Introduction

 

With the coming of the post-genome research era, more attention has been paid to the strategy of model biology research. The model plants A. thaliana and rice have been widely used in various research fields of botany. However, A. thaliana, which is a dicotyledon, has no biological characteristics and related agronomic traits of cereal crops, such as tillering and panicle differentiation, and is far away from monocotyledons (Blümel et al. 2015). Rice (Oryza sativa), whose genome sequencing has been completed early, is a model plant in Gramineae, but it has evolved very early from the cold-season Gramineae, and is far away from temperate crops.

Consequently, a novel model plant is needed to study some of the problems associated with gramineous plants. B. distachyon is a subfamily of Poaceae, which is closely related to many important cereal crops such as wheat, barley, and rice (Brutnell et al. 2015). In addition, compared to rice, whose genome is around 430 Mb (Sasaki and Antonio 2004), the genome of B. distachyon is just approximate 272 Mb (Initiative 2010). B. distachyon, as a model plant for grass crops, according to its biological characteristics of short life cycle, and easy to cultivate and genetically transform, could offer a suitable and meaningful foundation for many aspects of plant biology research (Scholthof et al. 2018).

Drought is one of the abiotic stresses that seriously affect plant growth and development, and is also the main limiting factor for global crop distribution and yield increase (Farooq et al. 2009; Matiu et al. 2017). Plants have evolved sophisticated mechanisms to respond to and adapt to drought stress at the molecular and physiological levels. However, previous molecular studies about stress response mainly concentrated on genes with protein-coding function (Sakuma et al. 2006; Huang et al. 2008). Recently, non-protein-encoding transcripts have also proven to be important regulators of gene expression (An et al. 2018; Nejat and Mantri 2018). Long non-coding RNA (lncRNA) is a kind of transcript with a length of more than 200 nucleotide (nt), lacking coding sequence (CDS) or open reading frame (ORF), and no or almost no protein coding potential (Do et al. 2019). Increasing evidences show that plant lncRNAs are critical regulators for growth and development as well as response to stresses. LncRNAs could regulate the flowering time (Heo and Sung 2011), pollen development (Huang et al. 2018), fruit senescence (Wang et al. 2018), photomorphogenesis (Wang et al. 2014b), as well as biotic and abiotic stress responses (Nejat and Mantri 2018). Functional analyses of lncRNAs suggest that they take part in gene expression regulation not only at transcriptional level, but also at post-transcriptional level through many complicated mechanisms, such as precursors for microRNAs (miRNAs) and other small RNAs or as target mimetics of miRNAs (Qin et al. 2017). At present, most studies on drought stress response in the main gramineous crops still focused on physiological levels or protein-coding functional genes, however, little is known about how non-coding transcripts exert an influence on the response and regulation of drought stress.

In this research, B. distachyon, the optimal model organism for gramineous crops, was taken as experimental material, and the high-throughput sequencing technology was employed to sequence the transcriptome of drought stress and normal growth plants, respectively. The drought-responsive lncRNAs were identified through the comparative transcriptome analyses to provide a theoretical foundation and new insight for understanding and improving the molecular regulation mechanism of drought response of gramineous crops.

 

Materials and Methods

 

Plant materials, growth conditions and drought stress treatment

 

Cultivar of B. distachyon Bd-21 was used as the experimental material. After removing the inferior palea, seeds were placed in a clean petri dish covered with moist filter paper to germinate. One day later, the petri dish was transferred to a refrigerator at 4°C for 8 days. Then the seeds were planted in nutrient pots in a light incubator, with growth conditions: 25°C, light for 16 h / dark for 8 h. When the seedlings grew to the 7th leaf stage, the drought treatment was carried out by means of water control. The control was watered once every other day, while the treated group grew seven days without watering. Then the leaves of the seedlings were sampled and frozen by the liquid nitrogen, storing in an ultra-low temperature refrigerator at -80°C for following testing.

RNA isolation and qualification

 

The methods of total RNA isolation and reverse transcription refer to the description of Ma et al. (2019). The degradation and contamination of total RNA were detected by 1% agarose gels. The purity, concentration and integrity of RNA were checked by the Nano Photometer® spectrophotometer (IMPLEN, CA, USA) and Qubit® RNA Assay Kit in Qubit® 2.0 Fluorometer (Life Technologies, CA, USA), and RNA Nano 6000 Assay Kit of the Bioanalyzer 2100 system (Agilent Technologies, CA, USA), respectively.

 

Library construction, sequencing and clustering

 

The construction of cDNA library, sequencing and clustering were conducted by the Novogene Bioinformatics Technology Co., Ltd. in Beijing, following the procedures of relevant kits, systems and machines.

 

Data processing and bioinformatics analysis

 

The data from Illumina sequencing were processed further and analyzed by the Novogene Bioinformatics Technology Co., Ltd.

(1)      Quality control (QC) was checked by raw data, clean data, Q20, Q30 and GC content of the clean data. Raw data (raw reads) of fastq format were firstly processed through in-house perl scripts. In this step, clean data (clean reads) were obtained by removing reads containing adapter, reads on containing ploy-N and low quality reads from raw data. At the same time, Q20, Q30 and GC content of the clean data were calculated. All the downstream analyses based on the clean data with high quality. The correlation of gene expression levels between samples is based on the square of the Pearson correlation coefficient (R2) > 0.92 (ideal sampling and test conditions), and R2 should be at least greater than 0.8 in a specific test operation.

(2)      Reference genome and gene model annotation files were downloaded from genome website (http://www.ncbi.nlm.nih.gov/nuccore/288572679) directly. Index of the reference genome was built using Bowtie v2.0.6 and paired-end clean reads were aligned to the reference genome using TopHat v2.0.9.

(3)      The mapped reads of each sample were assembled by Cufflinks (v2.1.1) (Fig. S1) (Trapnell et al. 2010). According to the assembly results of cufflinks, the structural features of lncRNA as well as the functional qualities of non-encoded proteins, a series of severe screening criteria were established. Then these lncRNAs could be selected and taken as the final candidate lncRNA set in terms of the following five steps (Fig. S2).

(4)      Coding potential analyses included Coding Potential Calculator (CPC) as well as Pfam-scan. The clarification of both coding and non-coding transcripts is mainly based on CPC (0.9-r2) to assess the extent and quality of ORFs in transcripts and search known protein sequence databases (Kong et al. 2007). Each transcript was translated in all three possible frames and used Pfam Scan (v1.3) to identify occurrence of any of the known protein family domains documented in the Pfam database (Punta et al. 2012). If a transcript has a Pfam hit, it would be removed in the following steps.

(5)      The function of lncRNA could be mainly achieved by cis or trans acting on the target gene with protein-encoding function (trans analysis is performed when the number of samples is more than 5). The basic principle of cis acting target gene prediction is that the function of lncRNA is related to the protein-coding gene adjacent to its coordinates, so the adjacent (upstream and downstream 10 k/100k) protein-coding genes are screened as their target genes. The main functions of lncRNA were predicted by functional enrichment analysis of target genes.

(6)      Quantification of gene expression level was carried out via the Cuffdiff (v2.1.1), which was used to calculate FPKMs of both lncRNAs and coding genes in each sample (Trapnell et al. 2010). FPKM means fragments per kilo-base of exon per million fragments mapped, calculated based on the length of the fragments and reads count mapped to this fragment. The threshold of differential expression varies from sample to sample. The transcript or gene with a P-adjust < 0.05 is the differential expression threshold for biological replicates, while the P-adjust <0.05 and |log2 (fold change)| > 1 is a significant difference threshold for non-biological replicates.

(7)      Gene Ontology (GO) enrichment analysis of genes with differential expression or lncRNA target genes were performed through the GO seq R package, in which gene length bias could be corrected. The corrected P value of GO term which was less than 0.05 would be considered significantly enriched by differential expressed genes. Kyoto Encyclopedia of Genes and Genomes (KEGG) is a database resource to recognize functions of the biological system (http://www.genome.jp/kegg/). The KOBAS software was used to detect the statistical enrichment of differentially expressed genes or target genes of lncRNA in KEGG pathways.

(8)      Alternative splicing (AS) events were divided into 12 basic forms via the software Asprofile v1.0 (Florea et al. 2013). The amount of AS events in each sample was estimated, respectively.

 

qRT-PCR validation

 

6 lncRNAs and 6 mRNAs were chosen to detect the expression level via qRT-PCR. The reaction system and procedures according to the description of Feng et al. (2015). The fold variation was calculated by 2-△△Ct method (Livak and Schmittgen 2001). The β-actin was taken as internal control.

Results

 

Evaluation of sequencing quality

 

The Illumina HiSeqTM2500 was performed in order to obtain the drought responsive lncRNAs in B. distachyon. In this experiment, 128,474,682 and 116,766,272 raw reads were acquired in CK and drought treatment, respectively. The original sequencing sequences produced by sequencing contain low-quality reads and some adapter. The raw reads must be filtered in order to attain clean reads to assure the quality of bioinformatic analysis, because clean reads are the basis of subsequent analysis. The clean reads acquired after filtering were 122,429,492 and 111,457,314, separately (Table S1).

The correlation of gene expression levels between samples is an essential index to verify the reliability of the experiment and the rationality of sample selection. The closer the correlation coefficient is to 1, the higher the similarity in expression patterns between samples (Fig. S3).

 

Assembly and screening of lncRNA

 

Cufflinks (Trapnell et al. 2010) used the results of the TopHat alignment to assemble transcripts, to obtain as small a collection of transcripts as possible, and to quantify transcripts. Based on the splicing results of cufflinks, the candidate lncRNAs were determined by a 5-step screening method. Fig. 1 shows the statistics of the number of transcripts produced by step1-step5 during the basic screening process of lncRNAs.

The candidate lncRNAs were analyzed by CPC and Pfam-scan, and the non-coding transcripts identified by the above analyses were counted. Fig. 2 visually displays the common and unique numbers of the predicted non-coding transcripts. The number of non-coding transcripts identified by CPC and Pfam-scan was 28,956 and 29,550, respectively. The common part of these prediction results (28,436) was taken as the lncRNAs data set predicted by this analysis for subsequent analysis.

 

Alignment analyses of expression level

 

By comparing the box plot (Fig. 3A) and density profile (Fig. 3B) of the quantitative results between CK and drought, the differences in the overall expression level of the samples under different treatments were visually reflected. Currently, among the RNA-seq technologies, FPKM is the most popular measure for evaluating gene expression levels (Trapnell et al. 2010). The FPKM values obtained by cuffdiff quantification could be used to compare the expression levels of mRNA and lncRNA. The results showed no molecular type preference (Table S2). The FPKM values of lncRNA and mRNA are shown in Table S3.

 

The expression values of lncRNA and mRNA transcripts of all samples were averaged, and the violin plot (Fig. 4) was drawn with the log10 (RPKM+1) values to visually reflect the difference in the whole expression level of lncRNA and mRNA. The results indicated that the expression level of lncRNA was lower than that of mRNA. Differential transcript analysis was performed using cuffdiff.

Using q-value<0.05 as the threshold of screening, the volcano plot can be used to visually see the whole distribution of differential transcripts or genes (Fig. 5). The red dot and blue dot in the figure indicate the up-regulated (2,497) and down-regulated (973) transcripts with significant differential expression, respectively (Fig. 5).

 

Functional prediction of lncRNA target gene

 

The (upstream and downstream 100 k) protein-coding genes adjoined lncRNAs were screened as their target genes, and the main functions of lncRNAs were predicted by functional enrichment analysis of target genes. The predicted results of cis target genes are shown in Table S4. The GO classification plot (Fig. 6A and Fig. 7A) represented differentially expresses lncRNA target genes and mRNAs, which intuitively reflected the number distribution of differential genes in GO term enriched in biological processes, cellular components, and molecular functions.

 

Fig. 1: Screening statistics of lncRNA. The x-axis is the screening step, and the y-axis is the number of transcripts after the corresponding step is filtered

 

 

Fig. 2: Venn diagram shows number of lncRNAs of predicted non-coding transcripts

 

 

Fig. 3: Comparison of gene expression levels under different treatments. A, FPKM box plot, the x-axis is the sample name, the y-axis is log10 (FPKM+1), and the box plot for each region corresponds to five statistics (top to bottom, maximum, upper quartile, median, lower quartile and minimum); B, FPKM density distribution profile, the x-axis is log10 (FPKM+1), and the y-axis is the density of the gene

 

In vivo, distinct genes collaborate with each other to exert their biological functions. Pathway significant enrichment analysis e mploys a hypergeometric test to ensure pathway which was significantly enriched in specific genes compared to the whole genome background. The scatter plots represent the results of the KEGG enrichment analyses (Fig. 6B and Fig. 7B). The degree of KEGG enrichment is measured by the rich factor, the q-value, and the number of genes enriched in this pathway. Rich factor refers to the ratio of the number of genes in the pathway entry in the lncRNA target genes with differential expression to the total number of genes in the pathway. The larger the rich factor, the higher the degree of enrichment. Q-value is the p-value after multiple hypothesis test correction. The value of q-value is [0, 1], the closer to zero, the more significant the enrichment. We selected the 20 most significant pathway entries for enrichment here (Fig. 6B and Fig. 7B). According to the results of KEGG enrichment analyses, clustering analysis was performed on the significantly enriched KEGG pathway. In the analysis process, we selected the most significant 20 metabolic pathways to analyze the difference in expression (Fig. 6C and Fig. 7C). The expression of lncRNA in the photosynthesis-antenna proteins pathway in CK was higher than that in drought treatment, while the expression of lncRNA in Glycolysis/Gluconeogenesis, Pentose and glucuronate interconversions, Glycerolipid metabolism pathways was lower than that in drought treatment. The mRNA expression levels in CK were higher than drought treatment in photosynthesis-antenna proteins, metabolic pathways, carbon fixation in photosynthetic organisms, and lower than drought treatment in pyruvate metabolism, ascorbate and aldarate metabolism, degradation of aromatic compounds pathways.

The response of plants to drought stress is a complex regulatory system including multiple genes and

 

 

Fig. 4: Comparison of lncRNA and mRNA expression levels under drought treatment. A, violin plot of lncRNA and mRNA expression, the x-axis is the gene name, the y-axis is log10 (FPKM+1), the width of each violin represents the number of points at that expression level; B, FPKM density distribution profile of lncRNA and mRNA, the x-axis is log10 (FPKM+1), and the y-axis is the density of the gene

 

 

Fig. 5: Volcano plot of differential expression transcript under drought treatment. Significantly differentially expressed transcripts are represented by red dots (upregulated expression transcripts) and green dots (downregulated expression of transcripts); x-axis represents fold change in transcripts in different samples; y-axis represents statistical significance of differences in transcript expression changes

 

regulatory networks. According to the functional annotations of predicted target genes, we divided them into five categories, including three photosynthesis-related genes, two transcription factors, one peptide, five ribosomal proteins, and one glycoside hydrolase (Table S5). A total of 90 lncRNAs with differential expression were involved. Except that the transcription factor wox11 was up-regulated under drought treatment, the remaining target genes were down-regulated (Table S5).

The regulation mechanism of lncRNA on its target gene is very complicated. The results of this study showed that lncRNA regulated its target genes mainly in two ways, one-to-many and many-to-one. The former refers to that one lncRNA could regulate multiple target genes (Table S6), while the latter, in contrast, means that one target gene could be regulated by multiple lncRNAs (Table S7).

 

AS analysis

 

The same gene may have different AS patterns, which greatly increase the ability to encode genes. Different AS patterns allow the same gene to produce multiple different mature mRNAs, ultimately producing different proteins. AS is an important pattern by which lncRNA regulates gene expression. The lncRNA could form a complementary double strand with the transcript of protein-encoding gene, which interferes with the splicing of mRNA and then could produce different AS events (Bardou et al. 2014; Gonzalez et al. 2015). The gene model predicted by Cufflinks (Trapnell et al. 2010) was employed to classify and quantify the AS events of each sample using AS profile (Florea et al. 2013) software. 12 categories of AS events include (1) TSS: Alternative 5’ first exon (transcription start site), (2) TTS: Alternative 3’ last exon (transcription terminal site), (3) SKIP: Skipped exon (SKIP_ON,SKIP_OFF pair), (4) XSKIP: Approximate SKIP (XSKIP_ON,XSKIP_OFF pair), (5) MSKIP: Multi-exon SKIP (MSKIP_ON,MSKIP_OFF pair), (6) XMSKIP: Approximate MSKIP (XMSKIP_ON,XMSKIP_OFF pair), (7) IR: Intron retention (IR_ON, IR_OFF pair), (8) XIR: Approximate IR (XIR_ON, XIR_OFF pair), (9) MIR: Multi-IR (MIR_ON, MIR_OFF pair), (10) XMIR: Approximate MIR (XMIR_ON, XMIR_OFF pair), (11) AE: Alternative exon ends (5’, 3’, or both), (12) XAE: Approximate AE. The number distribution of AS is similar in the two treatments, with the largest number of TSS and TTS and the least number of XMSKIPs (Fig. S4).

 

 

Fig. 6: Functional enrichment of differentially expressed lncRNA target gene under drought treatment. A, GO classification plot; B, scatter plot of the KEGG enrichment analysis. The top 20 enriched target genes of differentially expressed lncRNAs. The y-axis represents the pathway name and the x-axis represents the Rich factor. The size of the point indicates the number of differentially expressed genes in the pathway. The color of the point corresponds to a different q-value range; C, clustering analysis of the significantly enriched KEGG pathway. Highly expressed metabolic pathways were indicated in red, and relatively low metabolic pathways were indicated in green. Within each parenthesis after the metabolic pathway tag is the number of genes

Validation of lncRNAs and mRNA expression

 

6 lncRNAs and 6 mRNAs, with differential expression, were chosen randomly in order to validate the reliability of high-throughput sequencing by qRT-PCR. These results showed the similar expression patterns contrasted with the transcriptome profile (Fig. 8, Table S2), indicating that the transcriptome sequencing results were true and reliable.

 

Discussion

 

Drought is one of the severe abiotic stresses that influence plant growth and metabolism in various ways. For crops, the damage to photosynthetic systems caused by drought stress is the main cause of yield reduction (Bodner et al. 2015). Photosynthesis includes photoreaction and dark reaction. The photosynthetic electron transfer chain in photoreaction is mainly composed of four complexes on the thylakoid membrane: Photosynthetic System II (PSII), Cytochrome b6f complex (Cytb6f), Photosynthetic System I (PSI) and ATP synthase complex (Yadav et al. 2017). The Cytb6f mediates the electron transport from PSII to PSI, which is the primary part of the proton gradient formed on both sides of the thylakoid membrane. It is critical in regulating the electron flow direction of the electron transport chain as well as the state transition between PSII and PSI. In addition, the Cytb6f is also the regulatory center of photosynthetic phosphorylation (Finazzi et al. 2016). Ferredoxin (Fd) is a soluble protein that could accept electrons from PSI and contribute electrons to Fd: NADPH reductase (FNR), which produces NADPH for the Calvin cycle and ultimately generates ATP as well as releases oxygen (Zanetti and Allverti 2018). Two doubled haploid lines of winter triticale were treated with water stress by Hura et al. (2019). The results showed that in both lines, water stress caused a decrease in the content of Cytb6f protein, and the quantum efficiency of electron transport from the original electron acceptor to the final receptor was correspondingly reduced. In photosynthesis, excessive reactive oxygen species (ROS), produced by reduced protein levels of the Cytb6f, would damage the components and structures of the photosynthetic system, and ultimately limit plant growth and yield (Yamori et al. 2016; Hura et al. 2018). Under experimental conditions, the expression level of the Cytb6f was decreased, possibly to protect the PS I electron acceptor from excessive reduction. The expression levels of the above components in photosynthetic system

 

 

Fig.7: Functional enrichment of differentially expressed mRNA under drought treatment. A, GO classification plot; B, scatter plot of the KEGG enrichment analysis. The top 20 enriched mRNAs with differential expression. The y-axis represents the pathway name and the x-axis represents the Rich factor. The size of the point indicates the number of differentially expressed genes in the pathway. The color of the point corresponds to a different q-value range; C, clustering analysis of the significantly enriched KEGG pathway. Highly expressed metabolic pathways were indicated in red, and relatively low metabolic pathways were indicated in green. Within each parenthesis after the metabolic pathway tag is the number of genes

 

were down-regulated after drought stress treatment, which is in line with the results obtained in this study. YP_002000536.1, YP_002000516.1 and YP_002000499.1, which are target genes (the functions are annotated as FD, Cytb6f, and Cyt f large domain respectively, regulated by the corresponding lncRNA, see Table S6), were down-regulated after drought treatment, possibly related to the role of FD and Cytb6f in photosynthetic electron transport. The decreased expression under stress can reduce electron transfer to limit the production of ROS and avoid structural damage to protect the photosynthetic system.

 

Fig. 8: Relative expression of 12 selected genes measured by qRT-PCR under drought treatment. 6 mRNAs: BRADI1G47820, BRADI5G25810, BRADI1G46602, BRADI4G37860, BRADI3G39467, BRADI2G11730; 6 lncRNAs: XLOC_253993, XLOC_168454, XLOC_301959, XLOC_308502, XLOC_253993, XLOC_304290

 

 A large part of the plant genome is used for transcriptional regulation, for instance, 57 MADS-box genes have been identified in B. distachyon, involving in abiotic stress responses, including salt, drought, and low-temperature (Wei et al. 2014). The expression patterns of these TFs were different, among of which, 15 were up-regulated while 2 were down-regulated. APETALA1/FRUITFULL belongs to a family of the MADS-box TF which regulates the growth of sepals, petals and floral meristems (Li et al. 2016). AP1/FUL genes (also known as FUL-like genes) in monocot can be divided into three groups, including FUL1, FUL2 and FUL3 (Wu et al. 2017), which tend to have unnecessary functions in induction and differentiation of flower meristem. The expression patterns of FUL-like genes in monocot are more diverse than those of dicots, revealing that they may exert different functions in monocot (Wu et al. 2017). We employed the high-throughput sequencing to obtain a set of lncRNA whose targeted gene was FUL1, down-regulated in drought. Although most members of MADS-box TF show a positive expression to several abiotic stresses like cold, salt and drought, some of them present a negative regulatory relationship between response and stresses, such as hormones. The expression of 12 HbMADS-box genes was up-regulated after drought stress in rubber tree (Wei et al. 2018), while MADS-box gene ZMM7-L in maize down-regulated by exogenous ABA (Zhang et al. 2012). Therefore, we speculated that there are some MADS-box genes playing negative or reversed role in stress response. The WUSCHEL-related homeobox (WOX) TFs, which are specific in plant, play an essential part in a variety of physiological and developmental processes (Li et al. 2019). It may not only integrate auxin and cytokinin signaling to regulate crown root development, but also respond to various abiotic stresses, like drought, cold, and high salinity (Cheng et al. 2014). The expression of WOX11in rice could be quickly induced by drought stress, and the up-regulated expression further enhance the drought resistance, nevertheless the wox11 mutant is more sensitive to drought (Cheng et al. 2014). Cheng et al. (2016) showed that, compared to wild type, the transgenic plants with overexpressed OsWOX11 significantly increased root hair length and amount, which are conducive to water absorption and prompting drought resistance. Previous research also indicated that, as a high hierarchical regulator, WOX11 not only involves in root development and carbon metabolism, but also regulates other transcription factors as well as genes in different pathways (Jiang et al. 2017). From the current research, we screened a targeted mRNA named WOX11, which responded the drought stress and expressed higher in B. distachyon. According to the conservation of WOX genes, it might share the similar function like in rice and A. thaliana. However, the further research is required to verify its role in drought resistance.

Ribosome proteins function in two ways in organisms: they not only participate in protein synthesis as an important component of ribosomes; but also take part in the growth and development processes (Lee et al. 2018). The ribosomal protein genes highly expressed in rice plants infected with leaf blight, which showed a high response to stress signals, indicating that ribosomal proteins were a valuable resource to regulate plant stress resistance (Moin et al. 2016). Kang et al. (2016) isolated ribosomal P3 protein from Arabidopsis and proved that it can protect cells from high and low temperature stress. Although the expression of ribosomal proteins is relatively stable, environmental stress still affects its transcription. In this study, the genes encoding ribosomal proteins S2, S18, S19/S15, L33 and L23/25 are target genes for drought-responsive lncRNAs, but drought inhibited the expression of these target genes, which indicates that these ribosomal proteins may play a negative regulatory role under drought stress. However, the specific mechanisms and approaches still remain to be elucidated.

Apart from photosynthetic systems, riboproteins, MADS-box TFs and WOX TFs responding to drought stress in this experiment, some enzymes involved in nitrogen metabolism and carbon metabolism also showed distinct expression patterns. Caseinolytic protease (ClpP) is a serine protease with high conservation. The expression diversity of ClpP is particularly evident in higher plants, especially under stress conditions. Overexpression of rice ClpD1 protein in transgenic Arabidopsis plants could increase their tolerance to salt and drought stress (Mishra and Grover 2016). However, under long-term drought stress in the field, the expression level of ClpP protein remained stable in drought-tolerant wheat varieties Zlatitza and Yantar, but decreased in drought-sensitive wheat cultivars Miziya and Dobrudjanka (Vassileva et al. 2012). The target gene encoding ClpP was down-regulated after drought treatment in our experiment, similar to the results of Vassileva et al. (2012). As a sessile organism, the complexity of the Clp family in plant may be reflected in the response to multiple environmental stresses during its long-term evolution (Williams et al. 2019). The glycoside hydrolases in plant not only take part in the metabolism of various carbohydrates, such as cell wall polysaccharide metabolism, but also involve in the biosynthesis and regulation of glycans, adversity defense, signal transduction, secondary metabolism and so on (Yue et al. 2019). A glycoside hydrolase that responds to drought stress obtained in this experiment is described as an alpha amylase (AMY1), which is down-regulated after drought treatment but is inconsistent with previous studies. In terms of previous reports, three genes (aAmy1, aAmy3, and aAmy8) that encode α-amylase were up-regulated significantly in rice under submergence conditions (Du et al. 2014). It is speculated that the high expression of alpha amylase under stress conditions may be related to stress response, which can provide energy for plant growth by hydrolyzing starch into metabolizable sugars to overcome subsequent stresses (Shaw et al. 2017). However, the results of this study do not correspond to this speculation, suggesting that alpha-amylase may have additional regulatory mechanisms under adverse conditions.

Transcriptional regulation and post-transcriptional regulation are important ways to regulate gene expression (Gonçalves et al. 2017). In recent years, increasing genomics and genetic studies have focused on lncRNAs, which are the essential part of transcriptional and post-transcriptional regulation rather than transcriptional noise in plant. LncRNAs not only participate in the growth and development of plant, but also involve in the response of stresses (Di et al. 2014; Yuan et al. 2016). LncRNAs may be involved in the regulation of plants at transcriptional and post-transcriptional level through a variety of pathways. Some lncRNAs possess a binding site for miRNAs, so they can act as targets for specific miRNAs via target mimics, in order to prevent interactions between miRNAs and their targets (Wang et al. 2017a). For example, a lncRNA in rice, osa-eTM160, has been demonstrated as an endogenous target mimics of osa-miR160 and represses expression of osa-miR160 to up regulate the osa-ARF18, which is related with the reproductive development of rice (Wang et al. 2017b). A few plant lncRNAs could serve as precursors to small ncRNAs (sRNAs), such as miRNAs and siRNAs, to produce sRNA (Wang et al. 2017a). Certain antisense lncRNAs in plant would respond to environmental stress by interacting with sense mRNAs (Wang et al. 2017a). Antisense RNA might take part in regulation of gene expression at various levels, including genomic imprinting, transcription, transcriptional interference, RNA stability, alternative splicing, and chromatin modification (Wight and Werner 2013; Zhang et al. 2013). Wang et al. (2014a) confirmed antisense lncRNAs in A. thaliana in response to light, which may form a complex with sense mRNA affecting gene expression in the opposite strand. We also obtained some lncRNAs that responded to drought in the leaves of B. distachyon under drought treatment, and their expression patterns were quite different. Even if lncRNAs predicted to be the same target gene, there may be an opposite expression pattern. For example, the target gene described as WOX TF, induced by drought, is regulated by 27 lncRNAs, 12 of which are down-regulated and 15 are up-regulated (Table S7). Although these lncRNAs may work together on the same target gene, the results indicate that lncRNAs might regulate target gene expression through multiple pathways and participate in diverse biological processes. The fine regulation mechanism and action mode still need further research to explore.

 

Conclusion

 

A total of 90 lncRNAs with differential expression were screened in this research. According to the functional annotations, the predicted target genes were divided into five categories, including three photosynthesis-related genes, two transcription factors, one peptide, five ribosomal proteins, and one glycoside hydrolase. Most of the target genes were down-regulated under drought treatment except the transcription factor wox11. All of the valuable information provides a theoretical basis for further characterizing candidate genes involved in drought stress. However, further studies of annotated and novel transcripts are required to elucidate their function in adaptation mechanisms under drought stress. In conclusion, our research improves new perspectives into the defense mechanisms of drought response in B. distachyon, and also improves the comprehensive understanding concerning the drought tolerance as well as resistance in other temperate cereal.

 

Acknowledgements

 

This research was financially supported by the National Key R&D Program of China (Grant No. 2017YFD0301100), National Natural Science Foundation of China (Grant No. 31401323), the Natural Science Foundation of Henan Province (Grant No. 182300410040), the Key Research Projects of Higher Education in Henan Province (Grant No. 18B210002), the Fund of Henan University of Science and Technology (Grant Nos. 13480072 and 09001814), HAUST discipline improvement and promotion plan A (Grant No. 13660002), and Twelfth Five-Year National Science and Technology Pillar Programme (Grant No. 2011BAD16B07).

 

References

 

An N, S Fan, Y Wang, L Zhang, C Gao, D Zhang, MingyuHan (2018). Genome-wide identification, characterization and expression analysis of long non-coding RNAs in different tissues of apple. Gene 666:44‒57

Bardou F, F Ariel, CG Simpson, N Romero-Barrios, Laporte, S Balzergue, JWS Brown, M Crespi (2014). Long noncoding RNA modulates alternative splicing regulators in Arabidopsis. Dev Cell 30:166‒176

Blümel M, N Dally, C Jung (2015). Flowering time regulation in crops – what did we learn from Arabidopsis? Curr Opin Biotechnol 32:121‒129

Bodner G, A Nakhforoosh, HP Kaul (2015). Management of crop water under drought: a review. Agron Sustain Dev 35:401‒442

Brutnell TP, JL Bennetzen, JP Vogel (2015). Brachypodium distachyon and Setaria viridis: Model genetic systems for the grasses. Annu Rev Plant Biol 66:465‒485

Cheng S, DX Zhou, Y Zhao (2016). WUSCHEL-related homeobox gene WOX11 increases rice drought resistance by controlling root hair formation and root system development. Plant Signal Behav 11; Article e1130198

Cheng S, YL Huang, N Zhu, Y Zhao (2014). The rice WUSCHEL-related homeobox genes are involved in reproductive organ development, hormone signaling and abiotic stress response. Gene 549:266‒274

Di C, JP Yuan, Y Wu, JR Li, HX Lin, L Hu, T Zhang, YJ Qi, MB Gerstein, Y Guo, ZJ Lu (2014). Characterization of stress-responsive lncRNAs in Arabidopsis thaliana by integrating expression, epigenetic and structural features. Plant J 80:848‒861

Do T, Z Qu, I Searle (2019). Purification and Functional Analysis of Plant Long Noncoding RNAs. Humana Press, New York, USA

Du H, N Wu, F Cui, L You, XH Li, LZ Xiong (2014). A homolog of ETHYLENE OVERPRODUCER, OsETOL1, differentially modulates drought and submergence tolerance in rice. Plant J 78:834‒849

Feng YL, KT Wang, YY Zhao, C Ma, J Yin (2015). Virus-induced gene silencing-based functional verification of six genes associated with vernalization in wheat. Biochem Biophys Res Commun 458:928‒933

Finazzi G, J Minagawa, GN Johnson (2016). The Cytochrome b6f Complex: a regulatory Hub controlling electron flow and the dynamics of photosynthesis? Springer, Dordrecht, The Netherlands

Florea L, L Song, SL Salzberg (2013). Thousands of exon skipping events differentiate among splicing patterns in sixteen human tissues. F1000Res 2; Article 188

Gonçalves E, ZR Nakic, M Zampieri, O Wagih, D Ochoa, U Sauer, P Beltrao, J Saez-Rodriguez (2017). Systematic analysis of transcriptional and post-transcriptional regulation of metabolism in yeast. PLoS Comput Biol 13; Article e1005297

Gonzalez, I, R Munita, E Agirre, TA Dittmer, K Gysling, T Misteli, RF Luco (2015). A lncRNA regulates alternative splicing via establishment of a splicing-specific chromatin signature. Nat Str Mol Biol 22:370‒376

Heo JB, S Sung (2011). Vernalization-mediated epigenetic silencing by a long intronic noncoding RNA. Science 331:76‒79

Huang DQ, WR Wu, SR Abrams, AJ Cutler (2008). The relationship of drought-related gene expression in Arabidopsis thaliana to hormonal and environmental factors. J Exp Bot 59:2991‒3007

Huang L, H Dong, D Zhou, M Li, YH Liu, F Zhang, YY Feng, DL Yu, S Lin, JS Cao (2018). Systematic identification of long non-coding RNAs during pollen development and fertilization in Brassica rapa. Plant J 96:203‒222

Hura T, K Hura, A Ostrowska, J Gadzinowska, A Fiust (2019). Water stress-induced flag leaf senescence may be accelerated by rehydration. J Plant Physiol 236:109‒116

Hura T, K Hura, A Ostrowska, J Gadzinowska MT Grzesiaka, K Dziurka, E Dubas (2018). Rieske iron-sulfur protein of cytochrome-b6f is involved in plant recovery after drought stress. Environ Exp Bot 156:228‒239

Initiative IB (2010). Genome sequencing and analysis of the model grass Brachypodium distachyon. Nature 463:763–768

Jiang W, SL Zhou, Q Zhang, HZ Song, DX Zhou, Y Zhao (2017). Transcriptional regulatory network of WOX11 is involved in the control of crown root development, cytokinin signals, and redox in rice. J Exp Bot 68:2787‒2798

Kang CH, YM Lee, JH Park, GM Nawkar, HT Oh, MG Kim, SI Lee, WY Kim, DJ Yun, SY Lee (2016). Ribosomal P3 protein AtP3B of Arabidopsis acts as both protein and RNA chaperone to increase tolerance of heat and cold stresses. Plant Cell Environ 39:1631‒1642

Kong L, Y Zhang, ZQ Ye, XQ Liu, SQ Zhao, L Wei, G Gao (2007). CPC: assess the protein-coding potential of transcripts using sequence features and support vector machine. Nucl Acids Res 35:345‒349

Lee CH, M Kiparaki, J Blanco, V Folgado, Z Ji, A Kumar, G Rimesso, NE Baker (2018). A regulatory response to ribosomal protein mutations controls translation, growth, and cell competition. Dev Cell 46:456‒469

Li M, R Wang, Z Liu, X Wu, J Wang (2019). Genome-wide identification and analysis of the WUSCHEL-related homeobox (WOX) gene family in allotetraploid Brassica napus reveals changes in WOX genes during polyploidization. BMC Genomics 20; Article 317

Li Q, Y Wang, FX Wang, YY Guo, XQ Duan, JH Sun, HL An (2016). Functional conservation and diversification of APETALA1/FRUITFULL genes in Brachypodium distachyon. Physiol Plant 157:507‒518

Livak KJ, TD Schmittgen (2001). Analysis of relative gene expression data using real-time quantitative PCR and the 2(-ΔΔCT) method. Methods 25:402‒408

Ma C, J Zhang, JL Yuan, JR Guo, Y Xiong, YL Feng (2019). Differential expression of microRNAs are responsive to drought stress and exogenous methyl jasmonate in wheat (Triticum aestivum). Intl J Agric Biol 22:475‒486

Matiu M, DP Ankerst, A Menzel (2017). Interactions between temperature and drought in global and regional crop yield variability during 1961‒2014. PLoS One 12; Article e0178339

Mishra RC, RA Grover (2016). Constitutive over-expression of rice ClpD1 protein enhances tolerance to salt and desiccation stresses in transgenic Arabidopsis plants. Plant Sci 250:69‒78

Moin M, A Bakshi, A Saha, M Dutta, SM Madhav, PB Kirti (2016). Rice ribosomal protein large subunit genes and their spatio-temporal and stress regulation. Front Plant Sci 7; Article 1284

Nejat N, N Mantri (2018). Emerging roles of long non-coding RNAs in plant response to biotic and abiotic stresses. Crit Rev Biotechnol 38:93‒105

Punta M, PC Coggill, RY Eberhardt, J Mistry, J Tate, C Boursnell, N Pang, K Forslund, G Ceric, J Clements, A Heger, L Holm, ELL Sonnhammer, SR Eddy, A Bateman, RD Finn (2012). The Pfam protein families database. Nucl Acids Res 40:290–301

Qin T, HY Zhao, P Cui, N Albesher, LM Xiong (2017). A nucleus-localized long non-coding RNA enhances drought and salt stress tolerance. Plant Physiol 175:1321‒1336

Sakuma Y, K Maruyama, Y Osakabe, F Qin, M Seki, K Shinozaki, K Yamaguchi-Shinozaki (2006). Functional analysis of an Arabidopsis transcription factor, DREB2A, involved in drought-responsive gene expression. Plant Cell 18:1292‒1309

Sasaki T, BA Antonio (2004). Rice Genome as A Model System for Cereals. Springer, Dordrecht, The Netherlands

Scholthof KBG, S Irigoyen, P Catalan, KK Mandadi (2018). Brachypodium: a monocot grass model genus for plant biology. Plant Cell 30:1673‒1694

Shaw AK, PK Bhardwaj, S Ghosh, I Azahar, S Adhikari, A Adhikari, AR Sherpa, SK Saha, Z Hossain (2017). Profiling of BABA-induced differentially expressed genes of Zea mays using suppression subtractive hybridization. RSC Adv 7:43849‒43865

Trapnell C, BA Williams, G Pertea, A Mortazavi, G Kwan, MJV Baren, SL Salzberg, BJ Wold, L Pachter (2010). Transcript assembly and quantification by RNA-seq reveals unannotated transcripts and isoform switching during cell differentiation. Nat Biotechnol 28:511‒515

Vassileva V, K Demirevska, L Simova-Stoilova, T Petrova, N Tsenov, U Feller (2012). Long-term field drought affects leaf protein pattern and chloroplast ultrastructure of winter wheat in a cultivar-specific manner. J Agron Crop Sci 198:104‒117

Wang H, PJ Chung, J Liu, IC Jang, MJ Kean, J Xu, NH Chua (2014a). Genome-wide identification of long noncoding natural antisense transcripts and their responses to light in Arabidopsis. Genome Res 24:444‒453

Wang YQ, XD Fan, F Lin, GM He, W Terzaghi, DM Zhu, XW Deng (2014b). Arabidopsis noncoding RNA mediates control of photomorphogenesis by red light. Proc Natl Acad Sci USA 111:10359‒10364

Wang M, W Zhao, L Gao, L Zhao (2018). Genome-wide profiling of long non-coding RNAs from tomato and a comparison with mRNAs associated with the regulation of fruit ripening. BMC Plant Biol 18; Article 75

Wang JJ, XW Meng, OB Dobrovolskaya, YL Orlov, M Chen (2017a). Non-coding RNAs and their roles in stress response in plants. Genomics Proteomics Bioinform 15:301312

Wang M, HJ Wu, J Fang, CC Chu, XJ Wang (2017b). A long noncoding RNA involved in rice reproductive development by negatively regulating osa-miR160. Sci Bull 62:470‒475

Wei B, RZ Zhang, JJ Guo, DM Liu, AL Li, RC Fan, L Mao, XQ Zhang (2014). Genome-wide analysis of the mads-box gene family in Brachypodium distachyon. PLoS One, 9; e0084781

Wei M, Y Wang, R Pan, W Li (2018). Genome-wide identification and characterization of MADS-box family genes related to floral organ development and stress resistance in Hevea brasiliensis Müll. Arg. Forests 9:1-16

Wight M, A Werner (2013). The functions of natural antisense transcripts. Essays Biochem 54:91‒101

Williams AM, G Friso, KJV Wijk, DB Sloan (2019). Extreme variation in rates of evolution in the plastid Clp protease complex. Plant J 98:243‒259

Wu F, XW Shi, XL Lin, Y Liu, K Chong, G Theißen, Z Meng (2017). The ABCs of flower development: mutational analysis of AP1/FUL-like genes in rice provides evidence for a homeotic (A)-function in grasses. Plant J 89:310‒324

Yadav KNS, DA Semchonok, L Nosek, R Kouřil, G Fucile, EJ Boekema, LA Eichacker (2017). Supercomplexes of plant photosystem I with cytochrome b6f, light-harvesting complex II and NDH. Biochim Biophys Acta 1858:12‒20

Yamori W, E Kondo, D Sugiura, I Terashima, Y Suzuki, A Makino (2016). Enhanced leaf photosynthesis as a target to increase grain yield: insights from transgenic rice lines with variable Rieske FeS protein content in the cytochrome b6/f complex. Plant Cell Environ 39:80‒87

Yuan J, Y Zhang, J Dong, Y Sun, BL Lim, D Liu, ZJ Lu (2016). Systematic characterization of novel lncRNAs responding to phosphate, starvation in Arabidopsis thaliana. BMC Genomics 17; Article 655

Yue C, HL Cao, HZ Lin, J Hu, YJ Ye, JM Li, ZL Hao, XY Hao, Y Sun, YJ Yang, XC Wang (2019). Expression patterns of alpha-amylase and beta-amylase genes provide insights into the molecular mechanisms underlying the responses of tea plants (Camellia sinensis) to stress and postharvest processing treatments. Planta 0:281‒298

Zanetti G, A Allverti (2018). Chemistry and Biochemistry of Flavoenzymes. CRC Press, Boca Raton, Florida, USA

Zhang XM, YF Lii, ZG Wu, A Polishko, HM Zhang, V Chinnusamy, S Lonardi, JK Zhu, RY Liu, HL Jin (2013). Mechanisms of small RNA generation from cis-NATs in response to environmental and developmental cues. Mol Plant 6:704‒715

Zhang ZB, HY Li, DF Zhang, YH Liu, J Fu, YS Shi, YC Song, TY Wang, Y Li (2012). Characterization and expression analysis of six MADS-box genes in maize (Zea mays L.). J Plant Physiol 169:797‒806