p in the regulation of cold-induced nodular changes.
Figure 4. Expression levels of the validated target genes in low temperature-treated mature nodules. Total mRNA from cold-treated or untreated mature nodules was extracted, and RT-qPCR was used to analyze the relative expression of the selected target genes, including glyma07g01940 and glyma08g21620 targeted by gma-miR166u, glyma16g06160 and glyma19g25770 targeted by gma-miR2111f, glyma11g17490 and glyma01g18100 targeted by gma-miR171p and glyma02g35190 targeted by gma-miR169c. ELF1b was used as an internal control. The results given are averages ± SE of three biological replicates. Error bars indicate the standard deviation. Student’s t-tests were performed; statistically significant results are marked with * (p < 0.05); ** (p < 0.01); and *** (p < 0.001).
Figure 4. Expression levels of the validated target genes in low temperature-treated mature nodules. Total mRNA from cold-treated or untreated mature nodules was extracted, and RT-qPCR was used to analyze the relative expression of the selected target genes, including glyma07g01940 and glyma08g21620 targeted by gma-miR166u, glyma16g06160 and glyma19g25770 targeted by gma-miR2111f, glyma11g17490 and glyma01g18100 targeted by gma-miR171p and glyma02g35190 targeted by gma-miR169c. ELF1b was used as an internal control. The results given are averages ± SE of three biological replicates. Error bars indicate the standard deviation. Student’s t-tests were performed; statistically significant results are marked with * (p < 0.05); ** (p < 0.01); and *** (p < 0.001).
Ijms 15 13596 g004

2.2. Discussion

The recent genome-wide identification of miRNAs in low temperature-treated plants revealed the critical regulatory roles of miRNAs in plant responses to cold stress [5,6,7]. However, evidence for the miRNA-based regulation of cold responses in symbioses between plants and other organisms (e.g., bacteria) is lacking. It has been shown that miRNAs are involved in the establishment of legume-rhizobia symbioses, the functionality of mature nodules and the response to salt stress [22,45,46,47]. As a warm climate species, soybean exhibits high rates of nodulation and nitrogen fixation. The nitrogen fixation efficiency of soybean is substantially reduced at suboptimal temperatures [16,18]. Here, we demonstrated that the level of nitrogenase activity in functional nodules was decreased by more than 60% within 24 h of low temperature treatment (Figure 1D). Our results support the notion that low temperatures lead to a decreased metabolism in soybean-rhizobia symbioses and an immediate reduction in nitrogenase activity. Understanding the molecular mechanisms of temperature sensing and the signal transduction pathways in soybean-rhizobia symbioses, especially the nitrogen fixation function of nodules, will facilitate the improvement of the nitrogen fixation efficiency of soybean. In this study, we identified eleven low temperature-responsive miRNAs. We also validated the target genes for four miRNAs, and, more importantly, we identified the genes targeted by the miRNAs in nodules in response to cold. To our knowledge, this is the first report on the role of miRNAs in nodules and nitrogen fixation in response to low temperatures.
Among the eleven cold-responsive miRNAs we identified, four (miR166, miR171, miR2111 and miR169) are highly conserved in higher plants, and these miRNAs have been shown to regulate plant cold responses [8,10,11,48,49]. Interestingly, the expression patterns of these miRNAs vary in response to low temperatures in different plants. For example, miR169 is upregulated under cold stress in Arabidopsis, Brachypodium and rice [8,10,48,50]. By contrast, some specific members of the miR169 family are downregulated by cold treatment (e.g., miR169f/j/h/i in Populus and miR169d in Poncirus) [49,51]. Such findings indicate that these miRNAs or members of the miRNA family are involved in plant cold responses, but may have acquired diverse functions during evolution. Further, we found that miR169c expression was repressed by low-temperature treatment in nitrogen-fixing nodules (Figure 2B). Therefore, we propose that the conserved cold-responsive miRNAs and their mediated pathways play a key role in the cold responses of nitrogen-fixing nodules and soybean-rhizobia symbioses. Notably, previous studies have shown that miR166, miR171, miR2111 and miR397 are involved in nodulation in the legumes, Medicago and Lotus [52,53,54,55]. For example, Lotus miR171 and miR397 modulate rhizobial infection by targeting a transcription factor gene, Nodulation Signaling Pathway 2, and a gene encoding a laccase copper protein, respectively [53,56]. Thus, multiple miRNAs, including miR166, miR171, miR2111 and miR397, are required for nodule development, nitrogenase activity and nitrogen fixation efficiency control in nodules and in the adaptation of plants to adverse environmental conditions.
Our results also show that miRNAs are master regulators of the response of nodules to low temperatures. The miRNAs identified in this study target genes encoding transcription factors, enzymes and transporters that are involved in many biological processes, including the activation/repression of downstream genes, oxidation and reduction and the transportation of organic/inorganic molecules related to nitrogenase activity and nodule function (Table 1). miR166 is predicted to target homeodomain (HD)-ZIP transcription factors, and the target sites are conserved in Arabidopsis, rice, Brachypodium, maize and Populus [9,10,13,49,57,58]. In soybean, we found that Glyma08g21620 and Glyma07g01940 were cleaved by gma-miR166u (Figure 3), and both genes showed opposite expression patterns in response to cold (Figure 4). Glyma08g21620 and Glyma07g01940 encode a basic leucine zipper (bZIP) transcription factor and an HD-ZIP protein, respectively. Because bZIP and HD-ZIP transcription factors have diverse roles in plant development and stress responses [59,60,61], we speculate that gma-miR166u regulates the cold response of nitrogen-fixing nodules by modulating these transcription factors, allowing the signaling pathways to turn on/off. miR2111 has been reported to regulate plant responses to phosphate and nitrogen starvation by targeting Kelch domain-containing F-box proteins in Arabidopsis and soybean [62,63,64,65]. Here, we showed that miR2111f was also induced by low temperature treatment in mature nodules of soybean (Figure 2), and two Kelch-related genes (glyma19g25770 and glyma16g06160) were validated as targets of gma-miR2111f in cold-treated nodules (Figure 3 and Figure 4). These Kelch-related proteins share high homology with a Kelch domain-containing F-box protein that functions in the long-distance regulation of the Lotus-Rhizobium symbiosis (Takahara et al., 2013). It is likely that gma-miR2111 modulates soybean-rhizobia symbioses and stress adaptations. In addition, we found that gma-miR171p may regulate the cold response of nodules through the direct cleavage of a target gene (glyma11g17490) encoding a GRAS family transcription factor (Figure 4), although both predicted genes were confirmed to be targets of gma-miR171p (Figure 3). GRAS family transcription factors have been found to be essential for rhizobial Nod factor-induced transcription in rhizobial nodulation and mycorrhizal symbioses [66,67,68]. Moreover, five predicted target genes of miR169c were annotated as HAP2-like transcription factors (Table 1), and one of them (glyma02g35190) was validated as a direct target of gma-miR169c in cold-treated nodules (Figure 3 and Figure 4). In many plants, miR169 targets HAP2, which encodes a HAP2 transcription factor subunit and a component of the hetero-trimeric CCAAT-box-binding factor complex (CBF/NF-Y/HAP), which binds the CCAAT motif present in many eukaryotic promoters [69]. MtHAP2-1, which encodes a transcription factor belonging to the NFY-A family and which is targeted by miR169, has been shown to be required for nodule development [70]. However, the targeting of HAP2-like genes by miR169 has not been reported. It is likely that HAP2-like genes function as targets of miR169 to regulate cold responses (i.e., nitrogenase activity) in nodules.
Nitrogen fixation is a complex trait that is precisely and dynamically controlled by a complicated molecular network. Our results indicate that the transcription of cold-responsive miRNAs and their target genes may be controlled by cold-responsive transcription factors and transcriptional activators/repressors involved in auxin signaling (Figures S2 and S3). Our results provide novel insight into the regulatory networks underlying the nodular response to low temperatures. Functional analyses of miRNAs/targets and their signaling pathways, and of the molecular mechanisms of miRNA regulation in response to low temperatures, will help us to decipher the molecular mechanisms of nitrogen fixation efficiency control under low-temperature conditions. Because the majority of legumes are warm climate plants and nodulation and nitrogen fixation control are highly conserved in legumes, our research has important implications for the elucidation of nitrogen fixation efficiency control at low temperatures.

3. Experimental Section

3.1. Plant Materials and Low Temperature Treatment

Seeds of soybean (G. max cv. Williams 82) were surface-sterilized by treatment with 70% ethanol for 1 min, followed by 10% bleach for 4 min and then rinsed five times with sterile deionized water. The sterilized seeds were planted in a 6-inch pot filled with a 3:1 (v/v) mixture of sterilized vermiculite and perlite. The pots were pre-wetted with a low nitrogen plant nutrient solution and placed into a controlled growth chamber (24‒26 °C, 200 μmol·m−6·s−6, 16-h/8-h day/night photoperiod). The seedlings were watered every five days with the same low nitrogen solution. For rhizobium inoculation, B. japonicum strain USDA110 was grown in tryptone/yeast extract (TY) medium, centrifuged and resuspended in sterile deionized water to an optical density at 600 nm of 0.08. Five-day-old seedlings were flood-inoculated with the B. japonicum suspension culture (30 mL per plant).
For measurement of the Lb content and nitrogen use efficiency in mature nodules, seedlings inoculated with B. japonicum 28 days earlier were rinsed briefly in phosphate-buffered saline (PBS, pH 7.5) to remove the vermiculite and perlite particles and transferred to low nitrogen solution. The seedlings were chilled (CH) at 4 °C or left untreated (CK) for 24 h prior to analysis, and then, the nodules were collected. The harvested nodules were immediately frozen in liquid nitrogen for RNA extraction.

3.2. Analysis of the Leghemoglobin (Lb) Concentration

A pyridine-hemochrome assay was used to determine the Lb concentration [71]. Fresh nodules were collected and macerated with 2 mL of PBS. Nodular debris was discarded after filtration through two layers of cheesecloth. The filtrate from the last step was centrifuged at 10,000× g for 15 min. Next, 2 mL of the extract (suitably diluted) were mixed thoroughly with an equal volume of alkaline pyridine reagent. A few crystals of sodium were added to the greenish-yellow solution to reduce the hemochrome. The resulting solution was stirred without aeration, and the absorbance was read at 556 nm after 2 min against plain reagent as a blank using a NanoDrop 2000c spectrophotometer (Thermo Fisher Scientific, Waltham, MA, USA). The Lb concentration was estimated as described by Appleby and Bergersen [71].

3.3. Analysis of the Acetylene Reduction Activity (ARA) Level

After treatment, mature nodules (28 days post-inoculation (dpi)) were collected from the treated plants, and the ARA level was analyzed by gas chromatography, as described previously [36]. The mature nodules were collected in a 25-mL serum bottle sealed with an isobutyl rubber stopper. A syringe was used to replace 10% of the air in the bottle with the equivalent amount of C2H2, and the bottle was placed in the dark at 28 °C for 2 h. An Agilent gas chromatograph system (Agilent 7890 A, Agilent Technologies, Wilmington, DE, USA) was used to measure the ratio of C2H2 to ethane in the bottle (the parameters were as follows: KB Alumina capillary column, 30 m; column temperature, 80 °C; detector temperature, 200 °C; carrier gas, N2, air and H2; and purity of the standard gas (C2H2), >99.9%). The ARA of nitrogenase was calculated using the following formula [72]:
Ijms 15 13596 i001

3.4. Small RNA Library Construction, Solexa Sequencing and Data Analysis

For small RNA library construction and high-throughput sequencing, plants with mature nodules at 28 dpi were exposed to 4 °C for 24 h. Small RNAs were isolated from the treated (CH) and untreated nodules (CK) with a mirVana™-miRNA Isolation Kit (Ambion, Austin, TX, USA), followed by sequencing using an Illumina-Solexa 1 Genetic Analyzer at the Beijing Genomics Institute (Beijing, China). The read counts for miRNAs in the two libraries were normalized to one million (normalized expression = actual miRNA count/total count of clean reads × normalized one order of magnitude).
Data analysis was done as described by Dong [47]. Contaminating sequences, including adaptors, insert tags, sequences shorter than 18 nt or longer than 30 nt and those that could not be aligned with soybean genomic sequences were removed from the raw data. A BLAST analysis (E-value ≤ 0.01) was executed to exclude sequences matching exons and introns of mRNAs, rRNAs, tRNAs, scRNAs and snoRNAs in the NCBI GenBank database [73] and Rfam database [74]. The sequences filtered by the procedure described above were used for the following analysis.

3.5. Quantitative Reverse Transcription PCR (RT-qPCR)

Small RNAs were isolated from low temperature-treated or untreated soybean nodules using a mirVana™-miRNA Isolation Kit (Ambion) according to the manufacturer’s instructions. First-strand cDNA synthesis was performed using a miRcute miRNA first-strand cDNA synthesis kit with an oligo (dT) adaptor primer (5'-GCGAGCACAGAATTAATACGACTCACTATAGGT (A, G, or C) (A, G, C, or T)-3'). Quantitative real-time PCR was performed using an ABI PRISM 7500 Real-Time PCR System (ABI, Foster City, CA, USA) as follows: the reaction mixture (20 μL) contained 10 μL of SYBR Premix Ex Taq (Takara Bio Inc., Otsu, Japan), 2 μL of first-stand cDNA, 0.2 μM each primer, 0.4 μL of ROX Reference Dye II and 6.8 μL of ddH2O. The PCR mixtures were preheated at 95 °C for 30 s, followed by 40 cycles of amplification (95 °C for 5 s and 60 °C for 34 s). The forward primers used for the selected miRNAs and the reverse primer containing the adaptor sequence are included in Table S3. Our RT-qPCR data were analyzed using the software built into the 7500 system (ABI, Los Angeles, CA, USA). Each experiment contained three biological replicates.
For expression validation of the target genes, total RNA samples isolated using an RNase Plant Mini Kit (Tiangen Biotech, Beijing, China) were processed. Reverse transcription reactions were performed at 55 °C with 5 μg of RNA using SuperScript III Reverse Transcriptase (Invitrogen, Carlsbad, CA, USA). Oligo (dT) primers were used as reverse transcription primers according to the manufacturer’s instructions. The PCR primers used are included in Table S4. Real-time PCR was carried out using an ABI PRISM 7500 Real-Time PCR System (ABI), and SYBR Premix Ex-Taq (Takara Bio Inc.) was used to detect the PCR products. The reactions were incubated in a 96-well plate at 95 °C for 5 min, followed by 40 cycles of 95 °C for 5 s and 60 °C for 30 s. All reactions were run in triplicate.

3.6. miRNA Target Gene Prediction and Experimental Validation

Candidate miRNA target genes were determined using publicly available prediction algorithms, including psRNAtarget [43] and the target search tool in WMD3 [44]. Known G. max open reading frames downloaded from the soybean Genome Annotation Database (Phytozome) [75] were used for miRNA target prediction. To select putative miRNA-target pairs, only four mismatches between the mRNA targets and miRNAs were allowed [22].
For experimental validation of the putative target genes, total RNA was extracted from mature nodules and polyA mRNAs were isolated using a PolyATract® mRNA Isolation System III (Promega Corp., Fitchburg, WI, USA). PolyA mRNAs were then subjected to 5'-RACE assays using a GeneRacer Kit (Invitrogen), omitting the calf intestinal phosphatase and decapping reaction according to the manufacturer’s instructions. For each candidate target, two reverse gene-specific primers were designed. Nested PCR was performed to amplify the cleaved transcripts with reverse primers and GeneRacer adaptor-specific primers. The specific primers used for the putative targets are shown in Table S5. The PCR products were subsequently cloned into pMD19-T (Takara Bio Inc.) for sequencing.

4. Conclusions

SNF has been studied for a long time, but little is known about the molecular mechanism of SNF regulation under cold conditions. Sequencing analyses of soybean and other legumes have allowed researchers to investigate this issue. In this study, we showed the genome wide expression profiles of miRNAs in response to chilling in soybean nodules, described several miRNAs involved in the response of mature nodules to cold treatment and validated their target genes. miRNA-mediated gene regulation of the targets was confirmed to be part of this process in this tissue. Therefore, our results shed light on the roles of miRNAs in protection against chilling temperatures in soybean nodules. Functional analysis of miRNAs and their targets in response to low temperatures will help us to understand the molecular mechanisms by which nodules can cope with chilling injury and maintain a high nitrogen fixation rate in soybean under low temperatures.

Supplementary Files

Supplementary File 1

Acknowledgments

We are grateful to Wenxin Chen (China Agricultural University, Beijing, China) for kindly providing B. japonicum strain USDA110. This study was funded by the NSFC (31230050 and 30971797) and the Ministry of Agriculture of the People’s Republic of China (2009ZX08009-132B and 2014ZX0800929B). This work was also supported by the Youth Innovation Promotion Association, CAS.

Author Contributions

Senlei Zhang and Youning Wang designed the project and wrote the paper. Senlei Zhang, Youning Wang, Kexue Li and Liang Chen conducted the experiments. Kexue Li and Yanmin Zou were in charge of the statistical analysis. Xia Li and Youning Wang guided the conception and revision of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chinnusamy, V.; Zhu, J.; Zhu, J.K. Cold stress regulation of gene expression in plants. Trends Plant Sci. 2007, 12, 444–451. [Google Scholar] [CrossRef]
  2. Yamaguchi-Shinozaki, K.; Shinozaki, K. Transcriptional regulatory networks in cellular responses and tolerance to dehydration and cold stresses. Annu. Rev. Plant Biol. 2006, 57, 781–803. [Google Scholar] [CrossRef]
  3. Achard, P.; Gong, F.; Cheminant, S.; Alioua, M.; Hedden, P.; Genschik, P. The cold-inducible CBF1 factor-dependent signaling pathway modulates the accumulation of the growth-repressing DELLA proteins via its effect on gibberellin metabolism. Plant Cell 2008, 20, 2117–2129. [Google Scholar] [CrossRef]
  4. Chinnusamy, V.; Zhu, J.-K.; Sunkar, R. Gene regulation during cold stress acclimation in plants. Methods Mol. Biol. 2010, 639, 39–55. [Google Scholar] [CrossRef]
  5. Khraiwesh, B.; Zhu, J.; Zhu, J. Role of miRNAs and siRNAs in biotic and abiotic stress responses of plants. Biochim. Biophys. Acta 2012, 1819, 137–148. [Google Scholar] [CrossRef]
  6. Sunkar, R.; Li, Y.; Jagadeeswaran, G. Functions of microRNAs in plant stress responses. Cell 2012, 17, 196–203. [Google Scholar]
  7. Miura, K.; Furumoto, T. Cold signaling and cold response in plants. Int. J. Mol. Sci. 2013, 14, 5312–5337. [Google Scholar] [CrossRef]
  8. Zhou, X.; Wang, G.; Sutoh, K.; Zhu, J.; Zhang, W. Identification of cold-inducible microRNAs in plants by transcriptome analysis. Biochim. Biophys. Acta 2008, 1779, 780–788. [Google Scholar]
  9. Sunkar, R. Novel and stress-regulated microRNAs and other small rnas from Arabidopsis. Plant Cell 2004, 16, 2001–2019. [Google Scholar] [CrossRef]
  10. Zhang, J.; Xu, Y.; Huan, Q.; Chong, K. Deep sequencing of Brachypodium small RNAs at the global genome level identifies microRNAs involved in cold stress response. BMC Genomics 2009, 10, 449. [Google Scholar] [CrossRef]
  11. Tang, Z.; Zhang, L.; Xu, C.; Yuan, S.; Zhang, F.; Zheng, Y.; Zhao, C. Uncovering small RNA-mediated responses to cold stress in a wheat thermosensitive genic male-sterile line by deep sequencing. Plant Physiol. 2012, 159, 721–738. [Google Scholar] [CrossRef]
  12. Thiebaut, F.; Rojas, C.A.; Almeida, K.L.; Grativol, C.; Domiciano, G.C.; Lamb, C.R.C.; de Almeida Engler, J.; Hemerly, A.S.; Ferreira, P.C. Regulation of miR319 during cold stress in sugarcane. Plant Cell Environ. 2012, 35, 502–512. [Google Scholar] [CrossRef]
  13. Lu, S.; Sun, Y.-H.; Chiang, V.L. Stress-responsive microRNAs in Populus. Plant J. 2008, 55, 131–151. [Google Scholar] [CrossRef]
  14. Keyser, H.; Li, F. Potential for increasing biological nitrogen fixation in soybean. In Biological Nitrogen Fixation for Sustainable Agriculture; Ladha, J.K., George, T., Bohlool, B.B., Eds.; Springer: Berlin, Germany, 1992; pp. 119–135. [Google Scholar]
  15. Duke, S.; Scharader, L.; Miller, M. Low temperature effects on soybean (Glycine max) mitochondrial respiration and several dehydrogenases during imbibition and germination. Plant Physiol. 1977, 60, 716–722. [Google Scholar] [CrossRef]
  16. Walsh, K.; Layzell, D. Carbon and nitrogen assimilation and partitioning in soybeans exposed to low root temperatures. Plant Physiol. 1986, 80, 249–255. [Google Scholar] [CrossRef]
  17. Matamoros, M.A.; Baird, L.M.; Escuredo, P.R.; Dalton, D.A.; Minchin, F.R.; Iturbe-Ormaetxe, I.; Rubio, M.C.; Moran, J.F.; Gordon, A.J.; Becana, M. Stress-induced legume root nodule senescence. Physiological, biochemical, and structural alterations. Plant Physiol. 1999, 121, 97–112. [Google Scholar] [CrossRef]
  18. Van Heerden, P.D.R.; Kiddle, G.; Pellny, T.K.; Mokwala, P.W.; Jordaan, A.; Strauss, A.J.; de Beer, M.; Schluter, U.; Kunert, K.J.; Foyer, C.H. Regulation of respiration and the oxygen diffusion barrier in soybean protect symbiotic nitrogen fixation from chilling-induced inhibition and shoots from premature senescence. Plant Physiol. 2008, 148, 316–327. [Google Scholar] [CrossRef]
  19. Allen, D.J.; Ort, D.R. Impacts of chilling temperatures on photosynthesis in warm-climate plants. Trends Plant Sci. 2001, 6, 36–42. [Google Scholar] [CrossRef]
  20. Kingston-Smith, A.H.; Harbinson, J.; Williams, J.; Foyer, C.H. Effect of chilling on carbon assimilation, enzyme activation, and photosynthetic electron transport in the absence of photoinhibition in maize leaves. Plant Physiol. 1997, 114, 1039–1046. [Google Scholar]
  21. 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]
  22. 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]
  23. 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]
  24. Zhang, B.; Pan, X.; Stellwag, E.J. Identification of soybean microRNAs and their targets. Planta 2008, 229, 161–182. [Google Scholar] [CrossRef]
  25. Wong, C.E.; Zhao, Y.T.; Wang, X.J.; Croft, L.; Wang, Z.H.; Haerizadeh, F.; Mattick, J.S.; Singh, M.B.; Carroll, B.J.; Bhalla, P.L. MicroRNAs in the shoot apical meristem of soybean. J. Exp. Bot. 2011, 62, 2495–2506. [Google Scholar] [CrossRef]
  26. Kulcheski, F.R.; de Oliveira, L.F.V.; 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 Genomics 2011, 12, 307. [Google Scholar] [CrossRef] [Green Version]
  27. Pfeiffer, N.E.; Torres, C.M.; Wagner, F.W. Proteolytic activity in soybean root nodules activity in host cell cytosol and bacteroids throughout physiological development and senescence. Plant Physiol. 1983, 71, 797–802. [Google Scholar] [CrossRef]
  28. Chen, P.-C.; Phillips, D.A. Induction of root nodule senescence by combined nitrogen in Pisum sativum L. Plant Physiol. 1977, 59, 440–442. [Google Scholar] [CrossRef]
  29. Delgado, M.J.; Garrido, J.M.; Ligero, F.; Lluch, C. Nitrogen fixation and carbon metabolism by nodules and bacteroids of pea plants under sodium chloride stress. Physiol. Plant. 1993, 89, 824–829. [Google Scholar] [CrossRef]
  30. Delgado, M.J.; Ligero, F.; Lluch, C. Effects of salt stress on growth and nitrogen fixation by pea, faba-bean, common bean and soybean plants. Soil Biol. Biochem. 1994, 26, 371–376. [Google Scholar] [CrossRef]
  31. Balestrasse, K.B.; Gardey, L.; Gallego, S.M.; Tomaro, M. Response of antioxidant defence system in soybean nodules and roots subjected to cadmium stress. Funct. Plant Biol. 2001, 28, 497–504. [Google Scholar] [CrossRef]
  32. López, M.; Herrera-Cervera, J.A.; Iribarne, C.; Tejera, N.A.; Lluch, C. Growth and nitrogen fixation in Lotus japonicus and Medicago. truncatula under NaCl stress: Nodule carbon metabolism. J. Plant Physiol. 2008, 165, 641–650. [Google Scholar] [CrossRef]
  33. Sánchez-Pardo, B.; Carpena, R.O.; Zornoza, P. Cadmium in white lupin nodules: Impact on nitrogen and carbon metabolism. J. Plant Physiol. 2013, 170, 265–271. [Google Scholar] [CrossRef]
  34. Navascués, J.; Pérez-Rontomé, C.; Gay, M.; Marcos, M.; Yang, F.; Walker, F.A.; Desbois, A.; Abián, J.; Becana, M. Leghemoglobin green derivatives with nitrated hemes evidence production of highly reactive nitrogen species during aging of legume nodules. Proc. Natl. Acad. Sci. USA 2012, 109, 2660–2665. [Google Scholar] [CrossRef] [Green Version]
  35. Vorster, B.J.; Schlüter, U.; du Plessis, M.; van Wyk, S.; Makgopa, M.E.; Ncube, I.; Quain, M.D.; Kunert, K.; Foyer, C.H. The cysteine protease–cysteine protease inhibitor system explored in soybean nodule development. Agronomy 2013, 3, 550–570. [Google Scholar] [CrossRef]
  36. Boyd, E.S.; Lange, R.K.; Mitchell, A.C.; Havig, J.R.; Hamilton, T.L.; Lafrenière, M.J.; Shock, E.L.; Peters, J.W.; Skidmore, M. Diversity, abundance, and potential activity of nitrifying and nitrate-reducing microbial assemblages in a subglacial ecosystem. Appl. Environ. Microb. 2011, 77, 4778–4787. [Google Scholar] [CrossRef]
  37. Fenta, B.A.; Driscoll, S.P.; Kunert, K.J.; Foyer, C.H. Characterization of drought-tolerance traits in nodulated soya beans: The importance of maintaining photosynthesis and shoot biomass under drought-induced limitations on nitrogen metabolism. J. Agron. Crop. Sci. 2012, 198, 92–103. [Google Scholar] [CrossRef]
  38. Furlan, A.; Bianucci, E.; del Carmen, T.M.; Castro, S.; Dietz, K. Antioxidant enzyme activities and gene expression patterns in peanut nodules during a drought and rehydration cycle. Funct. Plant Biol. 2014, 41, 81–86. [Google Scholar]
  39. Hungria, M.; Kaschuk, G. Regulation of N2 fixation and NO3/NH4+ assimilation in nodulated and N-fertilized Phaseolus. vulgaris L. exposed to high temperature stress. Environ. Exp. Bot. 2014, 98, 32–39. [Google Scholar] [CrossRef]
  40. De Costa, W.; Becher, M.; Schubert, S. Effects of water stress on nitrogen fixation of common bean (Phaseolus. vulgaris L.). J. Nat. Sci. Found. Sri Lanka 2012, 25, 83–94. [Google Scholar]
  41. Palma, F.; Tejera, N.A.; Lluch, C. Nodule carbohydrate metabolism and polyols involvement in the response of Medicago sativa to salt stress. Environ. Exp. Bot. 2013, 85, 43–49. [Google Scholar] [CrossRef]
  42. Plantpan. Available online: http://plantpan.mbc.nctu.edu.tw/ (accessed on 17 July 2012).
  43. Dai, X.; Zhao, P. psRNATarget: A plant small RNA target analysis server. Nucleic Acids Res. 2011, 39, 155–159. [Google Scholar] [CrossRef]
  44. Ossowski, S.; Schwab, R.; Weigel, D. Gene silencing in plants using artificial microRNAs and other small RNAs. Plant J. 2008, 53, 674–690. [Google Scholar] [CrossRef]
  45. Simon, S.A.; Meyers, B.C.; Sherrier, D.J. MicroRNAs in the rhizobia legume symbiosis. Plant Physiol. 2009, 151, 1002–1008. [Google Scholar] [CrossRef]
  46. Devers, E.; 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]
  47. Dong, Z.; Shi, L.; Wang, Y.; Chen, L.; Cai, Z.; Wang, Y.; Jin, J.; Li, X. Identification and dynamic regulation of microRNAs involved in salt stress responses in functional soybean nodules by high-throughput sequencing. Int. J. Mol. Sci. 2013, 14, 2717–2738. [Google Scholar]
  48. Lv, D.-K.; Bai, X.; Li, Y.; Ding, X.-D.; Ge, Y.; Cai, H.; Ji, W.; Wu, N.; Zhu, Y.-M. Profiling of cold-stress-responsive miRNAs in rice by microarrays. Gene 2010, 459, 39–47. [Google Scholar] [CrossRef]
  49. Chen, L.; Zhang, Y.; Ren, Y.; Xu, J.; Zhang, Z.; Wang, Y. Genome-wide identification of cold-responsive and new microRNAs in Populus tomentosa by high-throughput sequencing. Biochem. Biophys. Res. Commun. 2012, 417, 892–896. [Google Scholar] [CrossRef]
  50. Li, Y.; Fu, Y.; Ji, L.; Wu, C.; Zheng, C. Characterization and expression analysis of the Arabidopsis mir169 family. Plant Sci. 2010, 178, 271–280. [Google Scholar] [CrossRef]
  51. Zhang, X.-N.; Li, X.; Liu, J.-H. Identification of conserved and novel cold-responsive microRNAs in trifoliate orange (Poncirus. trifoliata (L.) Raf.) using high-throughput sequencing. Plant Mol. Biol. Rep. 2014, 32, 328–341. [Google Scholar] [CrossRef]
  52. Boualem, A.; Laporte, P.; Jovanovic, 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]
  53. Høgslund, N.; Radutoiu, S.; Krusell, L.; Voroshilova, V.; Hannah, M.A.; Goffard, N.; Sanchez, D.H.; Lippold, F.; Ott, T.; Sato, S. Dissection of symbiosis and organ development by integrated transcriptome analysis of Lotus japonicus mutant and wild-type plants. PLoS One 2009, 4, e6556. [Google Scholar] [CrossRef]
  54. Lelandais-Briere, 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]
  55. Takahara, M.; Magori, S.; Soyano, T.; Okamoto, S.; Yoshida, C.; Yano, K.; Sato, S.; Tabata, S.; Yamaguchi, K.; Shigenobu, S. TOO MUCH LOVE, a Novel Kelch repeat-containing F-box protein, functions in the long-distance regulation of the legume–rhizobium symbiosis. Plant Cell Physiol. 2013, 54, 433–447. [Google Scholar]
  56. 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]
  57. Ding, Y.; Chen, Z.; Zhu, C. Microarray-based analysis of cadmium-responsive microRNAs in rice (Oryza sativa). J. Exp. Bot. 2011, 62, 3563–3573. [Google Scholar] [CrossRef]
  58. Ding, D.; Li, W.; Han, M.; Wang, Y.; Fu, Z.; Wang, B.; Tang, J. Identification and characterisation of maize microRNAs involved in developing ears. Plant Biol. 2014, 16, 9–15. [Google Scholar] [CrossRef]
  59. Uno, Y. Arabidopsis basic leucine zipper transcription factors involved in an abscisic acid-dependent signal transduction pathway under drought and high-salinity conditions. Proc. Natl. Acad. Sci. USA 2000, 97, 11632–11637. [Google Scholar] [CrossRef]
  60. Shin, D.; Koo, Y.D.; Lee, J.; Lee, H.-J.; Baek, D.; Lee, S.; Cheon, C.-I.; Kwak, S.-S.; Lee, S.Y.; Yun, D.-J. Athb-12, a homeobox-leucine zipper domain protein from Arabidopsis thaliana, increases salt tolerance in yeast by regulating sodium exclusion. Biochem. Biophys. Res. Commun. 2004, 323, 534–540. [Google Scholar] [CrossRef]
  61. Rodriguez-Uribe, L. A root-specific bZIP transcription factor is responsive to water deficit stress in tepary bean (Phaseolus. acutifolius) and common bean (P. vulgaris). J. Exp. Bot. 2006, 57, 1391–1398. [Google Scholar] [CrossRef]
  62. Kuo, H.-F.; Chiou, T.-J. The role of microRNAs in phosphorus deficiency signaling. Plant Physiol. 2011, 156, 1016–1024. [Google Scholar] [CrossRef]
  63. Liang, G.; He, H.; Yu, D. Identification of nitrogen starvation-responsive microRNAs in Arabidopsis thaliana. PLoS One 2012, 7, e48951. [Google Scholar] [CrossRef]
  64. 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]
  65. Xu, F.; Liu, Q.; Chen, L.; Kuang, J.; Walk, T.; Wang, J.; Liao, H. Genome-wide identification of soybean microRNAs and their targets reveals their organ-specificity and responses to phosphate starvation. BMC Genomics 2013, 14, 66. [Google Scholar]
  66. Smit, P.; Raedts, J.; Portyanko, V.; Debellé, F.; Gough, C.; Bisseling, T.; Geurts, R. NSP1 of the GRAS protein family is essential for rhizobial Nod factor-induced transcription. Science 2005, 308, 1789–1791. [Google Scholar] [CrossRef]
  67. Murakami, Y.; Miwa, H.; Imaizumi-Anraku, H.; Kouchi, H.; Downie, J.A.; Kawaguchi, M.; Kawasaki, S. Positional cloning identifies Lotus japonicus NSP2, a putative transcription factor of the GRAS family, required for NIN and ENOD40 gene expression in nodule initiation. DNA Res. 2007, 13, 255–265. [Google Scholar]
  68. Gobbato, E.; Marsh, J.F.; Vernié, T.; Wang, E.; Maillet, F.; Kim, J.; Miller, J.B.; Sun, J.; Bano, S.A.; Ratet, P. A GRAS-type transcription factor with a specific function in mycorrhizal signaling. Curr. Biol. 2012, 22, 2236–2241. [Google Scholar] [CrossRef]
  69. Maity, S.N.; de Crombrugghe, B. Role of the CCAAT-binding protein CBF/NF-Y in transcription. Trends Biochem. Sci. 1998, 23, 174–178. [Google Scholar] [CrossRef]
  70. Combier, J.P.; Frugier, F.; de Billy, F.; Boualem, A.; El-Yahyaoui, F.; Moreau, S.; Vernie, 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. Gene. Dev. 2006, 20, 3084–3088. [Google Scholar] [CrossRef]
  71. Appleby, C.; Bergersen, F. Preparation and experimental use of leghaemoglobin. In Methods for Evaluating Biological Nitrogen Fixation; Bergersen, F.J., Ed.; John Wiley & Sons: Chichester, UK, 1980; pp. 315–335. [Google Scholar]
  72. Shearer, G.; Kohl, D.H. N2-fixation in field settings: Estimations based on natural 15N abundance. Funct. Plant Biol. 1986, 13, 699–756. [Google Scholar]
  73. NCBI Genbank Database. Available online: http://www.ncbi.nlm.nih.gov/blast/Blast.cgi (accessed on 15 July 2012).
  74. Sanger. Available online: http://www.sanger.ac.uk/Software/Rfam (accessed on 15 July 2012).
  75. Phytozome. Available online: http://www.phytozome.net/ (accessed on 20 July 2012).

Share and Cite

MDPI and ACS Style

Zhang, S.; Wang, Y.; Li, K.; Zou, Y.; Chen, L.; Li, X. Identification of Cold-Responsive miRNAs and Their Target Genes in Nitrogen-Fixing Nodules of Soybean. Int. J. Mol. Sci. 2014, 15, 13596-13614. https://doi.org/10.3390/ijms150813596

AMA Style

Zhang S, Wang Y, Li K, Zou Y, Chen L, Li X. Identification of Cold-Responsive miRNAs and Their Target Genes in Nitrogen-Fixing Nodules of Soybean. International Journal of Molecular Sciences. 2014; 15(8):13596-13614. https://doi.org/10.3390/ijms150813596

Chicago/Turabian Style

Zhang, Senlei, Youning Wang, Kexue Li, Yanmin Zou, Liang Chen, and Xia Li. 2014. "Identification of Cold-Responsive miRNAs and Their Target Genes in Nitrogen-Fixing Nodules of Soybean" International Journal of Molecular Sciences 15, no. 8: 13596-13614. https://doi.org/10.3390/ijms150813596

Article Metrics

Back to TopTop