Next Article in Journal
Elucidating Pharmacological Mechanisms of Natural Medicines by Biclustering Analysis of the Gene Expression Profile: A Case Study on Curcumin and Si-Wu-Tang
Previous Article in Journal
The Effects of Endogenous Non-Peptide Molecule Isatin and Hydrogen Peroxide on Proteomic Profiling of Rat Brain Amyloid-β Binding Proteins: Relevance to Alzheimer’s Disease?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transcriptional Responses of a Bicarbonate-Tolerant Monocot, Puccinellia tenuiflora, and a Related Bicarbonate-Sensitive Species, Poa annua, to NaHCO3 Stress

1
Asian Natural Environmental Science Center (ANESC), the University of Tokyo, Nishitokyo-shi, Tokyo 188-0002, Japan
2
Graduate School of Agricultural and Life Sciences, the University of Tokyo, Bunkyo-ku, Tokyo 113-8657, Japan
3
Alkali Soil Natural Environmental Science Center (ASNESC), Northeast Forestry University, Harbin Hexing Road, Harbin 150040, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2015, 16(1), 496-509; https://doi.org/10.3390/ijms16010496
Submission received: 22 October 2014 / Accepted: 23 December 2014 / Published: 29 December 2014
(This article belongs to the Section Biochemistry)

Abstract

:
Puccinellia tenuiflora is an alkaline salt-tolerant monocot found in saline-alkali soil in China. To identify the genes which are determining the higher tolerance of P. tenuiflora compared to bicarbonate sensitive species, we examined the responses of P. tenuiflora and a related bicarbonate-sensitive Poeae plant, Poa annua, to two days of 20 mM NaHCO3 stress by RNA-seq analysis. We obtained 28 and 38 million reads for P. tenuiflora and P. annua, respectively. For each species, the reads of both unstressed and stressed samples were combined for de novo assembly of contigs. We obtained 77,329 contigs for P. tenuiflora and 115,335 contigs for P. annua. NaHCO3 stress resulted in greater than two-fold absolute expression value changes in 157 of the P. tenuiflora contigs and 1090 of P. annua contigs. Homologs of the genes involved in Fe acquisition, which are important for the survival of plants under alkaline stress, were up-regulated in P. tenuiflora and down-regulated in P. annua. The smaller number of the genes differentially regulated in P. tenuiflora suggests that the genes regulating bicarbonate tolerance are constitutively expressed in P. tenuiflora.

1. Introduction

Soil salinity and sodicity are major environmental stresses faced by crops. In Songnen Plain in Northeast China, about 3.73 × 106 ha contain elevated levels of alkaline salt, and the area is expanding at a rate of 1.4% annually [1,2]. In alkaline soils, plant stress factors include excess Na+ and HCO3/CO32 ions and high pH. In highly alkaline areas, only a small number of plants can grow.
Recently some transcriptome analyses under bicarbonate stress have been performed on both bicarbonate-tolerant and bicarbonate-sensitive species. The bicarbonate tolerant species include Leymus chinensis [3], Puccinellia tenuiflora [4,5,6,7], Tamarix hispida [8,9], Limonium bicolor [10] and the sensitive species include soybean [11], maize [2], Lotus japonicus [12] and flax [13]. However, most of these studies examined only a single species, which makes it difficult to compare the responses of bicarbonate-tolerant and bicarbonate-sensitive species.
Puccinellia tenuiflora (Griseb.) Scrib. et Merr. is a graminaceous plant found in saline-alkali soil in Songnen Plain, China. It is tolerant to both neutral and alkaline salts, and its tolerance mechanisms have been extensively studied. P. tenuiflora is able to maintain a high K+/Na+ ratio due to a high K+/Na+ selectivity of its plasma membrane [14] and a high ability to limit Na+ influx in the roots [15]. In addition, its leaves exude salts with wax through the stomata [16]. P. tenuiflora is able to maintain high photosynthetic activity under low NaCl stress, possibly through the activity of antioxidant enzymes [17]. P. tenuiflora also accumulates and exudes citric acid under alkaline salt stress, and secretes it from the roots, where it may adjust the pH of the rhizosphere [18]. Several P. tenuiflora genes that are presumably involved in the response to alkaline salt stress have been cloned and characterized [19,20,21,22,23,24,25,26]. EST [4,6,7], microarray [5,6] and proteomics [27] analyses have identified genes in various categories (metabolism, transcription regulation, signal transduction, transport etc.) that are presumably involved in the responses of P. tenuiflora to salt stress. However, it is unclear whether these responses are specific to P. tenuiflora or are shared with sensitive species.
Here, to obtain insights into the molecular mechanisms of alkaline salt tolerance, we compared the transcriptomes of Puccinellia tenuiflora and the salt-sensitive Poa annua L. (both of the tribe Poeae).

2. Results

2.1. Bicarbonate Stress Tolerance Test of P. annua

P. annua grew faster than P. tenuiflora under the control condition. Under the stress of 300 mM NaHCO3, P. tenuiflora survived but P. annua did not (Figure 1B), indicating that P. annua was a suitable species for a comparison of bicarbonate stress tolerance.
Figure 1. Stress tolerance of P. tenuiflora (left) and P. annua (right). Seedlings of P. annua and P. annua were hydroponically grown for nine days and were transferred to the nutrient solution containing 0 (A); or 300 mM NaHCO3 (B), and were grown for another five days. White bars = 1 cm
Figure 1. Stress tolerance of P. tenuiflora (left) and P. annua (right). Seedlings of P. annua and P. annua were hydroponically grown for nine days and were transferred to the nutrient solution containing 0 (A); or 300 mM NaHCO3 (B), and were grown for another five days. White bars = 1 cm
Ijms 16 00496 g001
P. annua roots continued to elongate at NaHCO3 concentrations up to 30 mM, but stopped growing and the root tips turned black at 40 mM NaHCO3 (Figure 2A). Root growth of P. tenuiflora was reduced at NaHCO3 concentrations as low as 10 mM, but was not stopped by concentrations up to 40 mM NaHCO3. Based on these results, we used 20 mM NaHCO3 for the RNA-seq analysis.
Figure 2. Root elongation during NaHCO3 treatment (A) and shoot length after five days of treatment (B) of P. tenuiflora and P. annua under the indicated concentrations of NaHCO3. Values show the average of three plates, each containing 7–10 seedlings. The error bars represent standard error (SE) and * indicate values that are significantly different from those under 0 mM NaHCO3 (p < 0.05 in Student’s t test).
Figure 2. Root elongation during NaHCO3 treatment (A) and shoot length after five days of treatment (B) of P. tenuiflora and P. annua under the indicated concentrations of NaHCO3. Values show the average of three plates, each containing 7–10 seedlings. The error bars represent standard error (SE) and * indicate values that are significantly different from those under 0 mM NaHCO3 (p < 0.05 in Student’s t test).
Ijms 16 00496 g002

2.2. De Novo Assembly of P. tenuiflora and P. annua Transcripts

After two sequencing runs using the same cDNA libraries (see Experimental Section 4.3 for details), we obtained 29,619,901 reads for P. tenuiflora and 47,700,089 reads for P. annua. After de novo assembly, 77,329 and 115,335 contigs were obtained for P. tenuiflora and P. annua, respectively. Setting stringent parameters for assembly to avoid trans chimeras resulted in relatively short contigs. Species distributions of the top BLASTX hits for the contigs of each species are shown in Figure S1. For both species, the grass Aegilops tauscii was the species with the most hits and the species distributions were very similar.

2.3. Read Mapping and Gene Annotation

The reads from the second run were mapped to the assembled contigs to calculate the expression values, and the expression values were compared between the control and the stressed samples. The numbers of the contigs whose expressions changed >2- or <0.5-fold under NaHCO3 stress compared to the control were 1090 in P. annua and 157 in P. tenuiflora (Table S2). In this table, fold changes of ∞ and −∞ indicate transcripts that were expressed only under the NaHCO3 condition or only under the control condition, respectively. However, transcripts detected only under one condition or the other may include artifacts caused by misassembled contigs.
The expressions of some of the genes which were differentially regulated in the RNA-seq analysis were checked by qRT-PCR. For each of the genes, the qRT-PCR results confirmed the RNA-seq results (Figure 3).
Figure 3. Fold change in expression values (expressed as Log2 values) of selected genes as determined by RNA-seq analysis and real-time quantitative RT-PCR. RNAs extracted from the same plant samples were used for the two analyses and tubulin was used as an internal control for the RT-PCR.
Figure 3. Fold change in expression values (expressed as Log2 values) of selected genes as determined by RNA-seq analysis and real-time quantitative RT-PCR. RNAs extracted from the same plant samples were used for the two analyses and tubulin was used as an internal control for the RT-PCR.
Ijms 16 00496 g003

3. Discussion

3.1. Effects of NaHCO3 Stress on P. tenuiflora and P. annua

P. annua was sensitive to NaHCO3 (Figure 1 and Figure 2). NaHCO3 concentrations as low as 40 mM killed the root tip of P. annua and totally blocked its elongation (Figure 2). On the other hand, roots of P. tenuiflora continued to elongate under 40 mM NaHCO3 treatment. The plants could survive 300 mM NaHCO3 stress treatment for five days, suggesting that the ability to protect root tips from bicarbonate stress is important for survival.
Twenty mM NaHCO3 treatment on agar plates had less effect on shoot growth than on root growth in both species. This finding is contrary to the finding that carbonate stress increased the root/shoot ratio of pea [28]. However, it should be noted that the stress treatment in this study was performed in sealed plates. The air inside the plates would have had a high CO2 concentration generated from bicarbonate, which might have affected the respiration and photosynthesis of the shoots.

3.2. RNA-Seq Analysis and de Novo Assembly

De novo assembly of the reads obtained in the first RNA-seq revealed the presence of many trans chimeras [29] that were later incorrectly assigned as up- or down-regulated genes, probably as the result of high expressions of rRNAs. Although rRNA removal tools and PCR duplicate removal tools have not been commonly used in other transcriptome studies, in the present study, they seemed to reduce the number of trans chimeras in the up- or down-regulated contigs. Setting stringent parameters for assembly also helped to reduce the number of chimeras, although it also resulted in the frequent appearance of possibly the same gene products in the up- and down-regulated genes lists. Because our goal was to identify the kinds of genes that are differentially regulated by bicarbonate stress, we concluded that multiple counts of the same gene products would have less effect on our results than the occurrence of trans chimeras. It is also possible that the contigs that yielded the same BLAST hits are actually orthologs or splice variants.

3.2.1. Genes Differentially Regulated in P. annua under NaHCO3 Treatment

Significantly more genes were differentially regulated in P. annua than in P. tenuiflora (Table S2), suggesting that P. annua was more severely stressed by NaHCO3 than P. tenuiflora. Some of the up-regulated genes were probable homologs of genes involved in stress tolerance and defense mechanisms such as glutathione S-transferase, disease resistance protein RPM1 and mitochondrial chaperone BCS1-B as well as genes involved in metabolism such as aspartic proteinase nepenthesin-2 and patatin group A-3 (Table 1 and “poaup_hit” tab in Table S2). Homologs of genes encoding flavonoid 3'-monooxygenase-like protein and anthocyanidin 5,3-O-glucosyltransferase-like protein were up-regulated, suggesting that secondary metabolites have a role in protecting against bicarbonate stress.
Table 1. Genes that were up-regulated under NaHCO3 treatment (fold change > 2, FDR-corrected p-value < 0.05). The annotations were assigned by BDA (Blast2GO Descriptor Annotator), showing the words most frequently counted of all blast hit descriptions (https://www.blast2go.com/b2gsupport/faqs).
Table 1. Genes that were up-regulated under NaHCO3 treatment (fold change > 2, FDR-corrected p-value < 0.05). The annotations were assigned by BDA (Blast2GO Descriptor Annotator), showing the words most frequently counted of all blast hit descriptions (https://www.blast2go.com/b2gsupport/faqs).
ContigAnnotationFold Change (EDGE Test)
P. annua
Poa_27190Glutathione S-transferase11.34
Poa_3657High affinity nitrate transporter53.21
Poa_12387Anthocyanidin 5,3- O-glucosyltransferase-like5.09
Poa_688ABC transporter b family member 4-like62.27
Poa_2744Aspartic proteinase nepenthesin-27.68
Poa_18988Cytochrome p45054.50
Poa_15731Disease resistance protein RPM118.52
Poa_4898Flavonoid 3-monooxygenase-like14.47
Poa_6299Mitochondrial chaperone BCS1-B2.93
Poa_8595Patatin group A-33.97
Poa_2480Phosphate transporter176.58
Poa_8940Phosphoenolpyruvate carboxykinase54.00
Poa_12214Potassium channel SKOR6.51
Poa_25791Probable WRKY transcription factor 70-like16.62
Poa_19228Subtilisin-like protease9.35
P. tenuiflora
Put_2050Boron transporter3.58
Put_11515Carbonic anhydrase3.90
Put_1357High-affinity nitrate transporter-like3.53
Put_763Leucine-rich repeat receptor-like protein kinase at2g19210-like3.48
Put_2064Long chain acyl-CoA synthetase 4-like5.03
Put_4625Metal-nicotianamine transporter YSL63.68
Put_22802Nicotianamine aminotransferase A-like4.65
Put_1383Phosphate transporter25.87
Put_1973Phosphoenolpyruvate carboxykinase7.47
Put_7021Sucrase-like protein6.08
Put_3399Sulfate transporter3.21
Put_4027Thionin-like peptide6.82
Put_10418Zinc transporter5.74
EDGE = Empirical analysis of digital gene expression [30].
High affinity nitrate transporters and predicted inorganic phosphate transporters were up-regulated in both P. annua and P. tenuiflora under bicarbonate stress. There have been conflicting reports on the effect of HCO3 on the uptake of NO3 [31,32,33,34,35,36,37] and on phosphorus status in plants [38,39], and the roles of these transporters under bicarbonate stress are not clear. However, the genes for nitrate transporters in alkaline salt-tolerant ecotype of Lotus japonicus has been reported to be up-regulated under bicarbonate stress [40]. Phosphoenolpyruvate carboxykinase (PEPCK) was strongly induced by NaHCO3 in both species, in agreement with previous studies [11,41], and may have a role in maintaining homeostasis of intracellular CO2/H2CO3/HCO3 and/or organic acids. However, its function in the response to NaHCO3 stress is unclear.
Transcripts involved in metal acquisition and homeostasis were down-regulated only in P. annua (Table S2). Down-regulated transcripts included homologs of iron-phytosiderophore transporter, nicotianamine aminotransferase A (NAAT-A) and nicotianamine synthase (NAS). Bicarbonate stress induces Fe chlorosis in plants [42,43]. High pH and HCO3 ion have been suggested to impede the solubility of Fe in the rhizosphere and plant apoplast, making Fe unavailable for plant cells [42,43,44]. NASs and NAATs are involved in synthesis of phytosiderophores, which are members of the mugineic acid family. Overexpression of OsNAS2 led to a higher tolerance of Fe deficiency and to high pH [45], and gene expression of NAAT was reported to be induced by Fe deficiency in barley [46]. Metal-nicotianamine transporter YSL (yellow stripe 1-like) transports Fe-chelated phytosiderophore complexes from rhizosphere into root cells. Putative homologs of transporters such as high-affinity potassium transporters [47] and zinc-induced facilitator-like 1 [48] were also down-regulated in P. annua, suggesting that NaHCO3 disrupts the ion homeostasis in P. annua. Down-regulation of β-expansin 1a precursor and cellulose synthase only in P. annua may partly account for the reduced root growth of the species under NaHCO3 stress.

3.2.2. Genes Differentially Regulated in P. tenuiflora under NaHCO3 Treatment

Metal-nicotianamine transporter and NAAT-A, which were down-regulated in P. annua, were up-regulated in P. tenuiflora. The difference in the responses of genes related to Fe-acquisition in P. tenuiflora and P. annua (Table 1 and Table S2) suggests that P. tenuiflora is better able to cope with Fe deficiency under alkaline salt stress. A putative homolog of BOR-like 2 was also induced in P. tenuiflora. BOR is involved in the export of boron from the cytoplasm to the apoplast [49]. Although 10 mM NaHCO3 treatment did not reduce the concentration of boron in leaves of Lotus tenuis [40], boron transporters may have an important function in saline-alkali fields since boron is said to be adsorbed to the soil in the presence of CaCO3 under high pH [50]. Other genes that were up-regulated only in P. tenuiflora included putative homologs of genes involved in metabolism (e.g., sucrase-like protein) and defense (e.g., thionin-like peptide [51]). Metabolic responses and defense responses must be important in both P. annua and P. tenuiflora under NaHCO3 stress, but the genes involved seem to be different.
Transcripts down-regulated in P. tenuiflora include members of families that are known to be involved in other stress responses (e.g., cytochrome p450 and WRKY transcription factors; Table 2 and Table S2). PutHKT2;1, which is described as a “salt transporter” in Table 2 and which has a high-affinity Na+-K+ symport activity [21], was also down-regulated. Thus, down-regulation of PutHKT2;1 might contribute to maintaining intracellular ion homeostasis under weak bicarbonate stress conditions.
Table 2. Genes that were down-regulated under NaHCO3 treatment (fold change < −2, FDR-corrected p-value < 0.05).
Table 2. Genes that were down-regulated under NaHCO3 treatment (fold change < −2, FDR-corrected p-value < 0.05).
ContigAnnotationFold Change (EDGE Test)
P. annua
Poa_403612'-Deoxymugineic-acid 2'-dioxygenase-like−3.90
Poa_7047Actin−3.31
Poa_13081Ammonium transporter−2.94
Poa_24077β-Expansin 1a precursor−8.33
Poa_8466Cellulose synthase−3.48
Poa_9099GDSL esterase lipase At5g45910-like−2.24
Poa_2507High-affinity potassium transporter−2.80
Poa_21856Iron-phytosiderophore transporter−3.51
Poa_36815Mate efflux family protein chloroplastic-like−6.77
Poa_2955Nicotianamine aminotransferase A-like−4.99
Poa_64935Nicotianamine synthase 3−90.37
Poa_58054Protein zinc induced facilitator-like 1-like−50.8
Poa_41522STOP1−9.44
Poa_6256Urea active transporter 1−2.47
Poa_9416Wall-associated receptor kinase 2-like−3.47
P. tenuiflora
Put_32521Cytochrome p450 716b1-like−3.72
Put_28631Sodium transporter−8.04
Put_52287WRKY transcription factor−5.15

4. Experimental Section

4.1. Plant Materials, Growth Conditions and Stress Treatments

Seeds of P. tenuiflora were collected in an alkaline soil area located in North-East China. Seeds of Poa annua were kindly provided by Masaru Ogasawara at University of Utsunomiya, Japan. Seeds were surface-sterilized by washing with 70% v/v ethanol for 5 min. and subsequently with 50% v/v NaClO for 15 min. Growth chamber was maintained at 28 °C during the day and 22 °C at night while the daily photoperiod of 350–400 µmol·m−2·s−1 was 12 h.
To compare the survival rates under strong alkaline salt stress, the surface-sterilized seeds were sown in tap water and were grown for nine days. The seedlings were transferred to nutrient solution containing 6 mg/L (NH4)2SO4, 2 mg/L K2SO4, 8.2 mg/L MgSO4, 2.3 mg/L KNO3, 7.5 mg/L Ca(NO3)2, 3.1 mg/L KH2PO4, 10 mg/L Fe-EDTA with 0 or 300 mM NaHCO3. Photos were taken after 5 days. The experiment was repeated twice and similar results were obtained.
For measuring the shoot and root length under weak alkaline salt stress, the surface-sterilized seeds were sown on 0.8% w/v agar plates containing the nutrient solution described above. After 4 days (P. annua) and 9 days (P. tenuiflora) of germination, the plants were transferred to 0.8% w/v agar plates containing the nutrients mentioned above and 0, 10, 20, 30 or 40 mM NaHCO3. The P. tenuiflora seedlings were allowed to germinate longer because they grow more slowly than P. annua seedlings. Shoot length and root elongation were measured after 5 days.
For RNA-seq analysis, surface-sterilized seeds were hydroponically grown for 19 days. Water was changed every 3–4 days. The plants were transferred to the nutrient solution containing 0 or 20 mM NaHCO3 and harvested after 48 h. Three biological replicates with >100 plants each were obtained.

4.2. RNA Extraction and cDNA Library Construction

Plants harvested in the previous subsection were separated into shoots and roots, frozen immediately using liquid nitrogen and ground to fine powder. In the first sequencing, mRNA was extracted directly from powdered tissue samples with a Dynabeads® DIRECT™ Micro Kit (Life Technologies, Carlsbad, CA, USA) following the manufacturer’s instructions. In the second sequencing, total RNA was extracted from frozen root samples using an RNeasy Plant Mini Kit (QIAGEN, Venlo, The Netherlands). mRNA was then isolated from total RNA with the Dynabeads Kit (See subsection 4.3 for details; Life Technologies). mRNA samples were fragmented, reverse transcribed and amplified to make barcoded whole transcriptome libraries using Ion Total RNA-seq Kit v2 (Life Technologies). Yield and size distribution of the fragmented RNA and the amplified cDNA were checked using an Agilent 2200 Tapestation with High Sensitivity RNA ScreenTape® and High Sensitivity D1000 ScreenTape® (Agilent Technologies, Palo Alto, CA, USA) at each step. For libraries whose peak sizes of amplified cDNAs were <200 bp (all the libraries in the first run and P. annua control #2, P. tenuiflora control #3 and P. tenuiflora control #1 in the second run), cDNAs with the sizes around 280 bp were selected using E-Gel® SizeSelect™ Agarose Gel (Life Technologies). Ion OneTouch™ System with Ion PI™ Template OT2 200 Kit v3 (Life Technologies) was used to prepare enriched, template-positive Ion PI™ Ion Sphere Particles. The whole library preparation and sequencing were performed twice from the same sample sets, obtaining 24 libraries from 12 sample sets (see the next section for detail).

4.3. Next Generation Sequencing and Data Analysis

Sequencing was performed using the Ion Proton™ System with an Ion PI™ Sequencing 200 Kit v3 (Life Technologies) following the manufacturer’s instructions. The sequencing results are summarized in Table S1 . The numbers of raw reads were different in the different libraries, which was caused by an error in diluting the cDNA libraries during the library preparation step. Since the expression values calculated in the RNA-seq analysis described later were normalized to the number of total reads in the given libraries, the difference in the number of total reads should not greatly impact the conclusions. Sequencing results were imported into CLC Genomic Workbench 7.5 (CLC bio, Aarhus, Denmark) as FASTQ files for further analysis. On CLC Genomic Workbench, the raw reads with the quality score less than 0.05 were trimmed using the “Trim Sequences” tool. Reads shorter than 15 bp were discarded. The average quality scores of the trimmed reads were 22.89 for P. annua and 23.08 for P. tenuiflora. In the first trial of RNA-seq, the trimmed reads were de novo assembled just after this step. However, BLASTN analysis revealed that many of the resulting contigs were trans chimeras [29] resulting from misassembly of different gene products into one contig. Many of them seemed to be chimeras of rRNA and other transcripts, so in the second trial we tried to reduce rRNA contamination by extracting mRNAs from total RNA instead of directly from tissue samples. This way rRNA contamination was reduced substantially (Table S1), but not totally removed. Thus we used SortMeRNA (ver. 1.99 beta, [52]) to filter out probable rRNA contamination from the trimmed reads. The average quality scores of the trimmed reads from the second run were 23.60 for P. annua and 23.27 P. tenuiflora. The trimmed reads were exported as FASTQ files from CLC Genomics Workbench. rRNA was removed from the reads with SortMeRNA with the databases supplied with the software (silva-euk-18s-database-id95.fasta & silva-euk-28s-database-id95.fasta). The “length of the sliding window” option was set (“-L 14”) to allow reads longer than 14 bp. Default settings were used for other parameters. Reads not assigned as rRNA were imported back to CLC Genomics Workbench. The same process was applied to the reads from the first run, and the reads from the two runs were combined to be used for de novo assembly on CLC Genomics Workbench. The reads were de novo assembled for each plant species using 12 libraries each with the word and bubble size automatically set by the software to yield contigs (word size 23 and 24 bp, bubble size 108 and 119 bp for P. annua and P. tenuiflora, respectively). Minimum contig length was set to 200 bp. To raise the precision of the contigs, the trimmed reads were mapped back to the resulting contigs with the following settings: mismatch cost 2, insertion cost 3, deletion cost 3, length fraction 0.8 and similarity fraction 0.95. The contig sequences were updated occasionally according to the mapped reads. After the removal of probable PCR duplicates from the trimmed reads using Duplicate Read Removal plugin (ver. 1.0, beta, CLC bio) of the CLC Genomics Workbench, expression analysis was performed with RNA-Seq Analysis Tool of CLC Genomic Workbench for each sample groups of the second run, using de novo assembled contigs as references. Parameters for read mapping were set as follows: Mismatch cost 2, insertion cost 3, deletion cost 3, length fraction 0.8, similarity fraction 0.8 and maximum number of hits for a read 10. Expression levels were compared between control groups and stressed groups using Empirical Analysis of DGE [30,53] tool on CLC Genomics Workbench 7.5. Reads per kilobase of exon per million mapped reads (RPKM) was also calculated. Contigs whose absolute fold changes were >2 (FDR-corrected p-values <0.05) by EDGE test were selected as up-regulated or down-regulated transcripts. For those contigs, homology searches using BLASTX (BLAST+ ver. 2.2.30, NCBI) against the Poaceae (taxid: 4479) protein sequences from NCBI nr protein database and annotation with Blast2GO (ver. 2.7.2; [54]) were performed. Summary of rRNA removal and duplicate removal are shown in Table S1.
The FASTQ files of the raw reads and the sequences of differentially regulated contigs were deposited to DDBJ Sequence Read Archive (BioProject ID: PRJDB3227).

4.4. Real-Time PCR Analysis

Selected contigs that were differentially regulated by the NaHCO3 stress were examined by real-time PCR analysis. Total RNA was extracted from the same plant root samples that were used in RNA-seq analysis with RNeasy Plant Mini Kit (QIAGEN), and reverse transcribed using High-Capacity RNA-to-cDNA™ Kit (Life Technologies) following the manufacturer’s instructions. The cDNA was diluted 20 times and 1 μL of the diluted cDNA was used as the template for quantitative RT-PCR analysis. SYBR® Premix Ex Taq™ II (Tli RNase H Plus) (TaKaRa, Shiga, Japan) and StepOne™ Real-Time PCR System (Applied Biosystems, Framingham, MA, USA) were used for the PCR. A tubulin gene from P. tenuiflora was used as an internal standard to normalize the expression data [21]. The primer pairs (Table S3) were designed to yield products with the sizes of 80–200 bp based on the contig sequences. The PCR was performed as follows: 95 °C for 30 s, followed by 40 cycles of 95 °C for 5 s and 60 °C for 30 s. The experiments were carried out in triplicate. Ten-fold serial dilution of 0.2× cDNA mixture of all the samples were used for drawing the standard curve.

5. Conclusions

The present RNA-seq analysis revealed the transcriptome of P. tenuiflora on a larger scale than did previous studies, and showed significant differences in the numbers of genes that respond to NaHCO3 stress between related NaHCO3-tolerant and NaHCO3-sensitive species. The tolerant P. tenuiflora seemed to adapt to bicarbonate stress by regulating a small number of genes including those important for Fe acquisition, which suggests that P. tenuiflora has a high tolerance to NaHCO3, even when grown under unstressed conditions.

Supplementary Materials

Supplementary materials can be found at https://www.mdpi.com/1422-0067/16/01/0496/s1.

Acknowledgments

We are grateful to Professor Masaru Ogasawara at Weed Science Center, Utsunomiya University, for kindly providing the seeds of P. annua. This work was supported by a Grant-in-Aid for JSPS Fellows for S.K. (25-7241).

Author Contributions

Tetsuo Takano and Shio Kobayashi conceived and designed the experiments; Shio Kobayashi, Hina Satone and Engkong Tan performed the experiments; Shio Kobayashi, Hiroyuki Kurokochi and Engkong Tan analyzed the data; Shuichi Asakawa and Shenkui Liu contributed reagents/materials; Shio Kobayashi wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, L.; Seki, K.; Miyazaki, T.; Ishihama, Y. The causes of soil alkalinization in the Songnen Plain of Northeast China. Paddy Water Environ. 2009, 7, 259–270. [Google Scholar] [CrossRef]
  2. Zhang, L.M.; Liu, X.G.; Qu, X.N.; Yu, Y.; Han, S.P.; Dou, Y.; Xu, Y.Y.; Jing, H.C.; Hao, D.Y. Early transcriptomic adaptation to Na2CO3 stress altered the expression of a quarter of the total genes in the maize genome and exhibited shared and distinctive profiles with NaCl and high pH stresses. J. Integr. Plant Biol. 2013, 55, 1147–1165. [Google Scholar] [CrossRef] [PubMed]
  3. Jin, H.; Plaha, P.; Park, J.Y.; Hong, C.P.; Lee, I.S.; Yang, Z.H.; Jiang, G.B.; Kwak, S.S.; Liu, S.K.; Lee, J.S.; et al. Comparative EST profiles of leaf and root of Leymus chinensis, a xerophilous grass adapted to high pH sodic soil. Plant Sci. 2006, 170, 1081–1086. [Google Scholar] [CrossRef]
  4. Wang, Y.; Chu, Y.; Liu, G.; Wang, M.H.; Jiang, J.; Hou, Y.; Qu, G.; Yang, C. Identification of expressed sequence tags in an alkali grass (Puccinellia tenuiflora) cDNA library. J. Plant Physiol. 2007, 164, 78–89. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, Y.; Yang, C.; Liu, G.; Jiang, J. Development of a cDNA microarray to identify gene expression of Puccinellia tenuiflora under saline-alkali stress. Plant Physiol. 2007, 45, 567–576. [Google Scholar]
  6. Wang, Y.; Yang, C.; Liu, G.; Zhang, G.; Ban, Q. Microarray and suppression subtractive hybridization analyses of gene expression in Puccinellia tenuiflora after exposure to NaHCO3. Plant Sci. 2007, 173, 309–320. [Google Scholar] [CrossRef]
  7. Zhang, X.; Wei, L.; Wang, Z.; Wang, T. Physiological and molecular features of Puccinellia tenuiflora tolerating salt and alkaline-salt stress. J. Integr. Plant Biol. 2012, 55, 262–276. [Google Scholar] [CrossRef]
  8. Gao, C.; Wang, Y.; Liu, G.; Yang, C.; Jiang, J.; Li, H. Expression profiling of salinity-alkali stress responses by large-scale expressed sequence tag analysis in Tamarix hispid. Plant Mol. Biol. 2008, 66, 245–258. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, C.; Gao, C.; Wang, L.; Zheng, L.; Yang, C.; Wang, Y. Comprehensive transcriptional profiling of NaHCO3-stressed Tamarix hispida roots reveals networks of responsive genes. Plant Mol. Biol. 2013, 84, 145–157. [Google Scholar] [CrossRef]
  10. Wang, Y.; Ma, H.; Liu, G.; Xu, C.; Zhang, D.; Ban, Q. Analysis of gene expression profile of Limonium bicolor under NaHCO3 stress using cDNA microarray. Plant Mol. Biol. Report. 2008, 26, 241–254. [Google Scholar] [CrossRef]
  11. Ge, Y.; Li, Y.; Zhu, Y.M.; Bai, X.; Lv, D.K.; Guo, D.; Ji, W.; Cai, H. Global transcriptome profiling of wild soybean (Glycine soja) roots under NaHCO3 treatment. BMC Plant Biol. 2010, 10, 153. [Google Scholar] [CrossRef] [PubMed]
  12. Babuin, M.F.; Campestre, M.P.; Rocco, R.; Bordenave, C.D.; Escaray, F.J.; Antonelli, C.; Calzadilla, P.; Gárriz, A.; Serna, E.; Carrasco, P.; et al. Response to long-term NaHCO3-derived alkalinity in model Lotus japonicus ecotypes Gifu B-129 and Miyakojima MG-20: Transcriptomic profiling and physiological characterization. PLoS One 2014, 9, e97106. [Google Scholar] [CrossRef] [PubMed]
  13. Yu, Y.; Huang, W.; Chen, H.; Wu, G.; Yuan, H.; Song, X.; Kang, Q.; Zhao, D.; Jiang, W.; Liu, Y.; et al. Identification of differentially expressed genes in flax (Linum usitatissimum L.) under saline-alkaline stress by digital gene expression. Gene 2014, 549, 113–122. [Google Scholar] [CrossRef] [PubMed]
  14. Peng, Y.H.; Zhu, Y.F.; Mao, Y.Q.; Wang, S.M.; Su, W.A.; Tang, Z.C. Alkali grass resists salt stress through high [K+] and an endodermis barrier to Na+. J. Exp. Bot. 2004, 55, 939–949. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, C.M.; Zhang, J.L.; Liu, X.S.; Li, Z.; Wu, G.Q.; Cai, J.Y.; Flowers, T.J.; Wang, S.M. Puccinellia tenuiflora maintains a low Na+ level under salinity by limiting unidirectional Na+ influx resulting in a high selectivity for K+ over Na+. Plant. Cell Environ. 2009, 32, 486–496. [Google Scholar] [CrossRef] [PubMed]
  16. Guorong, S.; Yongzhen, P.; Hongbo, S.; Liye, C.; Xining, Z.; Haiyan, M.; Wenzhong, C.; Cunxu, W. Does Puccinelia tenuiflora have the ability of salt exudation? Coll. Surf. B. Biointerfaces 2005, 46, 197–203. [Google Scholar] [CrossRef]
  17. Wang, Y.; Sun, G.; Suo, B.; Chen, G.; Wang, J.; Yan, Y. Effects of Na2CO3 and NaCl stresses on the antioxidant enzymes of chloroplasts and chlorophyll fluorescence parameters of leaves of Puccinellia tenuiflora (Turcz.) scribn.et Merr. Acta Physiol. Plant. 2007, 30, 143–150. [Google Scholar] [CrossRef]
  18. Guo, L.Q.; Shi, D.C.; Wang, D.L. The key physiological response to alkali stress by the alkali-resistant halophyte Puccinellia tenuiflora is the accumulation of large quantities of organic acids and into the rhyzosphere. J. Agron. Crop Sci. 2010, 196, 123–135. [Google Scholar] [CrossRef]
  19. Ardie, S.W.; Liu, S.; Takano, T. Expression of the AKT1-type K+ channel gene from Puccinellia tenuiflora, PutAKT1, enhances salt tolerance in Arabidopsis. Plant Cell Rep. 2010, 29, 865–874. [Google Scholar] [CrossRef] [PubMed]
  20. Ardie, S.W.; Nishiuchi, S.; Liu, S.; Takano, T. Ectopic expression of the K+ channel β subunits from Puccinellia tenuiflora (KPutB1) and rice (KOB1) alters K+ homeostasis of yeast and Arabidopsis. Mol. Biotechnol. 2011, 48, 76–86. [Google Scholar] [CrossRef] [PubMed]
  21. Ardie, S.W.; Xie, L.; Takahashi, R.; Liu, S.; Takano, T. Cloning of a high-affinity K+ transporter gene PutHKT2;1 from Puccinellia tenuiflora and its functional comparison with OsHKT2;1 from rice in yeast and Arabidopsis. J. Exp. Bot. 2009, 60, 3491–502. [Google Scholar] [CrossRef] [PubMed]
  22. Bu, Y.; Sun, B.; Zhou, A.; Zhang, X.; Lee, I.; Liu, S. Identification and characterization of a PutAMT1;1 gene from Puccinellia tenuiflora. PLoS One 2013, 8, e83111. [Google Scholar] [CrossRef] [PubMed]
  23. Chang-Qing, Z.; Shunsaku, N.; Shenkui, L.; Tetsuo, T. Characterization of two plasma membrane protein 3 genes (PutPMP3) from the alkali grass, Puccinellia tenuiflora, and functional comparison of the rice homologues, OsLti6a/b from rice. BMB Rep. 2008, 41, 448–454. [Google Scholar] [CrossRef] [PubMed]
  24. Kobayashi, S.; Abe, N.; Yoshida, K.T.; Liu, S.; Takano, T. Molecular cloning and characterization of plasma membrane- and vacuolar-type Na+/H+ antiporters of an alkaline-salt-tolerant monocot, Puccinellia tenuiflora. J. Plant Res. 2012, 125, 587–594. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, H.; Zhang, X.; Takano, T.; Liu, S. Characterization of a PutCAX1 gene from Puccinellia tenuiflora that confers Ca2+ and Ba2+ tolerance in yeast. Biochem. Biophys. Res. Commun. 2009, 383, 392–396. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, X.; Yang, R.; Wang, B.; Liu, G.; Yang, C.; Cheng, Y. Functional characterization of a plasma membrane Na+/H+ antiporter from alkali grass (Puccinellia tenuiflora). Mol. Biol. Rep. 2010, 38, 4813–4822. [Google Scholar] [CrossRef] [PubMed]
  27. Yu, J.; Chen, S.; Wang, T.; Sun, G.; Dai, S. Comparative proteomic analysis of Puccinellia tenuiflora leaves under Na2CO3 stress. Int. J. Mol. Sci. 2013, 14, 1740–1762. [Google Scholar] [CrossRef] [PubMed]
  28. Zribi, K.; Gharsalli, M. Effect of bicarbonate on growth and iron nutrition of pea. J. Plant Nutr. 2002, 25, 2143–2149. [Google Scholar] [CrossRef]
  29. Yang, Y.; Smith, S.A. Optimizing de novo assembly of short-read RNA-seq data for phylogenomics. BMC Genomics 2013, 14, 328. [Google Scholar] [CrossRef] [PubMed]
  30. Robinson, M.D.; Oshlack, A. A scaling normalization method for differential expression analysis of RNA-seq data. Genome Biol. 2010, 11, R25. [Google Scholar] [CrossRef] [PubMed]
  31. Cramer, M.D.; Lewis, O.A.M.; Lips, S.H. Inorganic carbon fixation and metabolism in maize roots as affected by nitrate and ammonium nutrition. Physiol. Plant. 1993, 89, 632–639. [Google Scholar] [CrossRef]
  32. Zioni, A.B.; Vaadia, Y.; Lips, S.H. Nitrate uptake by roots as regulated by nitrate reduction products of the shoot. Physiol. Plant. 1971, 15, 288–290. [Google Scholar] [CrossRef]
  33. Van der Westhuizen, M.M.; Cramer, M.D. The influence of elevated rhizosphere dissolved inorganic carbon concentrations on respiratory O2 and CO2 flux in tomato roots. J. Exp. Bot. 1998, 49, 1977–1985. [Google Scholar] [CrossRef]
  34. Van der Merwe, C.A.; Cramer, M.D. Effect of enriched rhizosphere carbon dioxide on nitrate and ammonium uptake in hydroponically grown tomato plants. Plant Soil 2000, 221, 5–11. [Google Scholar] [CrossRef]
  35. Gao, Z.F.; Lips, S.H. Effects of increasing inorganic carbon supply to roots on net nitrate uptake and assimilation in tomato seedlings. Physiol. Plant. 1997, 101, 206–212. [Google Scholar] [CrossRef]
  36. Alhendawi, R.A.; Römheld, V.; Kirkby, E.A.; Marschner, H. Influence of increasing bicarbonate concentrations on plant growth, organic acid accumulation in roots and iron uptake by barley, sorghum, and maize. J. Plant Nutr. 1997, 20, 1731–1753. [Google Scholar] [CrossRef]
  37. Cai, C.; Wang, J.Y.; Zhu, Y.G.; Shen, Q.R.; Li, B.; Tong, Y.P.; Li, Z.S. Gene structure and expression of the high-affinity nitrate transport system in rice roots. J. Integr. Plant Biol. 2008, 50, 443–451. [Google Scholar] [CrossRef] [PubMed]
  38. Kolesch, H.; Oktay, M.; Höfner, W. Effect of iron chlorosis-inducing factors on the pH of the cytoplasm of sunflower (Helianthus annuus). Plant Soil 1984, 82, 215–221. [Google Scholar] [CrossRef]
  39. Guo, R.; Shi, L.; Ding, X.; Hu, Y.; Tian, S.; Yan, D.; Shao, S.; Gao, Y.; Liu, R.; Yang, Y. Effects of saline and alkaline stress on germination, seedling growth, and ion balance in wheat. Agron. J. 2010, 102, 1252. [Google Scholar] [CrossRef]
  40. Paz, R.C.; Rocco, R.A.; Reinoso, H.; Menéndez, A.B.; Pieckenstain, F.L.; Ruiz, O.A. Comparative study of alkaline, saline, and mixed saline-alkaline stresses with regard to their effects on growth, nutrient accumulation, and root morphology of Lotus tenuis. J. Plant Growth Regul. 2012, 31, 448–459. [Google Scholar] [CrossRef]
  41. Gong, B.; Zhang, C.; Li, X.; Wen, D.; Wang, S.; Shi, Q.; Wang, X. Identification of NaCl and NaHCO3 stress responsive proteins in tomato roots using iTRAQ-based analysis. Biochem. Biophys. Res. Commun. 2014, 446, 417–422. [Google Scholar] [CrossRef] [PubMed]
  42. Mengel, K. Iron availability in plant tissues-iron chlorosis on calcareous soils. Plant Soil 1994, 165, 275–283. [Google Scholar] [CrossRef]
  43. Bavaresco, L.; Giachino, E.; Colla, R. Iron chlorosis paradox in grapevine. J. Plant Nutr. 1999, 22, 1589–1597. [Google Scholar] [CrossRef]
  44. Römheld, V. The chlorosis paradox: Fe inactivation as a secondary event in chlorotic leaves of grapevine. J. Plant Nutr. 2000, 23, 1629–1643. [Google Scholar] [CrossRef]
  45. Lee, S.; Kim, Y.S.; Jeon, U.S.; Kim, Y.K.; Schjoerring, J.K.; An, G. Activation of rice nicotianamine synthase 2 (OsNAS2) enhances iron availability for biofortification. Mol. Cells 2012, 33, 269–275. [Google Scholar] [CrossRef] [PubMed]
  46. Takahashi, M.; Yamaguchi, H.; Nakanishi, H.; Shioiri, T.; Nishizawa, N.K.; Mori, S. Cloning two genes for nicotianamine aminotransferase, a critical enzyme in iron acquisition (Strategy II) in graminaceous plants. Plant Physiol. 1999, 121, 947–956. [Google Scholar] [CrossRef] [PubMed]
  47. Rodríguez-Navarro, A.; Rubio, F. High-affinity potassium and sodium transport systems in plants. J. Exp. Bot. 2006, 57, 1149–1160. [Google Scholar] [CrossRef] [PubMed]
  48. Haydon, M.J.; Cobbett, C.S. A novel major facilitator superfamily protein at the tonoplast influences zinc tolerance and accumulation in Arabidopsis. Plant Physiol. 2007, 143, 1705–1719. [Google Scholar] [CrossRef] [PubMed]
  49. Miwa, K.; Wakuta, S.; Takada, S.; Ide, K.; Takano, J.; Naito, S.; Omori, H.; Matsunaga, T.; Fujiwara, T. Roles of BOR2, a boron exporter, in cross linking of rhamnogalacturonan II and root elongation under boron limitation in Arabidopsis. Plant Physiol. 2013, 163, 1699–1709. [Google Scholar] [CrossRef] [PubMed]
  50. Goldberg, S.; Forster, H. Boron sorption on calcareous soils and reference calcites. Soil Sci. 1991, 152, 304–310. [Google Scholar] [CrossRef]
  51. Epple, P.; Apel, K.; Bohlmann, H. Overexpression of an endogenous thionin enhances resistance of Arabidopsis against Fusarium oxysporum. Plant Cell 1997, 9, 509–520. [Google Scholar] [CrossRef] [PubMed]
  52. Kopylova, E.; Noé, L.; Touzet, H. SortMeRNA: Fast and accurate filtering of ribosomal RNAs in metatranscriptomic data. Bioinformatics 2012, 28, 3211–3217. [Google Scholar] [CrossRef] [PubMed]
  53. Robinson, M.D.; McCarthy, D.J.; Smyth, G.K. edgeR: A Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 2010, 26, 139–140. [Google Scholar] [CrossRef] [PubMed]
  54. Conesa, A.; Götz, S.; García-Gómez, J.M.; Terol, J.; Talón, M.; Robles, M. Blast2GO: A universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics 2005, 21, 3674–3676. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Kobayashi, S.; Satone, H.; Tan, E.; Kurokochi, H.; Asakawa, S.; Liu, S.; Takano, T. Transcriptional Responses of a Bicarbonate-Tolerant Monocot, Puccinellia tenuiflora, and a Related Bicarbonate-Sensitive Species, Poa annua, to NaHCO3 Stress. Int. J. Mol. Sci. 2015, 16, 496-509. https://doi.org/10.3390/ijms16010496

AMA Style

Kobayashi S, Satone H, Tan E, Kurokochi H, Asakawa S, Liu S, Takano T. Transcriptional Responses of a Bicarbonate-Tolerant Monocot, Puccinellia tenuiflora, and a Related Bicarbonate-Sensitive Species, Poa annua, to NaHCO3 Stress. International Journal of Molecular Sciences. 2015; 16(1):496-509. https://doi.org/10.3390/ijms16010496

Chicago/Turabian Style

Kobayashi, Shio, Hina Satone, Engkong Tan, Hiroyuki Kurokochi, Shuichi Asakawa, Shenkui Liu, and Tetsuo Takano. 2015. "Transcriptional Responses of a Bicarbonate-Tolerant Monocot, Puccinellia tenuiflora, and a Related Bicarbonate-Sensitive Species, Poa annua, to NaHCO3 Stress" International Journal of Molecular Sciences 16, no. 1: 496-509. https://doi.org/10.3390/ijms16010496

Article Metrics

Back to TopTop