Next Article in Journal
Selenium Accumulation, Speciation and Localization in Brazil Nuts (Bertholletia excelsa H.B.K.)
Previous Article in Journal
Integrated Use of Aureobasidium pullulans Strain CG163 and Acibenzolar-S-Methyl for Management of Bacterial Canker in Kiwifruit
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification of Genes Differentially Expressed in Response to Cold in Pisum sativum Using RNA Sequencing Analyses

1
ICV—Institut Charles Viollette, EA 7394, USC 1411, Univ. Lille, F-59000 Lille, France
2
ICV—Adaptation au Froid du Pois, USC 1411, INRA, F-80300 Péronne, France
3
GCIE, UE 0972, INRA, F-80203 Péronne, France
4
GeT-PlaGe, Genotoul, US 1426, INRA, F-31326 Castanet-Tolosan, France
5
Evo-Eco-Paleo, UMR 8198, Univ. Lille, CNRS, F-59000 Lille, France
*
Author to whom correspondence should be addressed.
Plants 2019, 8(8), 288; https://doi.org/10.3390/plants8080288
Submission received: 11 June 2019 / Revised: 30 July 2019 / Accepted: 9 August 2019 / Published: 15 August 2019
(This article belongs to the Section Plant Physiology and Metabolism)

Abstract

:
Low temperature stress affects growth and development in pea (Pisum sativum L.) and decreases yield. In this study, RNA sequencing time series analyses performed on lines, Champagne frost-tolerant and Térèse frost-sensitive, during a low temperature treatment versus a control condition, led us to identify 4981 differentially expressed genes. Thanks to our experimental design and statistical analyses, we were able to classify these genes into three sets. The first one was composed of 2487 genes that could be related to the constitutive differences between the two lines and were not regulated during cold treatment. The second gathered 1403 genes that could be related to the chilling response. The third set contained 1091 genes, including genes that could be related to freezing tolerance. The identification of differentially expressed genes related to cold, oxidative stress, and dehydration responses, including some transcription factors and kinases, confirmed the soundness of our analyses. In addition, we identified about one hundred genes, whose expression has not yet been linked to cold stress. Overall, our findings showed that both lines have different characteristics for their cold response (chilling response and/or freezing tolerance), as more than 90% of differentially expressed genes were specific to each of them.

1. Introduction

Cold stress is one of the most important factors that limit plant productivity around the world. Understanding the molecular bases of the cold response is thus essential to breed cold-tolerant varieties. To survive winter frosts, plants need to acquire frost tolerance which depends on the duration and time of the exposition to low temperatures [1] and varies according to species [2,3] and genotypes [4]. Plants can adopt two strategies to overcome frost. The first one consists of escaping the main frost periods, which can be obtained by different adaptive mechanisms in the natural population like developing a rosette form or reducing the hypocotyl length, which has been well documented in Arabidopsis thaliana [5,6,7]. The second one concerns the acquisition of freezing tolerance (FT), through a phenomenon called cold acclimation [8]. Following exposure to low temperatures, plants increase their ability to tolerate cold temperatures. Cold acclimation reveals two mechanisms of tolerance, which include: chilling tolerance and the induction of freezing tolerance (FT). Chilling tolerance represents the ability of a plant to respond to low but positive temperatures inferior to 15 C, and FT is an induced response where plants acquire an increased freezing tolerance upon a prior low but non-freezing temperature treatment [9]. FT is a multigenic and quantitative trait that depends on more numerous metabolic changes than chilling [10]. The acquisition of freezing tolerance during acclimation depends on the duration and time of the exposition to cold [1], and varies according to species [11]. The cold acclimation process consists of a large number of changes at the molecular and metabolic levels [12,13].
Many transcriptomic studies have been undertaken in A. thaliana in order to decipher cold responses [14,15]. Some high throughput transcriptomic analyses have also been performed on cold stress in other plants, among which some have been conducted in legumes (Fabaceae) e.g., Ammopiptanthus mongolicus [16], Glycine max [17], Lotus japonicas [18], Vignia unguiculata, subspecies sesquipedalis [19], Vigna subterranea [20], Medicago truncatula [4], Medicago falcata [21], Medicago sativa [22,23], and Cicer arietinum [24]. All of these studies show, as in A. thaliana, the important role of transcription factors, including CBFs (C-repeat Binding Factors), kinases, and COR (Cold-Responsive) genes in cold regulation and acclimation. CBFs recognize and bind to cis-elements in the promoters of cold-COR genes, thus triggering their expression [25].
Among legumes, we are especially interested in dry pea (Pisum sativum L. (Ps)), which is an important source of proteins and starch for animal feeding and human food. Although dry peas are usually sown in spring in Europe, autumn sowings (winter peas) are desirable as they would allow for increasing and stabilizing the grain yield [26]. However, to permit autumn sowings, plants have to resist winter frost. In order to decipher the cold response in pea, we have already carried out transcriptomic approaches using microarray and suppression subtractive hybridization (SSH) [27,28]. However, these first approaches were limited by their relatively low throughput with only 11,930 non-cold specific ESTs available on the microarray and 5000 ESTs displayed within the SSH libraries.
In this present investigation, using a RNA-sequencing (RNA-seq) approach, we took advantage of the Illumina high throughput technology, allowing us to detect low modulation of expression and to obtain a broader overview of the cold responses in pea. RNA-seq analyses performed on 24 mRNA libraries obtained from cold-treated and control samples of the pea lines Champagne (Ch) and Térèse (Te) led us to obtain nearly 900 million reads, and the resulting quantitative analysis allowed us to identify 793,583,651 clean reads ending up with 4981 differentially expressed genes (DEGs). The use of a known frost-tolerant line (Ch) and a frost-sensitive one (Te) allowed us, by comparing both lines submitted or not to a low temperature treatment, to classify DEGs into three sets: the first one corresponds to genes related to the constitutive differences between the two lines and that are not regulated during cold treatment (Line Response), the second one corresponds to genes in which expression is affected by the cold treatment and show similar expression patterns in the two lines (Temperature Common Response i.e., chilling response), and the last one corresponds to genes in which expression is affected by the cold treatment and present different patterns of expression in the two lines, including genes putatively related to FT (Temperature Line-Specific Response).

2. Materials and Methods

2.1. Plant Materials, Experimental Design, and Cold Stress

Two pea lines with contrasted characteristics for cold tolerance were subjected to a low temperature (LT) treatment. Ch is a frost-tolerant winter forage variety whereas Te is a frost-sensitive spring dry pea variety. Only Ch is able to cold acclimate and survive frost, while Te dies when submitted to negative temperatures even if prior subjected to a cold acclimation period [29]. Both lines were grown in a climatic chamber within eight isolating trays to prevent roots from freezing, each one having one hundred holes containing a Jiffy block. Both lines were grown in two different experiments respectively characterized by a LT period (Figure 1A) and by a control (N) temperature condition (Figure 1B). For the LT experiment, plants were first exposed to 20 °C day/14 °C night at 500 µmol m−2s−1 photosynthetic photon flux (PPF) with a 10 h photoperiod over 21 days. This initial phase was called the nursery period. It was followed by a LT period of 16 days with 8 °C day/2 °C night at 250 µmol m−2s−1 PPF and with a 10 h photoperiod. For the N experiment, the nursery period was extended up to 27 days, in order to be able to sample at the same developmental stages as in the LT experiment. In both experiments (LT and N), after the previous periods, the plants were exposed to freezing over 8 days at 4 °C day/−9 °C night at 150 µmol m−2s−1 PPF with a 10 h photoperiod. Afterwards, a recovery period of 16 days was applied with 16 °C day/5 °C night at 500 µmol m−2s−1 PPF with a 10 h photoperiod.
Three samplings were performed in each experiment (LT and N). For the N experiment, the samplings were performed at the 20th (T0), 22nd (T1), and 27th (T2) day of the nursery period. For the LT experiment the samplings were realized at the 20th day (T0) of the nursery period and at the 3rd (T1) and 16th (T2) day of the LT period (Figure 1). The developmental stage of the plants was regularly checked by the state of leaf unfolding, allowing the sampling of plants at the same developmental stage in both LT and N conditions, with the same final number of degree-days [30] despite differences in number of days until sampling. At the end of the recovery period, we confirmed the phenotype of the two lines, because the Ch plants survived the frost period, unlike the Te ones.
A total of 24 samples were harvested, corresponding to 2 lines × 2 treatments (LT and N) × 3 sampling times (T0, T1, and T2) × 2 biological replicates. For each sample, aerial parts of 3 plants were pooled and were immediately frozen using liquid nitrogen and stored at −80 °C until RNA extraction.

2.2. RNA Extraction and High Throughput Sequencing

Total RNA was isolated with the Plant RNAeasy Mini kit (Qiagen) and quantified using a NanoDrop. The paired-end sequencing was performed using an Illumina HiSeq2000 sequencer at the Next Generation Sequencing (NGS) platform of GeT: http://get.genotoul.fr/. About 900 million raw reads generated from the 24 samples were deposited to the National Center for Biotechnology Information (NCBI) in the SRA (Short Read Archive) database (BioProject #PRJNA543764).

2.3. RNA-Sequencing Analysis, Assembly, and Annotation

The quality of the raw reads was checked using FastQC v0.11.4 [31]. After removing adapter sequences with Cutadapt v1.0 [32], reads were cleaned using Prinseq v0.20.3 [33]. The clean reads of the twenty-four samples were de novo assembled using Trinity v20140717 [34]. The contigs obtained by Trinity were filtered using TransRate v1.03 [35], which detects chimeric sequences, structural errors, incomplete assembly, and base errors. Then, Quast v2.3 [36] was used to assess the quality of the assembly. The transcriptome was annotated with Blastx searches [37] against A. thaliana (TAIR 10) protein databases.
Gene ontology (GO) (http://geneontology.org/ Gene Ontology Consortium and http://pantherdb.org/), and Kyoto Encyclopedia of Genes and Genomes (KEGG) (http://www.kegg.jp/kegg/tool/map_pathway2.html KEGG Mapper) information was assigned based on the A. thaliana homologous proteins, and GO functional classification was performed (https://www.arabidopsis.org/tools/bulk/go/index.jsp). Given the importance of protein kinases [38,39,40,41,42], transcription factors and transcription regulators (TF) [43,44,45,46] in cold stress responses, a specific annotation was made for these classes of genes by combining analyses from iTAK v1.2 (http://bioinfo.bti.cornell.edu/cgi-bin/itak/index.cgi) [47], PlnTFDB (http://plntfdb.bio.uni-potsdam.de/v3.0/), [48] and AGRIS (Arabidopsis Gene Regulatory Information Server) (http://arabidopsis.med.ohio-state.edu/, [49]).

2.4. Differential Expression Analysis and Statistical Tools

For the differential expression analysis, Illumina-cleaned reads from the 24 samples were pseudo-aligned on the de novo transcriptome assembly using Kallisto v0.43.1 [50]. Count data were analyzed using a multifactorial design (line (Ch and Te), treatment (LT and N), and time (T0, T1, and T2)) with the R package DESeq2 [51]. The multiple steps of statistical and clustering analyses led to three sets of distinct DEGs (Differentially Expressed Genes) (Figure S1). Firstly, only unigenes fulfilling the criterion TRUE (when the number of reads is sufficient to perform the statistical test) in DESeq2 were conserved for the next analyses. Then, unigenes represented by less than or equal to 48 normalized counts among the 24 samples were discarded. The p-values were calculated for nine combinations (ChNT0:TeNT0, ChNT0:ChNT1, ChNT0:ChNT2, ChLT0:ChLT1, ChLT0:ChLT2, TeNT0:TeNT1, TeNT0:TeNT2, TeLT0:TeLT1, and TeLT0:TeLT2). Only unigenes with an adjusted p-value ≤ 0.02 in at least one of the nine combinations were conserved. A sorting was then realized according to the significance of the functional annotation (E-value ≤ 9 × 10−4). The resulting set of unigenes was submitted to three successive analyses of variance (ANOVA) using Multi Experiment Viewer [52] statistical tools, in order to identify unigenes related to the differences between the two lines or related to the cold responses. First, a two factor ANOVA was performed allowing us to retain only DGEs which varied according to the line and/or the treatment. Then a one-way ANOVA was realized, providing a set of unigenes related to the differences between the two lines (Line Response). The remaining genes were submitted to an additional one-way analysis allowing us to decipher genes responding identically to the LT in the two lines (Temperature Common Response: TCR) from those which presented different cold responses in Ch and Te (Temperature Line Specific Response: TLSR).
Afterwards, gene expression patterns were built and classified with the MeV clustering tools. After log2 transformation of normalized count and mean-centered reduced fit, hierarchical clustering (HCL) were performed using Pearson’s correlation and average linkage clustering method. GO term enrichment analyses of the different sets of DEG were performed using AmiGO2 (http:/amigo1.geneontology.org/cgi-bin/amigo/term_enrichment).

2.5. Statistical Differentially Expressed Genes (DEGs) Corroboration

In order to support the expression of DEGs identified in this study, we used quantitative polymerase chain reaction (qPCR) data from an earlier published experiment [28] which presented only a few differences in the environmental conditions and the sampling times. The T0, T1, and T2 samples of the RNA-seq study were compared to the T0, T6, and T10 of the qPCR analysis, respectively (Figure 1). Blastn searches (E-value ≤ 1 × 10−25, coverage ≥ 200 pb, ≥ 95% of identity) were performed to link the unigenes representing DEGs in the RNA-seq study to the ESTs used to define primers for the qPCR analysis.

3. Results

3.1. Samples, Sequencing, and Assembly Assessment

Twenty-four RNA libraries were built from RNA samples extracted from the two contrasted pea lines, Ch (frost-tolerant) and Te (frost-sensitive), under LT (a low but non-freezing temperature regime is applied before submitting plants to frost) and N (frost is applied without any prior LT treatment). The Illumina sequencing of the 24 samples led us to obtain 886,477,626 paired-end reads. After removing low-quality sequences, a total of 793,583,651 clean reads were assembled into 150,342 contigs with a minimum length of 201 bp (Table S1). Following an analysis of the quality of the transcriptome assembly, 122,194 high-quality contigs, for a total of 118,787,279 bp and representing 73,225 unigenes were conserved for the further analyses. The largest contig was 11,608 bp long and the N50 value of the assembly was 1904 (Table S1). Blastx searches against A. thaliana protein sequences allowed to annotate 34% (24,854/73,225) of the unigenes (E-value ≤ 9 × 10−4). The distribution of the biological processes of the annotated unigenes resembled those reported for A. thaliana genes, suggesting that the construction of the pea RNA-seq libraries did not induce an enrichment of sequences related to a particular class of function (Figure S2).

3.2. Differential Expression Analysis and Clustering

The threshold applied on DESeq2 results (adjusted p-value ≤ 0.02) retained 11,076 unigenes, presenting a modulation of expression within the 24 different samples. Among them, 9676 unigenes were annotated using A. thaliana proteins. A two-way ANOVA performed on the 9676 annotated unigenes permitted us to detect 4981 genes, in which expression was significantly modulated according to the line and/or to the low temperature treatment (Figure S1). Among them, 33 were compared with data previously obtained by qPCR. We observed that RNA-seq and qPCR data are correlated (R = 0.71) (Figure 2).
The examination of the 33 expression patterns reflects the correlation between the data from the two methods of gene expression analysis and supports the robustness of the whole transcriptomic analysis (Figure S3). Then, the 4981 DEGs were submitted to a one-way ANOVA based on the line factor, which led to identify 2487 genes differentially expressed between Ch and Te but not modulated during the cold treatment (Line Response, LR). This set represents a part of the constitutive differences between the two lines. The hierarchical clustering (HCL) analysis divided this set into two blocks. The first block contained 906 genes, which were more expressed in the Ch than in the Te samples (LR (a), Figure 3A (a)) and inversely, the second one was composed of 1581 genes which were more expressed in the Te than in the Ch samples (LR b, Figure 3A (b)). Then a second one-way ANOVA based on the cold treatment factor was performed with the remaining 2494 (4981 − 2487 = 2494) DEGs (Figure S1) and revealed 1403 genes related to the “Temperature Common Response” (TCR) of the two lines. The HCL analysis carried out on these 1403 significant genes could separate 520 (Figure 3B (a)) and 883 (Figure 3B (b)) genes up-expressed in the N condition and the LT condition, respectively. The remaining 1091 genes (2494 − 1403 = 1091), which responded differently in Ch and Te under the LT treatment were considered to be associated to the “Temperature Line Specific Response” (TLSR). The HCL performed on this last set (Figure 3C) showed four distinct expression patterns. The first subset (TLSR (a)) gathered 253 genes which were more expressed in Ch than in Te and down-regulated during the LT conditions in both lines (Figure 3C (a)). The second subset (TLSR (b)) was composed of 228 genes which were also more expressed in Ch than in Te but up-regulated during the LT conditions (Figure 3C (b)). The third one (TLSR (c)) contained 131 genes that were less expressed in Ch than in Te and down-regulated during the LT conditions (Figure 3C (c)) and the last of the four subsets (TLSR (d)) was composed of 479 genes which were likewise less expressed in Ch than in Te but up-regulated during the LT treatment (Figure 3C (d)).

3.3. Functional Annotation, Gene Ontology (GO) Term Enrichment, and Kyoto Encyclopedia of Genes and Genomes (KEGG) Pathways

For the three sets of DEGs described above (LR, TCR, and TLSR) and distributed into eight subsets, the GO term enrichment analysis is detailed in Figure S4, the mapping on KEGG pathways in Table S2, the list of genes annotated as coding putative kinases in Table S3, and those annotated as coding transcription factors in Figure 4. In addition, a summary of all these results is provided in Table 1 and the list of genes known to be related to cold stress in Table 2. This comprehensive analysis of the function of DEGs allowed us to present an overview of the differences between the two lines and to decipher their respective cold responses.
We observed that the five most significant enriched GO terms related to biological processes were the same for all of the eight subsets. These terms were “metabolic process”, “cellular process”, “localization”, “cellular component organization or biogenesis”, and “response to stimulus”. This result seems to show that there is no difference in biological processes involved in the different subsets, at least not at this level of functional assignment. On the other hand, the enriched GO terms related to cellular localization were different within the subsets. In two cases, LR (a) and TCR (a), the five most significant terms related to localization concerned chloroplast and plastid components. This was also the case for a part of the terms enriched in the TLSR (a) subset. For all other subsets, the major enriched terms were related to cell, cytoplasm, and/or intercellular part.
The mapping of all DEGs on the KEGG pathways revealed relatively important differences in the number of mapped genes according to the subset (Table S2, Table 1). Indeed, 36% of the DEGs of the TLSR (d) subset were mapped on the pathways versus 80% of those of the TLSR (a) subset. This analysis, unlike the enrichment of GO terms, revealed a diversity of pathway representations between the subsets. For example, the most represented pathways in the LR (b) and TCR (b) subsets were related to the ribosome metabolism, while those of the LR (a) and TLSR (b) subsets were related to the amino acid metabolism.
The analysis, using iTak, allowed us to annotate 193 putative kinases among the DEGs of the three sets, including 96, 54, and 43 in the LR, TCR, and TLSR sets, respectively (Table S3, Table 1). Kinases were the most represented in the TLSR (b) and (d) subsets (4.8% for each) and the less represented in the TLSR (c) subset (1.5%). A total of 397 putative TFs was also annotated, with 188, 114, and 95 in the LR, TCR, and TLSR subsets, respectively (Figure 4, Table 1). TFs were the most represented in the TLSR (d) subset (9.6%) and the less represented in the TCR (a) subset (4.8%). The most represented family differs according to the subset, bHLH and AP2-EREBP being the most represented in three subsets each (LR (b), TCR (a) and (b) for bHLH and LR (a), TLSR (b) and (c) for AP2-EREBP). We noted that some of the kinases and TFs are known to be related to cold stress, including Myb, CBF, and WRKY (Table 2 and see below).
The functional classification using GO knowledgebase allowed us to classify 145 genes in GO terms related to the cold response, with 128, 11, and 6 in “response to cold” GO:0009409, “cold acclimation” GO:0009631, and “cellular response to cold” GO:0070417, respectively (Table 1, Table 2). Even if it was expected to identify genes related to these GO terms within the TCR (44 genes: 3.1%) and TLSR (36 genes: 3.3%) sets, it was more surprising to identify 65 genes (2.6%) in the LR set. This could mean that several genes that are known to be induced during cold stress in other species could be constitutively expressed in Ch or Te. The proportion of genes related to these GO terms in the different subsets was between 2.3% (LR (b)) and 6.9% (TLSR (c)).

4. Discussion

The genetic studies previously carried out by our team have shown that the Champagne and Térèse lines differ in their response to cold. Following a period of acclimation, the Ch line becomes tolerant to frost while Te remains sensitive [29,126]. This present study aims to provide molecular elements to explain the cold responses in pea. The statistical treatments of the RNA-seq data of 73,225 unigenes led us to distinguish 4981 DEGs, divided into three main sets according to their expression patterns. The analysis of these genes, their modulation of expression, and their affiliation to metabolic pathways has allowed us to enrich our knowledge on the behavior of these two lines regarding their response to cold.

4.1. Differences in Gene Expression Between The Two Lines Independently of the LT Treatment: Predispositions to Face Cold Stress?

The class of genes considered here corresponds to genes which are differentially expressed in the two lines at T0, i.e., genes which are more expressed in Ch than in Te samples, or inversely. The opposition of gene expression patterns was observed throughout the time course study with an almost constant expression level in each line, independently of the LT treatment (Figure 3A). We can therefore consider these genes as being part of the constitutive differences between the two lines. In both Ch and Te up-expressed subsets (Figure 3A (a) and (b), respectively), the KEGG pathway annotation revealed that a majority of these genes are related to metabolism (84.4% and 68.9% for the Ch and Te up-expressed subsets, respectively) suggesting important intrinsic differences between the two lines regarding cellular metabolism. In particular, 73 and 49 genes were related to amino-acid and energy metabolisms among genes up-expressed in Ch and Te, respectively. In addition, 102 genes in the Ch up-expressed subset and 39 in the Te up-expressed subset were related to RNA and protein metabolism. The GO-enrichment analyses were consistent with these results, highlighting the assignment of genes related to metabolism and to RNA and RNA surveillance pathways in the Ch and Te up-expressed subsets, respectively. Interestingly, even if the majority of the genes considered here did not present any modulation of expression during the time course study, 50/906 (5.5%) and 108/1581 (6.8%) of genes from the Ch and Te up-expressed subsets respectively, are known to be involved in responses to diverse stimuli, notably to responses to stress. In particular, within the Ch up-expressed subset, we identified genes coding a cold-regulated 413 inner membrane protein 1 (AT1G29395), and a 3-oxoacyl-synthase II (AT1G74960), which are known to accumulate in cold/freezing conditions in order to preserve chloroplast membranes integrity [110,111]. We also noted in the Ch up-expressed subset the presence of two genes, the first being a mediator of the RNA polymerase II transcriptional subunit 32 (AT1G11760) which regulates the expression of the second, a dehydration element B1A (CBF3, AT4G25480). Both are well known for their involvement in cold acclimation in A. thaliana [112]. More surprisingly, several genes are also known to be involved in the cold responses within the Te up-subset. For example, we found in this subset three genes coding AGAMOUS-like proteins (AT3G61120, AT4G22950, AT2G45660), known to lead to early flowering in A. thaliana and induced by an extended cold treatment [127]. We also identified genes coding a calmodulin-binding transcription activator protein (AT5G64220), which is involved in the rapid induction of CBF factors [53], and an adenine nucleotide alpha hydrolase-like super family (AT1G09740), which enhances freezing tolerance in Arabidopsis after a short period of cold-acclimation [54]. Moreover, genes coding a raffinose synthase (AT5G20250, [45]), a xyloglucan endotransglucosylas/hydrolase (AT4G03210 [55]), and an enolase (in pea [128] and LOS2, AT2G36530 [56]) were scored in the Te up-expressed subset, all of them being involved in cold stress in A. thaliana. The fact that some genes related to cold stress are up-expressed in Te could also be expected, since this line could possess deficient alleles that would not confer to Te an effective defense against cold.
Concerning the kinases, we observed within the Ch up-expressed subset one gene encoding a MPK3 (mitogen-activated protein kinase 3, AT3G45640) shown to be up-expressed in response to cold stress in A. thaliana [57]. Within the Te up-expressed subset, we identified another MAP3K (MAPK/ERK kinase kinase 1, AT4G08500) that plays a major role in cold stress signaling in A. thaliana [113]. More interestingly, 50 (5.52%) and 138 (8.73%) TFs were identified in Ch and Te up-expressed subsets, respectively. Among them, 50 distinct TF families were represented, including 19 TF families which gathered preferentially expressed genes in Ch and Te (i.e., common TF families) and 7 and 24 families which were represented by genes preferentially expressed in Ch or Te, respectively (i.e., line-specific TF families). Among them, several TFs were previously signaled as differentially expressed in cold stress in legume species [16,18,19,22] or are known to be involved in cold acclimation or freezing tolerance in A. thaliana, e.g., DREB1 (CBF3), CAMTA, NAC, and WRKY [25,129]. The other TFs, including ARID, BSD, mTERF, RWP-RK, S1Fa-like, SOH1, SRS, SWI/SNF-BAF60b, TAZ, and TUB are reported for the first time in pea and may be also involved in the intrinsic phenotypic differences between the two lines.
A recent study, using genetic structure and linkage disequilibrium in a large collection of pea germplasm, also highlighted the line-dependent differences [130]. These differences may be due to the fact that Ch belongs to the winter forage lines cohort and Te is issued from the spring lines set. Overall, our analysis reflects the importance of intrinsic differences between lines at the gene expression level [28,128,131]. Among this “Line Response” set, many genes coding proteins related to cold responses were identified. These results suggest that Ch and Te have constitutive defenses against cold stress, which have been described in the majority of cases as induced in other species and particularly in A. thaliana.

4.2. How Pea Faces Cold Stress

The identification of genes showing a significant increase/decrease of expression in Ch and Te during the LT and no fluctuation in the N experiment suggests that both lines have the capacity to undertake molecular modifications in response to cold stress, that can be considered as the chilling response shared by both genotypes. These genes were gathered into the TCR set. In addition, genes showing variation of expression only in Ch during the LT treatment could be attributed to the FT capacity of this frost tolerant line and were clustered within the TLSR set. Elsewhere, the expression pattern variations are more diverse for this set of TLSR, both at the beginning of the LT and also over time, as well as in N condition between these two lines. Furthermore, it is important to note that for both TCR and TLSR sets, we have scored more up-expressed genes in Te than in Ch.

4.2.1. Chilling Response

The TCR set was separated into two subsets, gathering almost the same number of genes which were down- or up-regulated during the LT treatment. Concerning down-expressed genes, it should be noted that the enrichment of GO terms related to chloroplast could provide evidence that genes involved in the photosynthetic system are hugely affected by chilling (Figure S4 (F)). These observations are in agreement with the fact that cold stress leads to a disruption and/or dysfunction of photosynthesis and causes damages to thylakoid membrane and chloroplastic envelopes [132]. The up-expressed subset, for its part gathered, in particular, several RNA-binding proteins, which are involved in RNA and RNA surveillance pathways, operating in plant responses to abiotic stress [58,133]. Concerning the GO annotations, we observed that 30 (3.40%) and 14 (2.7%) genes of the up- and down-expressed subsets respectively, were associated to cold (Table 2). For example, genes representing a STARCH EXCESS 1 (AT1G10760 [114]), a Glycine-rich RNA-binding protein7 (AT2G21660 [59]), two low temperature and salt-responsive protein LTI6A and LTI6B (AT3G05880 and AT3G05890 [60]), a Late Embryogenesis Abundant 4-5 (AT5G06760 [61]), and a Galactinol synthase 2 (AT1G56600 of the raffinose pathway [62]) were found in the up-expressed subset and are known to present an increase of expression in response to cold condition in A. thaliana. At last, a transcriptional adapter ADA2b (AT4G16420), which may repress freezing tolerance and does not require the expression of CBF or COR genes in A. thaliana [115] was scored in the up-subset. Moreover, we found also in the up-expressed subset genes coding a VERNALIZATION INDEPENDENCE 4 protein (AT5G61150, homologous to LEO1), involved in vernalization response in many A. thaliana ecotypes [134] and a sensitive to freezing 2 protein (AT3G06510, SFR2, a constitutively expressed b-glucosidase), which is conserved in all land plants [63] and involved in the response to freezing by protecting chloroplast membrane from damages.
Overall, concerning kinases and TFs annotations, we observed that most of the annotated kinases in up- and down-expressed subsets are known to be differentially expressed in cold stress [16,18,19,22]. In particular, we identified a gene coding calcium/calmodulin-regulated receptor-like kinase 1 (AT5G54590) in the up-expressed subset that is involved in freezing tolerance in A. thaliana [113], and 5 casein kinases I and 3 SnRK which were involved in stress response [38,135]. Elsewhere, more TFs were scored in the up- than in the down-expressed subset, with seven families being represented in both subsets, and eight and 31 families being specific to down- or up-expressed subsets, respectively. In particular, we identified specifically into the up-expressed subset a MADS protein (AT4G24540, AGAMOUS-LIKE 24 protein), known to be implicated as a transcription activator mediating floral transition in response to vernalization [7]. Here, we report for the first time the involvement of genes coding ARID, DBP, EIL3, mTERF, RWP-RK, SWI, and TAZ proteins in the chilling response.

4.2.2. Champagne Specific Responses to Cold and Acquisition of Freezing Tolerance

As we have seen previously, the specific response to cold concerns a greater variety of genes than the common response. We observed, whatever the modulation of gene expression levels (increase or decrease), many genes presented different level of expression in Ch and Te at the starting point of the time course study (T0). These observations again highlight the importance of the constitutive differences, even in the specific responses of the two lines. Indeed, the second subset (TLSR (b), Figure 3C (b)) gathered up-regulated genes in Ch and in Te during LT, but which are more expressed in Ch, as a result of a higher expression level at T0 for this line. Similarly, genes within the third subset (TLSR (c), Figure 3C (c)) are down-regulated during LT and are less expressed in Ch than in Te during LT as a result of a lower level of expression in Ch at T0. Hence, genes from clusters TLSR (b) and TLSR (c) could be related to the specific responses of Ch taking part in the freezing tolerance capacity of this line. In a same manner, first and forth subsets (Figure 3C (a) and (d)) gathered genes that are related to the specific responses of Te. Considering that these responses (qualitatively and/or quantitatively) are ineffective to bring freezing tolerance to Te, they won’t be discussed below. We noted that only very few genes presented a modulation of expression between the two sampling times (3 and 16 days) during LT, suggesting that at day three, most of the defenses against cold stress are implemented.
Within the TLSR (b) subset, six genes were assigned to GO terms related to cold including a gene coding a pre-mRNA-processing factor 31 homolog (AT1G60170, PRP31), which is involved in the regulation of expression of cold-responsive genes (CORs) in A. thaliana [120]. The PRP31 protein possesses a HAT domain, which is also found in 4 tetratricopeptide repeat protein (AT3G53560, AT2G37400, AT3G46790 and AT3G23020, TPR and/or PRP1) present in this subset [64]. The five other genes code a RNA-binding family protein (AT1G70200, presenting RRM/RBD/RNP motifs), which is implicated in cold tolerance by 23S ribosomal RNA processing in A. thaliana [64], two chloroplastic extern membrane protein 16-1 (AT2G28900) induced in low temperature in land plants [65], a Fibrillin-1a (AT4G04020) which is involved in response to freezing [66], and a Glycine-rich RNA-binding protein 2 (AT4G13850), which confers freezing tolerance after a cold acclimation period [116]. The KEGG annotation revealed that 28.3% of the genes were mapped on other pathways than those related to “Metabolism” and were rather related to RNA and RNA-related pathways. Thus, the RNA metabolism and more particularly RNA-binding with two Pentatricopeptide repeat-containing (AT1G11290 and AT1G20300, PPR motif), one RNA-binding (AT3G08620, KH domain) family, two RNA-binding (AT1G70200 and AT3G20890, PRM/RBD/RNP motifs), one Helicase (AT3G08620, KH domain), and one Glycine-rich containing domain (AT4G13850, RRM domain) proteins in this subset seem to be an important component of the specific cold response in Ch pea line, as well as in A. thaliana [133,136] and Oryza sativa [137].
The post-translational regulation using kinase proteins is considered as a key feature in plant response against cold stress [138]. Among the 11 up-expressed genes coding kinases in the subset TLSR b, two code calcium-dependent protein kinase 6 (AT4G14580 [57] and AT2G17290 [139]) and one a receptor-like protein kinase FERONIA (AT3G51550 [140]) that are involved in stomatal closure control in relation to cold conditions. Moreover, a gene coding a leucine-rich repeat receptor-like serine/threonine-protein kinase implicated in jasmonic acid and ethylene-dependent systemic resistance (AT3G14840 [141]) is scored. In the earlier studies [28,142], the evidence that jasmonate metabolism could play a role in freezing tolerance was provided. TF constitute 9.21% of the genes of this subset. Excepted SWI/SNF-BAF60b, which is signaled in this study for the first time in cold responses, all other are already known to be up or down regulated under cold treatment in legumes species [16,19,22].
The KEGG annotation within the TLSR (c) subset revealed that most of the genes were associated to “Metabolism”. A total of 9 genes were assigned to GO terms related to cold, including genes coding an acyl-CoA-binding domain-containing protein 6 (AT1G31812) that binds phosphatidylcholine in phospholipid metabolism [143], a chlorophyll a/b-binding protein 3-1, chloroplastic (AT1G61520) belonging to the light-harvesting complex in photosystem I [67], two NADPH-dependent aldo-keto reductase, chloroplastic (AT2G37770) that detoxifies a range of toxic aldehydes and ketones produced during stress [68], a bidirectional sugar transporter SWEET17 (AT4G15920) involved in fructose transport [121], and a 3-hydroxyisobutyryl-CoA hydrolase 1 (AT5G65940) that plays a role in peroxisomal metabolism in cold stress signaling and plant tolerance to cold stress, by the degradation of valine [69]. We noticed also the presence of two genes coding a PLAT domain-containing protein 1 (AT4G39730), which functions as positive regulator of abiotic stress tolerance [70], a serine/thereonine protein kinase 2 (AT3G08720), which is a downstream effector of the target of rapamycin signaling pathway (TOR) that presents an increase of protein activity via a phosphorylation induced under cold treatment [71]. All of the TF of TLSR (c) subset were already recorded in cold and freezing tolerance in legumes species, validating once again their participation as key factors of the cold acclimation process in Ch.

5. Conclusions

In most published studies dealing with cold acclimation, gene expression analyses have been realized at the beginning (1 to 3 h) [12,14] or within the 24 h [19] after the acclimation period. Since we look at gene expression after three and sixteen days of acclimation, we expected to identify novel differentially expressed genes. Hence the expressed genes in this work should be involved in subsequent stages of the cold response and more downstream of metabolic chains.
From a total of 4489 differentially expressed genes, we observed the importance of the constitutive differences in gene expression between the two lines. In particular, we identified more preferentially expressed genes related to RNA metabolism in Te, and to protein metabolism in Ch.
Elsewhere, we observed that most of the genes we identified as involved in the freezing tolerance presented similar modulations (activation/repression) in Ch and Te, but with different levels of expression at the beginning of the time course study. This reveals again the importance of initial differences, in the specific responses of the two lines. According to the examination of clusters gathering genes related to chilling and FT, we were surprised to observe less genes modulated by the cold stress in Ch compared to Te. This probably means that Ch presents constitutive and/or induced mechanisms that are more efficient to get over cold than those implemented in Te.
Furthermore, we have identified many TFs, which are linked here for the first time to cold responses. Overall, genes whose expression is for the first time correlated with cold response could open new horizons in the use of genetic diversity of low temperature responses in pea.

Supplementary Materials

The following are available online at https://www.mdpi.com/2223-7747/8/8/288/s1, Figure S1: Workflow for statistical analyses, Figure S2: Functional classification by Gene Ontology, Figure S3: Comparison of the qPCR and the RNA-seq data, Figure S4: GO terms enrichment of Biological Processes and Cellular Component, Table S1: Statistical overview of sequencing and transcriptome assembly data, Table S2: KEGG pathways repartition and classification, Table S3: Kinases repartition in each subset of DEGs.

Author Contributions

N.B., I.L.H., B.D. and S.L. designed the study, contributed to the interpretation of the results and wrote the manuscript. E.H., I.L.H., O.J., F.D., J.-F.H. and N.B. conducted experiments in climatic chamber. N.B. and S.L. conducted the RNA extraction and the bioinformatic analyses. O.B. prepared the NGS libraries and performed RNA sequencing. All authors read and approved the manuscript.

Funding

The funds of this work were provided by the Région Hauts-de-France (Program “projet emergent”) # 2013.1471/7.

Acknowledgments

We thank C. Devaux for her technical expertise and G. Aubert and J. Kreplak for their advices in computing tools. We also thank the high performance computing service of Bilille, University of Lille, for providing computing resources. This work was performed in collaboration with the GeT core facility, Toulouse, France (http://get.genotoul.fr), and was supported by the France Génomique national infrastructure, funded as part of an “Investissement d’avenir” program managed by the Agence Nationale pour la Recherche (contract ANR-10-INBS-09).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pennycooke, J.C.; Cheng, H.; Stockinger, E.J. Comparative genomic sequence and expression analyses of Medicago truncatula and alfalfa subspecies falcata COLD-ACCLIMATION-SPECIFIC genes. Plant Physiol. 2008, 146, 1242–1254. [Google Scholar] [CrossRef] [PubMed]
  2. Carvallo, M.A.; Pino, M.T.; Jeknić, Z.; Zou, C.; Doherty, C.J.; Shiu, S.-H.; Chen, T.H.H.; Thomashow, F. A comparison of the low temperature transcriptomes and CBF regulons of three plant species that differ in freezing tolerance: Solanum commersonii, Solanum tuberosum, and Arabidopsis thaliana. J. Exp. Bot. 2011, 62, 3807–3819. [Google Scholar] [CrossRef] [PubMed]
  3. Tayeh, N.; Bahrman, N.; Devaux, R.; Bluteau, A.; Prosperi, J.-M.; Delbreil, B.; Lejeune-Hénaut, I. A high-density genetic map of the Medicago truncatula major freezing tolerance QTL on chromosome 6 reveals colinearity with a QTL related to freezing damage on Pisum sativum linkage group VI. Mol. Breed. 2013, 32, 279–289. [Google Scholar] [CrossRef]
  4. Zhang, L.L.; Zhao, M.G.; Tian, Q.Y.; Zhang, W.H. Comparative studies on tolerance of Medicago truncatula and Medicago falcata to freezing. Planta 2011, 234, 445–457. [Google Scholar] [CrossRef] [PubMed]
  5. Koornneef, M.; Alonso-Blanco, C.; Vreugdenhil, D. Naturally occurring genetic variation in Arabidopsis thaliana. Ann. Rev. Plant Biol. 2004, 55, 141–172. [Google Scholar] [CrossRef] [PubMed]
  6. Maloof, J.N.; Borevitz, J.O.; Dabi, T.; Lutes, J.; Nehring, R.B.; Redfern, J.L.; Trainer, G.T.; Wilson, J.M.; Asami, T.; Berry, C.C.; et al. Natural variation in light sensitivity of Arabidopsis. Nat. Genet. 2001, 29, 441–446. [Google Scholar] [CrossRef] [PubMed]
  7. Michaels, S.D.; Ditta, G.; Gustafson-Brown, C.; Pelaz, S.; Yanofsky, M.; Amasino, R.M. AGL24 acts as a promoter of flowering in Arabidopsis and is positively regulated by vernalization. Plant J. 2003, 33, 867–874. [Google Scholar] [CrossRef] [PubMed]
  8. Thomashow, M.F. Plant cold acclimation: Freezing tolerance genes and regulatory mechanisms. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1999, 50, 571–599. [Google Scholar] [CrossRef]
  9. Levitt, J. Responses of plants to environmental stresses. In Water, Radiation, Salt, and Other Stresses; Academic Press: Cambridge, MA, USA, 1980; Volume 2. [Google Scholar]
  10. Guy, C.L. Cold acclimation and freezing stress tolerance: Role of protein metabolism. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1990, 41, 187–223. [Google Scholar] [CrossRef]
  11. Guy, C.; Kaplan, F.; Kopka, J.; Selbig, J.; Hincha, D.K. Metabolomics of temperature stress. Physiol. Plant. 2008, 132, 220–235. [Google Scholar] [CrossRef]
  12. Ruelland, E.; Vaultier, M.N.; Zachowski, A.; Hurry, V. Cold Signalling and Cold Acclimation in Plants. Adv. Bot. Res. 2009, 49, 35–150. [Google Scholar] [CrossRef]
  13. Bertrand, A.; Bipfubusa, M.; Claessens, A.; Rocher, S.; Castonguay, Y. Effect of photoperiod prior to cold acclimation on freezing tolerance and carbohydrate metabolism in alfalfa (Medicago sativa L.). Plant Sci. 2017, 264, 122–128. [Google Scholar] [CrossRef] [PubMed]
  14. Kreps, J.A.; Wu, Y.; Chang, H.-S.; Zhu, T.; Wang, X.; Harper, J.F. Transcriptome Changes for Arabidopsis in Response to Salt, Osmotic, and Cold Stress. Plant Physiol. 2002, 130, 2129–2141. [Google Scholar] [CrossRef] [PubMed]
  15. Chinnusamy, V.; Ohta, M.; Kanrar, S.; Lee, B.-H.; Hong, X.; Agarwal, M.; Zhu, J.-K. ICE1: A regulator of cold-induced transcriptome and freezing tolerance in Arabidopsis. Genes Dev. 2003, 17, 1043–1054. [Google Scholar] [CrossRef]
  16. Pang, T.; Ye, C.-Y.; Xia, X.; Yin, W. De novo sequencing and transcriptome analysis of the desert shrub, Ammopiptanthus mongolicus, during cold acclimation using Illumina/Solexa. BMC Genom. 2013, 14, 488. [Google Scholar] [CrossRef]
  17. Kidokoro, S.; Watanabe, K.; Ohori, T.; Moriwaki, T.; Maruyama, K.; Mizoi, J.; Htwe, N.M.P.S.; Fujita, Y.; Sekita, S.; Shinozaki, K.; et al. Soybean DREB1/CBF-type transcription factors function in heat and drought as well as cold stress-responsive gene expression. Plant J. 2015, 81, 505–518. [Google Scholar] [CrossRef]
  18. Calzadilla, P.I.; Maiale, S.J.; Ruiz, O.A.; Escaray, F.J. Transcriptome Response Mediated by Cold Stress in Lotus japonicus. Front. Plant Sci. 2016, 7, 374. [Google Scholar] [CrossRef]
  19. Tan, H.; Huang, H.; Tie, M.; Tang, Y.; Lai, Y.; Li, H. Transcriptome profiling of two Asparagus Bean (Vigna unguiculata subsp. sesquipedalis) cultivars differing in chilling tolerance under cold stress. PLoS ONE 2016, 11, e0151105. [Google Scholar] [CrossRef]
  20. Bonthala, V.S.; Mayes, K.; Moreton, J.; Blythe, M.; Wrigh, T.V.; May, S.T.; Massawe, F.; Mayes, S.; Twycross, J. Identification of Gene Modules Associated with Low Temperatures Response in Bambara Groundnut by Network-Based Analysis. PLoS ONE 2016, 11, e0148771. [Google Scholar] [CrossRef]
  21. Miao, Z.; Xu, W.; Li, D.; Hu, X.; Liu, J.; Zhang, R.; Wang, T. De novo transcriptome analysis of Medicago falcata reveals novel insights about the mechanisms underlying abiotic stress-responsive pathway. BMC Genom. 2015, 16, 818. [Google Scholar] [CrossRef]
  22. Song, L.; Jiang, L.; Chen, Y.; Shu, Y.; Bai, Y.; Guo, C. Deep-sequencing transcriptome analysis of field-grown Medicago sativa L. crown buds acclimated to freezing stress. Funct. Integr. Genom. 2016, 16, 495–511. [Google Scholar] [CrossRef] [PubMed]
  23. Zhou, Q.; Luo, D.; Chai, X.; Wu, Y.; Wang, Y.; Nan, Z.; Yang, Q.; Liu, W.; Liu, Z. Multiple Regulatory Networks Are Activated during Cold Stress in Medicago sativa L. Int. J. Mol. Sci. 2018, 19, 3169. [Google Scholar] [CrossRef] [PubMed]
  24. Garg, R.; Bhattacharjee, A.; Jain, M. Genome-scale transcriptomic insights into molecular aspects of abiotic stress responses in chickpea. Plant Mol. Biol. Rep. 2015, 33, 388–400. [Google Scholar] [CrossRef]
  25. 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]
  26. Lejeune-Hénaut, I.; Delbreil, B.; Devaux, R.; Guilioni, L. Cold temperatures and the functioning of the canopy in pea. In Physiology of the Pea Crop; Science Publishers: Enfield, NH, USA, 2010; pp. 168–181. [Google Scholar]
  27. Lucau-Danila, A.; Toitot, C.; Goulas, E.; Blervacq, A.-S.; Hot, D.; Bahrman, N.; Sellier, H.; Lejeune-Hénaut, I.; Delbreil, B. Transcriptome analysis in pea allows to distinguish chilling and acclimation mechanisms. Plant Physiol. Biochem. 2012, 58, 236–244. [Google Scholar] [CrossRef] [PubMed]
  28. Legrand, S.; Marque, G.; Blassiau, C.; Bluteau, A.; Canoy, A.-S.; Fontaine, V.; Jaminon, O.; Bahrman, N.; Mautord, J.; Morin, J.; et al. Combining gene expression and genetic analyses to identify candidate genes involved in cold responses in pea. J. Plant Physiol. 2013, 170, 1148–1157. [Google Scholar] [CrossRef] [PubMed]
  29. Lejeune-Hénaut, I.; Hanocq, E.; Béthencourt, L.; Fontaine, V.; Delbreil, B.; Morin, J.; Petit, A.; Devaux, R.; Boilleau, M.; Stempniak, J.-J.; et al. The flowering locus Hr colocalizes with a major QTL affecting winter frost tolerance in Pisum sativum L. Theor. Appl. Genet. 2008, 116, 1105–1116. [Google Scholar] [CrossRef]
  30. Bonhomme, R. Bases and limits to using “degree.day” units. Eur. J. Agron. 2000, 13, 1–10. [Google Scholar] [CrossRef]
  31. Andrews, S. FastQC: A Quality Control Tool for High Throughput Sequence Data. 2010. Available online: http://www.bioinformatics.babraham.ac.uk/projects/fastqc/ (accessed on 12 August 2019).
  32. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet. J. 2011, 17, 10–12. [Google Scholar] [CrossRef]
  33. Schmieder, R.; Edwards, R. Quality control and preprocessing of metagenomic datasets. Bioinformatics 2011, 27, 863–864. [Google Scholar] [CrossRef] [Green Version]
  34. Grabherr, M.G.; Haas, B.J.; Yassour, M.; Levin, J.Z.; Thompson, D.A.; Amit, I.; Adiconis, X.; Fan, L.; Raychowdhury, R.; Zeng, Q.; et al. Trinity: Reconstructing a full-length transcriptome without a genome from RNA-Seq data. Nat. Biotechnol. 2011, 29, 644–652. [Google Scholar] [CrossRef] [PubMed]
  35. Smith-Unna, R.; Boursnell, C.; Patro, R.; Hibberd, J.M.; Kelly, S. TransRate: Referance-free quality asssessment of de novo transcriptome assemblies. Genome Res. 2016, 26, 1134–1144. [Google Scholar] [CrossRef] [PubMed]
  36. Gurevich, A.; Saveliev, V.; Vyahhi, N.; Tesler, G. QUAST: Quality assessment tool for genome assemblies. Bioinformatics 2013, 29, 1072–1075. [Google Scholar] [CrossRef] [PubMed]
  37. Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Basic local alignment search tool. J. Mol. Biol. 1990, 215, 403–410. [Google Scholar] [CrossRef]
  38. Li, J.; Liu, H.; Xia, W.; Mu, J.; Feng, Y.; Liu, R.; Yan, P.; Wang, A.; Lin, Z.; Guo, Y.; et al. De Novo Transcriptome Sequencing and the Hypothetical Cold Response Mode of Saussurea involucrata in Extreme Cold Environments. Int. J. Mol. Sci. 2017, 18, 1155. [Google Scholar] [CrossRef] [PubMed]
  39. 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] [PubMed]
  40. Huang, X.; Zhang, Y.; Zhang, X.; Shi, Y. Long-chain base kinase1 affects freezing tolerance in Arabidopsis thaliana. Plant Sci. 2017, 259, 94–103. [Google Scholar] [CrossRef]
  41. Pareek, A.; Khurana, A.; Sharma, A.; Kumar, R. An overview of signaling regulons during cold stress tolerance in plants. Curr. Genom. 2017, 18, 498–511. [Google Scholar] [CrossRef]
  42. Ye, Y.; Ding, Y.; Jiang, Q.; Wang, F.; Sun, J.; Zhu, C. The role of receptor-like protein kinases (RLKs) in abiotic stress response in plants. Plant Cell Rep. 2017, 36, 235–242. [Google Scholar] [CrossRef]
  43. Nakashima, K.; Yamaguchi-Shinozaki, K.; Shinozaki, K. The transcriptional regulatory network in the drought response and its crosstalk in abiotic stress responses including drought, cold, and heat. Front. Plant Sci. 2014, 5, 170. [Google Scholar] [CrossRef] [Green Version]
  44. Huang, G.-T.; Ma, S.-L.; Bai, L.-P.; Zhang, L.; Ma, H.; Jia, P.; Liu, J.; Zhong, M.; Guo, Z.-F. Signal transduction during cold, salt, and drought stresses in plants. Mol. Biol. Rep. 2012, 39, 969–987. [Google Scholar] [CrossRef] [PubMed]
  45. Janská, A.; Maršík, P.; Zelenková, S.; Ovesná, J. Cold stress and acclimation—What is important for metabolic adjustment? Plant Biol. 2010, 12, 395–405. [Google Scholar] [CrossRef] [PubMed]
  46. Miura, K.; Furumoto, T. Cold Signaling and Cold Response in Plants. Int. J. Mol. Sci. 2013, 14, 5312–5337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Zheng, Y.; Jiao, C.; Sun, H.; Rosli, H.G.; Pombo, M.A.; Zhang, P.; Banf, M.; Dai, X.; Martin, G.B.; Giovannoni, J.J.; et al. iTAK: A program for genome-wide prediction and classification of plant transcription factors, transcriptional regulators, and protein kinases. Mol. Plant 2016, 9, 1667–1670. [Google Scholar] [CrossRef] [PubMed]
  48. Pérez-Rodríguez, P.; Riaño-Pachón, D.M.; Corrêa, L.G.G.; Rensing, S.A.; Kersten, B.; Mueller-Roeber, B. PlnTFDB: Updated content and new features of the plant transcription factor database. Nucleic Acids Res. 2010, 38, D822–D827. [Google Scholar] [CrossRef] [PubMed]
  49. Davuluri, R.V.; Sun, H.; Palaniswamy, S.K.; Matthews, N.; Molina, C.; Kurtz, M.; Grotewold, E. AGRIS: Arabidopsis Gene Regulatory Information Server, an information resource of Arabidopsis cis-regulatory elements and transcription factors. BMC Bioinform. 2003, 4, 25. [Google Scholar] [CrossRef]
  50. Bray, N.L.; Pimentel, H.; Melsted, P.; Pachter, L. Near-optimal probabilistic RNA-seq quantification. Nat. Biotechnol. 2016, 34, 525–527. [Google Scholar] [CrossRef] [PubMed]
  51. Love, M.I.; Huber, W.; Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014, 15, 550. [Google Scholar] [CrossRef] [Green Version]
  52. Saeed, A.I.; Sharov, V.; White, J.; Li, J.; Liang, W.; Bhagabati, N.; Braisted, J.; Klapa, M.; Currier, T.; Thiagarajan, M.; et al. TM4: A free, open-source system for microarray data management and analysis. Biotechniques 2003, 34, 374–378. [Google Scholar] [CrossRef]
  53. Kim, Y.; Park, S.; Gilmour, S.J.; Thomashow, M.F. Roles of CAMTA transcription factors and salicylic acid in configuring the low-temperature transcriptome and freezing tolerance of Arabidopsis. Plant J. 2013, 75, 364–376. [Google Scholar] [CrossRef]
  54. Kawamura, Y.; Uemura, M. Mass spectrometric approach for identifying putative plasma membrane proteins of Arabidopsis leaves associated with cold acclimation. Plant J. 2003, 36, 141–154. [Google Scholar] [CrossRef] [PubMed]
  55. Lee, B.-H.; Henderson, D.A.; Zhu, J.-K. The Arabidopsis Cold-Responsive Transcriptome and Its Regulation by ICE. Plant Cell 2005, 17, 3155–3175. [Google Scholar] [CrossRef] [PubMed]
  56. Lee, H.; Guo, Y.; Ohta, M.; Xiong, L.; Stevenson, B.; Zhu, J.-K. LOS2, a genetic locus required for cold-responsive gene transcription encodes a bi-functional enolase. EMBO J. 2002, 21, 2692–2702. [Google Scholar] [CrossRef] [PubMed]
  57. Mizoguchi, T.; Irie, K.; Hirayama, T.; Hayashida, N.; Yamaguchi-Shinozaki, K.; Matsumoto, K.; Shinozaki, K. A gene encoding a mitogen-activated protein kinase kinase kinase is induced simultaneously with genes for a mitogen-activated protein kinase and an S6 ribosomal protein kinase by touch, cold, and water stress in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 1996, 93, 765–769. [Google Scholar] [CrossRef] [PubMed]
  58. Vermel, M.; Guermann, B.; Delage, L.; Grienenberger, J.-M.; Maréchal-Drouard, L.; Gualberto, J.M. A family of RRM-type RNA-binding proteins specific to plant mitochondria. Proc. Natl. Acad. Sci. USA 2002, 99, 5866–5871. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Kwak, K.J.; Kim, Y.O.; Kang, H. Characterization of transgenic Arabidopsis plants overexpressing GR-RBP4 under high salinity, dehydration, or cold stress. J. Exp. Bot. 2005, 56, 3007–3016. [Google Scholar] [CrossRef]
  60. Capel, J.; Jarillo, J.A.; Salinas, J.; Martínez-Zapate, J.M. Two homologous low-temperature-inducible genes from Arabidopsis encode highly hydrophobic proteins. Plant Physiol. 1997, 115, 569–576. [Google Scholar] [CrossRef]
  61. Reyes, J.L.; Campos, F.; Wei, H.; Arora, R.; Yang, Y.; Karlson, D.T.; Covarrubias, A.A. Functional dissection of Hydrophilins during in vitro freeze protection. Plant Cell Environ. 2008, 31, 1781–1790. [Google Scholar] [CrossRef]
  62. Nishizawa, A.; Yabuta, Y.; Shigeoka, S. Galactinol and Raffinose Constitute a Novel Function to Protect Plants from Oxidative Damage. Plant Physiol. 2008, 147, 1251–1263. [Google Scholar] [CrossRef] [Green Version]
  63. Fourrier, N.; Bédard, J.; Lopez-Juez, E.; Barbrook, A.; Bowyer, J.; Jarvis, P.; Warren, G.; Thorlby, G. A role for sensitive to freezing2 in protecting chloroplasts against freeze-induced damage in Arabidopsis. Plant J. 2008, 55, 734–745. [Google Scholar] [CrossRef]
  64. Wang, S.; Bai, G.; Wang, S.; Yang, L.; Yang, F.; Wang, Y.; Zhu, J.-K.; Hua, J. Chloroplast RNA-Binding Protein RBD1 Promotes Chilling Tolerance through 23S rRNA Processing in Arabidopsis. PLoS Genet. 2016, 12, e1006027. [Google Scholar] [CrossRef] [PubMed]
  65. Drea, S.C.; Lao, N.T.; Wolfe, K.H.; Kavanagh, T.A. Gene duplication, exon gain and neofunctionalization of OEP16-related genes in land plants. Plant J. 2006, 46, 723–735. [Google Scholar] [CrossRef] [PubMed]
  66. Goulas, E.; Schubert, M.; Kieselbach, T.; Kleczkowski, L.A.; Gardeström, P.; Schröder, W.; Hurry, V. The chloroplast lumen and stromal proteomes of Arabidopsis thaliana show differential sensitivity to short- and long-term exposure to low temperature. Plant J. 2006, 47, 720–734. [Google Scholar] [CrossRef] [PubMed]
  67. Ganeteg, U.; Klimmek, F.; Jansson, S. Lhca5—An LHC-type protein associated with photosystem I. Plant Mol. Biol. 2004, 54, 641–651. [Google Scholar] [CrossRef] [PubMed]
  68. Simpson, P.J.; Tantitadapitak, C.; Reed, A.M.; Mather, O.C.; Bunce, C.M.; White, S.A.; Ride, J.P. Characterization of Two Novel Aldo–Keto Reductases from Arabidopsis: Expression Patterns, Broad Substrate Specificity, and an Open Active-Site Structure Suggest a Role in Toxicant Metabolism Following Stress. J. Mol. Biol. 2009, 392, 465–480. [Google Scholar] [CrossRef] [PubMed]
  69. Dong, C.H.; Zolman, B.K.; Bartel, B.; Lee, B.H.; Stevenson, B.; Agarwal, M.; Zhu, J.-K. Disruption of Arabidopsis CHY1 reveals an important role of metabolic status in plant cold stress signaling. Mol. Plant 2009, 2, 59–72. [Google Scholar] [CrossRef] [PubMed]
  70. Hyun, T.K.; Van Der Graaff, E.; Albacete, A.; Eom, S.H.; Großkinsky, D.K.; Böhm, H.; Janschek, U.; Rim, Y.; Ali, W.W.; Kim, S.Y.; et al. The Arabidopsis PLAT Domain Protein1 Is Critically Involved in Abiotic Stress Tolerance. PLoS ONE 2014, 9, e112946. [Google Scholar] [CrossRef]
  71. Mizoguchi, T.; Hayashida, N.; Yamaguchi-Shinozaki, K.; Kamada, H.; Shinozaki, K. Two genes that encode ribosomal-protein S6 kinase homologs are induced by cold or salinity stress in Arabidopsis thaliana. FEBS Lett. 1995, 358, 199–204. [Google Scholar] [CrossRef]
  72. Amme, S.; Matros, A.; Schlesier, B.; Mock, H.-P. Proteome analysis of cold stress response in Arabidopsis thaliana using DIGE-technology. J. Exp. Bot. 2006, 57, 1537–1546. [Google Scholar] [CrossRef]
  73. Joubès, J.; Raffaele, S.; Bourdenx, B.; Garcia, C.; Laroche-Traineau, J.; Moreau, P.; Domergue, F.; Lessire, R. The VLCFA elongase gene family in Arabidopsis thaliana: Phylogenetic analysis, 3D modelling and expression profiling. Plant Mol. Biol. 2008, 67, 547–566. [Google Scholar] [CrossRef]
  74. Du, Z.-Y.; Xiao, S.; Chen, Q.-F.; Chye, M.-L. Depletion of the Membrane-Associated Acyl-Coenzyme A-Binding Protein ACBP1 Enhances the Ability of Cold Acclimation in Arabidopsis. Plant Physiol. 2010, 152, 1585–1597. [Google Scholar] [CrossRef] [PubMed]
  75. Sheldon, C.C.; Finnegan, E.J.; Dennis, E.S.; Peacock, W.J. Quantitative effects of vernalization on FLC and SOC1 expression. Plant J. 2006, 45, 871–883. [Google Scholar] [CrossRef] [PubMed]
  76. Jarillo, J.A.; Leyva, A.; Salinas, J.; Martinez-Zapater, J.M. Low Temperature lnduces the Accumulation of Alcohol Dehydrogenase mRNA in Arabidopsis thaliana, a Chilling-Tolerant Plant. Plant Physiol. 1993, 101, 833–837. [Google Scholar] [CrossRef] [PubMed]
  77. Cantero, A.; Barthakur, S.; Bushart, T.J.; Chou, S.; Morgan, R.O.; Fernandez, M.P.; Clark, G.B.; Roux, S.J. Expression profiling of the Arabidopsis annexin gene family during germination, de-etiolation and abiotic stress. Plant Physiol. Biochem. 2006, 44, 13–24. [Google Scholar] [CrossRef] [PubMed]
  78. Doukhanina, E.V.; Chen, S.; Van Der Zalm, E.; Godzik, A.; Reed, J.; Dickman, M.B. Identification and Functional Characterization of the BAG Protein Family in Arabidopsis thaliana. J. Biol. Chem. 2006, 281, 18793–18801. [Google Scholar] [CrossRef] [PubMed]
  79. Bae, M.S.; Cho, E.J.; Choi, E.-Y.; Park, O.K. Analysis of the Arabidopsis nuclear proteome and its response to cold stress. Plant J. 2003, 36, 652–663. [Google Scholar] [CrossRef] [PubMed]
  80. Yang, T.; Chaudhuri, S.; Yang, L.; Chen, Y.; Poovaiah, B.W. Calcium/Calmodulin Up-regulates a Cytoplasmic Receptor-like Kinase in Plants. J. Biol. Chem. 2004, 279, 42552–42559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Pandey, G.K.; Cheong, Y.H.; Kim, B.-G.; Grant, J.J.; Li, L.; Luan, S. CIPK9: A calcium sensor-interacting protein kinase required for low-potassium tolerance in Arabidopsis. Cell Res. 2007, 17, 411–421. [Google Scholar] [CrossRef] [PubMed]
  82. Kolesinski, P.; Piechota, J.; Szczepaniak, A. Initial characteristics of RbcX proteins from Arabidopsis thaliana. Plant Mol. Biol. 2011, 77, 447–459. [Google Scholar] [CrossRef] [PubMed]
  83. Carpenter, C.D.; Kreps, J.A.; Simon, A.E. Genes encoding Glycine - Rich Arabidopsis thaliana proteins with RNA-Binding motifs are influenced by cold treatment and an endogenous circadian rhythm. Plant Physiol. 1994, 104, 1015–1025. [Google Scholar] [CrossRef] [PubMed]
  84. Li, X.; Ma, D.; Lu, S.-X.; Hu, X.; Huang, R.; Liang, T.; Xu, T.; Tobin, E.M.; Liu, H. Blue Light-and Low Temperature-Regulated COR27 and COR28 Play Roles in the Arabidopsis Circadian Clock. Plant Cell 2016, 28, 2755–2769. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, X.; Liu, S.; Takano, T. Two cysteine proteinase inhibitors from Arabidopsis thaliana, AtCYSa and AtCYSb, increasing the salt, drought, oxidation and cold tolerance. Plant Mol. Biol. 2008, 68, 131–143. [Google Scholar] [CrossRef] [PubMed]
  86. Gong, Z.; Lee, H.; Xiong, L.; Jagendorf, A.; Stevenson, B.; Zhu, J.-K. RNA helicase-like protein as an early regulator of transcription factors for plant chilling and freezing tolerance. Proc. Natl. Acad. Sci. USA 2002, 99, 11507–11512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Gómez-Merino, F.C.; Brearley, C.A.; Ornatowska, M.; Abdel-Haliem, M.E.F.; Zanor, M.-I.; Mueller-Roeber, B. AtDGK2, a Novel Diacylglycerol Kinase from Arabidopsis thaliana, phosphorylates 1-Stearoyl-2-arachidonoyl-sn-glycerol and 1, 2-Dioleoyl-sn-glycerol and Exhibits Cold-inducible Gene Expression. J. Biol. Chem. 2004, 279, 8230–8241. [Google Scholar] [CrossRef] [PubMed]
  88. Dong, C.-H.; Agarwal, M.; Zhang, Y.; Xie, Q.; Zhu, J.-K. The negative regulator of plant cold responses, HOS1, is a RING E3 ligase that mediates the ubiquitination and degradation of ICE1. Proc. Natl. Acad. Sci. USA 2006, 103, 8281–8286. [Google Scholar] [CrossRef] [Green Version]
  89. Vergnolle, C.; Vaultier, M.-N.; Taconnat, L.; Renou, J.-P.; Kader, J.-C.; Zachowski, A.; Ruelland, E. The Cold-Induced Early Activation of Phospholipase C and D Pathways Determines the Response of Two Distinct Clusters of Genes in Arabidopsis Cell Suspensions. Plant Physiol. 2005, 139, 1217–1233. [Google Scholar] [CrossRef] [PubMed]
  90. Rae, L.; Lao, N.T.; Kavanagh, T.A. Regulation of multiple aquaporin genes in Arabidopsis by a pair of recently duplicated DREB transcription factors. Planta 2011, 234, 429–444. [Google Scholar] [CrossRef]
  91. Kim, Y.-O.; Kim, J.S.; Kang, H. Cold-inducible zinc finger-containing glycine-rich RNA-binding protein contributes to the enhancement of freezing tolerance in Arabidopsis thaliana. Plant J. 2005, 42, 890–900. [Google Scholar] [CrossRef]
  92. Chen, C.-N.; Chu, C.-C.; Zentella, R.; Pan, S.-M.; Ho, T.-H.D. AtHVA22 gene family in Arabidopsis: Phylogenetic relationship, ABA and stress regulation, and tissue-specific expression. Plant Mol. Biol. 2002, 49, 633–644. [Google Scholar]
  93. Torabinejad, J.; Donahue, J.L.; Gunesekera, B.N.; Allen-Daniels, M.J.; Glenda, E.; Gillaspy, G.E. VTC4 Is a Bifunctional Enzyme That Affects Myoinositol and Ascorbate Biosynthesis in Plants. Plant Physiol. 2009, 150, 951–961. [Google Scholar] [CrossRef] [Green Version]
  94. Xu, D.; Huang, X.; Xu, Z.-Q.; Schläppi, M. The HyPRP gene EARLI1 has an auxiliary role for germinability and early seedling development under low temperature and salt stress conditions in Arabidopsis thaliana. Planta 2011, 234, 565–577. [Google Scholar] [CrossRef] [PubMed]
  95. Nordin, K.; Vahala, T.; Palva, E.T. Differential expression of two related, low-temperature-induced genes in Arabidopsis thaliana (L.) Heynh. Plant Mol. Biol. 1993, 21, 641–653. [Google Scholar] [CrossRef] [PubMed]
  96. Li, H.; Ding, Y.; Shi, Y.; Zhang, X.; Zhang, S.; Gong, Z.; Yang, S. MPK3-and MPK6-Mediated ICE1 Phosphorylation Negatively Regulates ICE1 Stability and Freezing Tolerance in Arabidopsis. Dev. Cell 2017, 43, 630–642. [Google Scholar] [CrossRef] [PubMed]
  97. Dong, M.A.; Farré, E.M.; Thomashow, M.F. Circadian clock-associated 1 and late elongated HYPOCOTYL regulate expression of the c-repeat binding factor (CBF) pathway in Arabidopsis. Proc. Natl. Acad. Sci. USA 2011, 108, 7241–7246. [Google Scholar] [CrossRef] [PubMed]
  98. Hirayama, T.; Ohto, C.; MizoguchiI, T.; Shinozaki, K. A gene encoding a phosphatidylinositol-specific phospholipase C is induced by dehydration and salt stress in Arabidopsis thaliana. Proc. Natl. Acad. Sci. USA 1995, 92, 3903–3907. [Google Scholar] [CrossRef] [PubMed]
  99. Li, W.; Li, M.; Zhang, W.; Welti, R.; Wang, X. The plasma membrane–bound phospholipase Dδ enhances freezing tolerance in Arabidopsis thaliana. Nat. Biotechnol. 2004, 22, 427–433. [Google Scholar] [CrossRef] [PubMed]
  100. Covington, M.F.; Panda, S.; Liu, X.L.; Strayer, C.A.; Wagner, D.R.; Kay, S.A. ELF3 Modulates Resetting of the Circadian Clock in Arabidopsis. Plant Cell 2001, 13, 1305–1315. [Google Scholar] [CrossRef]
  101. Xin, Z.; Mandaokar, A.; Chen, J.; Last, R.L.; Browse, J. Arabidopsis ESK1 encodes a novel regulator of freezing tolerance. Plant J. 2007, 49, 786–799. [Google Scholar] [CrossRef]
  102. Cao, S.; Ye, M.; Jiang, S. Involvement of GIGANTEA gene in the regulation of the cold stress response in Arabidopsis. Plant Cell Rep. 2005, 24, 683–690. [Google Scholar] [CrossRef]
  103. Salleh, F.M.; Evans, K.; Goodall, B.; Machin, H.; Mowla, S.B.; Mur, L.A.J.; Runions, J.; Theodoulou, F.L.; Foyer, C.H.; Rogers, H.J. A novel function for a redox-related LEA protein (SAG21/AtLEA5) in root development and biotic stress responses. Plant Cell Environ. 2012, 35, 418–429. [Google Scholar] [CrossRef]
  104. Maruyama, K.; Takeda, M.; Kidokoro, S.; Yamada, K.; Sakuma, Y.; Urano, K.; Fujita, M.; Yoshiwara, K.; Matsukura, S.; Morishita, Y.; et al. Metabolic Pathways Involved in Cold Acclimation Identified by Integrated Analysis of Metabolites and Transcripts Regulated by DREB1A and DREB2A. Plant Physiol. 2009, 150, 1972–1980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Yamazaki, T.; Kawamura, Y.; Minami, A.; Uemura, M. Calcium-Dependent Freezing Tolerance in Arabidopsis Involves Membrane Resealing via Synaptotagmin SYT1. Plant Cell 2008, 20, 3389–3404. [Google Scholar] [CrossRef] [PubMed]
  106. Kim, M.J.; Shin, R.; Schachtman, D.P. A Nuclear Factor Regulates Abscisic Acid Responses in Arabidopsis. Plant Physiol. 2009, 151, 1433–1445. [Google Scholar] [CrossRef] [PubMed]
  107. Chu, B.; Snustad, D.P.; Carter, J.V. Alteration of β-Tubulin Gene Expression during Low-Temperature Exposure in Leaves of Arabidopsis thaliana. Plant Physiol. 1993, 103, 371–377. [Google Scholar] [CrossRef] [PubMed]
  108. Jiang, Y.; Deyholos, M.K. Functional characterization of Arabidopsis NaCl-inducible WRKY25 and WRKY33 transcription factors in abiotic stresses. Plant Mol. Biol. 2009, 69, 91–105. [Google Scholar] [CrossRef] [PubMed]
  109. Polisensky, D.H.; Braam, J. Cold-Shock Regulation of the Arabidopsis TCH Genes and the Effects of Modulating lntracellular Calcium Levels. Plant Physiol. 1996, 111, 1271–1279. [Google Scholar] [CrossRef] [PubMed]
  110. Okawa, K.; Nakayama, K.; Kakizaki, T.; Yamashita, T.; Inaba, T. Identification and characterization of Cor413im proteins as novel components of the chloroplast inner envelope. Plant Cell Environ. 2008, 31, 1470–1483. [Google Scholar] [CrossRef] [PubMed]
  111. Carlsson, A.S.; LaBrie, S.T.; Kinney, A.J.; Von Wettstein-Knowles, P.; Browse, J. A KAS2 cDNA complements the phenotypes of the Arabidopsis fab1 mutant that differs in a single residue bordering the substrate binding pocket. Plant J. 2002, 29, 761–770. [Google Scholar] [CrossRef] [PubMed]
  112. Hemsley, P.A.; Hurst, C.H.; Kaliyadasa, E.; Lamb, R.; Knight, M.R.; De Cothi, E.A.; Steele, J.F.; Knight, H. The Arabidopsis Mediator Complex Subunits MED16, MED14, and MED2 Regulate Mediator and RNA Polymerase II Recruitment to CBF-Responsive Cold-Regulated Genes. Plant Cell 2014, 26, 465–484. [Google Scholar] [CrossRef]
  113. Furuya, T.; Matsuoka, D.; Nanmori, T. Phosphorylation of Arabidopsis thaliana MEKK1 via Ca2+ signaling as a part of the cold stress response. J. Plant Res. 2013, 126, 833–840. [Google Scholar] [CrossRef]
  114. Yano, R.; Nakamura, M.; Yoneyama, T.; Nishida, I. Starch-Related a-Glucan/Water Dikinase Is Involved in the Cold-Induced Development of Freezing Tolerance in Arabidopsis. Plant Physiol. 2005, 138, 837–846. [Google Scholar] [CrossRef] [PubMed]
  115. Vlachonasios, K.E.; Thomashow, M.F.; Triezenberg, S.J. Disruption Mutations of ADA2b and GCN5 Transcriptional Adaptor Genes Dramatically Affect Arabidopsis Growth, Development, and Gene Expression. Plant Cell 2003, 15, 626–638. [Google Scholar] [CrossRef] [PubMed]
  116. Kim, J.Y.; Park, S.J.; Jang, B.; Jung, C.-H.; Ahn, S.J.; Goh, C.-H.; Cho, K.; Han, O.; Kang, H. Functional characterization of a glycine-rich RNA-binding protein 2 in Arabidopsis thaliana under abiotic stress conditions. Plant J. 2007, 50, 439–451. [Google Scholar] [CrossRef] [PubMed]
  117. Roudier, F.; Gissot, L.; Beaudoin, F.; Haslam, R.; Michaelson, L.; Marion, J.; Molino, D.; Lima, A.; Bach, L.; Morin, H.; et al. Very-Long-Chain Fatty Acids Are Involved in Polar Auxin Transport and Developmental Patterning in Arabidopsis. Plant Cell 2010, 22, 364–375. [Google Scholar] [CrossRef] [PubMed]
  118. Novillo, F.; Medina, J.; Salinas, J. Arabidopsis CBF1 and CBF3 have a different function than CBF2 in cold acclimation and define different gene classes in the CBF regulon. Proc. Natl. Acad. Sci. USA 2007, 104, 21002–21007. [Google Scholar] [CrossRef] [PubMed]
  119. Nakai, Y.; Nakahira, Y.; Sumida, H.; Takebayashi, K.; Nagasawa, Y.; Yamasaki, K.; Akiyama, M.; Ohme-Takagi, M.; Fujiwara, S.; Shiina, T.; et al. Vascular plant one-zinc-finger protein 1/2 transcription factors regulate abiotic and biotic stress responses in Arabidopsis. Plant J. 2013, 73, 761–775. [Google Scholar] [CrossRef] [PubMed]
  120. Du, J.-L.; Zhang, S.-W.; Huang, H.-W.; Cai, T.; Li, L.; Chen, S.; He, X.-J. The Splicing Factor PRP31 Is Involved in Transcriptional Gene Silencing and Stress Response in Arabidopsis. Mol. Plant 2015, 8, 1053–1068. [Google Scholar] [CrossRef] [PubMed]
  121. Chardon, F.; Bedu, M.; Calenge, F.; Klemens, P.A.W.; Spinner, L.; Clement, G.; Chietera, G.; Léran, S.; Ferrand, G.; Lacombe, B.; et al. Leaf Fructose Content Is Controlled by the Vacuolar Transporter SWEET17 in Arabidopsis. Curr. Biol. 2013, 23, 697–702. [Google Scholar] [CrossRef] [PubMed]
  122. Chen, M.; Markham, J.E.; Cahoon, E.B. Sphingolipid Δ8 unsaturation is important for glucosylceramide biosynthesis and low-temperature performance in Arabidopsis. Plant J. 2012, 69, 769–781. [Google Scholar] [CrossRef] [PubMed]
  123. Meyerhoff, O.; Müller, K.; Roelfsema, M.R.G.; Latz, A.; Lacombe, B.; Hedrich, R.; Dietrich, P.; Becker, D. AtGLR3.4, a glutamate receptor channel-like gene is sensitive to touch and cold. Planta 2005, 222, 418–427. [Google Scholar] [CrossRef]
  124. Küfner, I.; Koch, W. Stress regulated members of the plant organic cation transporter family are localized to the vacuolar membrane. BMC Res. Notes 2008, 1, 43. [Google Scholar] [CrossRef] [PubMed]
  125. Schlaen, R.G.; Mancini, E.; Sanchez, S.E.; Perez-Santangelo, S.; Rugnone, M.L.; Simpson, C.G.; Brown, J.W.S.; Zhang, X.; Chernomoretz, A.; Yanovsky, M.J. The spliceosome assembly factor GEMIN2 attenuates the effects of temperature on alternative splicing and circadian rhythms. Proc. Natl. Acad. Sci. USA 2015, 112, 9382–9387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Dumont, E.; Fontaine, V.; Vuylsteker, C.; Sellier, H.; Bodèle, S.; Voedts, N.; Devaux, R.; Frise, M.; Avia, K.; Hilbert, J.-L.; et al. Association of sugar content QTL and PQL with physiological traits relevant to frost damage resistance in pea under field and controlled conditions. Theor. Appl. Genet. 2009, 118, 1561–1571. [Google Scholar] [CrossRef] [PubMed]
  127. Schönrock, N.; Bouveret, R.; Leroy, O.; Borghi, L.; Köhler, C.; Gruissem, W.; Hennig, L. Polycomb-group proteins repress the floral activator AGL19 in the FLC-independent vernalization pathway. Genes Dev. 2006, 20, 1667–1678. [Google Scholar] [CrossRef] [PubMed]
  128. Dumont, E.; Bahrman, N.; Goulas, E.; Valot, B.; Sellier, H.; Hilbert, J.-L.; Vuylsteker, C.; Lejeune-Hénaut, I.; Delbreil, B. A proteomic approach to decipher chilling response from cold acclimation in pea (Pisum sativum L.). Plant Sci. 2011, 180, 86–98. [Google Scholar] [CrossRef] [PubMed]
  129. Thomashow, M.F. Role of cold-responsive genes in plant freezing tolerance. Plant Physiol. 1998, 118, 1–8. [Google Scholar] [CrossRef] [PubMed]
  130. Siol, M.; Jacquin, F.; Chabert-Martinello, M.; Smýkal, P.; Le Paslier, M.-C.; Aubert, G.; Burstin, J. Patterns of Genetic Structure and Linkage Disequilibrium in a Large Collection of Pea Germplasm. Genes Genomes Genet. 2017, 7, 2461–2471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Grimaud, F.; Renaut, J.; Dumont, E.; Sergeant, K.; Lucau-Danila, A.; Blervacq, A.-S.; Sellier, H.; Bahrman, N.; Lejeune-Hénaut, I.; Delbreil, B.; et al. Exploring chloroplastic changes related to chilling and freezing tolerance during cold acclimation of pea (Pisum sativum L.). J. Proteom. 2013, 80, 145–159. [Google Scholar] [CrossRef]
  132. Liu, X.; Zhou, Y.; Xiao, J.; Bao, F. Effects of Chilling on the Structure, Function and Development of Chloroplasts. Front. Plant Sci. 2018, 9, 1715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Ambrosone, A.; Costa, A.; Leone, A.; Grillo, S. Beyond transcription: RNA-binding proteins as emerging regulators of plant response to environmental constraints. Plant Sci. 2012, 182, 12–18. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, H.; Van Nocker, S. The Vernalization Independence 4 gene encodes a novel regulator of flowering locus C. Plant J. 2002, 31, 663–673. [Google Scholar] [CrossRef] [PubMed]
  135. Hey, S.J.; Byrne, E.; Halford, N.G. The interface between metabolic and stress signaling. Ann. Bot. 2010, 105, 197–203. [Google Scholar] [CrossRef] [PubMed]
  136. Lee, B.-H.; Kapoor, A.; Zhu, J.; Zhu, J.-K. STABILIZED1, a Stress-Upregulated Nuclear Protein, Is Required for Pre-mRNA Splicing, mRNA Turnover, and Stress Tolerance in Arabidopsis. Plant Cell 2006, 18, 1736–1749. [Google Scholar] [CrossRef] [PubMed]
  137. Kim, J.Y.; Kim, W.Y.; Kwak, K.J.; Oh, S.H.; Han, Y.S.; Kang, H. Zinc finger-containing glycine-rich RNA-binding protein in Oryza sativa has an RNA chaperone activity under cold stress conditions. Plant Cell Environ. 2010, 33, 759–768. [Google Scholar] [PubMed]
  138. Benschop, J.J.; Mohammed, S.; O’Flaherty, M.; Heck, A.J.R.; Slijper, M.; Menke, F.L.H. Quantitative Phosphoproteomics of Early Elicitor Signaling in Arabidopsis. Mol. Cell. Proteom. 2007, 6, 1198–1214. [Google Scholar] [CrossRef] [PubMed]
  139. Mori, I.C.; Murata, Y.; Yang, Y.; Munemasa, S.; Wang, Y.F.; Andreoli, S.; Tiriac, H.; Alonso, J.M.; Harper, J.F.; Ecker, J.R.; et al. CDPKs CPK6 and CPK3 function in ABA regulation of guard cell S-type anion- and Ca 2þ -permeable channels and stomatal closure. PLoS Biol. 2006, 4, e327. [Google Scholar] [CrossRef]
  140. Yu, Y.; Chakravorty, D.; Assmann, S.M. The G Protein b-Subunit, AGB1, interacts with FERONIA in RALF1-Regulated St; omatal Movement. Plant Physiol. 2018, 176, 2426–2440. [Google Scholar] [CrossRef]
  141. Le, M.H.; Cao, Y.; Zhang, X.-C.; Stacey, G. LIK1, A CERK1-Interacting Kinase, Regulates Plant Immune Responses in Arabidopsis. PLoS ONE 2014, 9, e102245. [Google Scholar] [CrossRef]
  142. Hu, Y.; Jiang, Y.; Han, X.; Wang, H.; Pan, J.; Yu, D. Jasmonate regulates leaf senescence and tolerance to cold stress: Crosstalk with other phytohormones. J. Exp. Bot. 2017, 68, 1361–1369. [Google Scholar] [CrossRef]
  143. Chen, Q.-F.; Xiao, S.; Chye, M.-L. Overexpression of the Arabidopsis 10-Kilodalton Acyl-Coenzyme A-Binding Protein ACBP6 Enhances Freezing Tolerance. Plant Physiol. 2008, 148, 304–315. [Google Scholar] [CrossRef]
Figure 1. Scheme of the experiments and samplings. RNA-sequencing (RNA-seq) experiments (A): low temperature (LT) treatment and (B): control (N), nursery in magenta, low temperature in cyan, freezing in sky blue, and recovery period in orange; quantitative polymerase chain reaction (qPCR) experiments from Reference [28], as mentioned in § 2.5, (C): LT and (D): N, color codes as in RNA-seq. The sampling dates are in green and the numbers refer to the days.
Figure 1. Scheme of the experiments and samplings. RNA-sequencing (RNA-seq) experiments (A): low temperature (LT) treatment and (B): control (N), nursery in magenta, low temperature in cyan, freezing in sky blue, and recovery period in orange; quantitative polymerase chain reaction (qPCR) experiments from Reference [28], as mentioned in § 2.5, (C): LT and (D): N, color codes as in RNA-seq. The sampling dates are in green and the numbers refer to the days.
Plants 08 00288 g001
Figure 2. Scatter plot between the two sets of transcripts issued from the present RNA-seq and qPCR data from Reference [28]. The x-axis corresponds to log2 of qPCR ratios and the y-axis shows log2 of RNA-seq ratios for 33 transcripts. The linear relationship between the 2 variables and their correlation coefficient, R = ∑xy/√ ∑x2 ∑y2 are reported in the top of the graph.
Figure 2. Scatter plot between the two sets of transcripts issued from the present RNA-seq and qPCR data from Reference [28]. The x-axis corresponds to log2 of qPCR ratios and the y-axis shows log2 of RNA-seq ratios for 33 transcripts. The linear relationship between the 2 variables and their correlation coefficient, R = ∑xy/√ ∑x2 ∑y2 are reported in the top of the graph.
Plants 08 00288 g002
Figure 3. Hierarchical clustering of the three sets of differentially expressed genes. (A): “Line Response” containing two subsets, a (Ch up) and b (Te up); (B): “Temperature Common Response” having two subsets, a (TCR down) and b (TCR up) and (C): “Temperature Line Specific Response” including four subsets, TLSR a, b, c, and d. The order of the columns from left to right are ChNT0, ChNT1, ChNT2, ChLT0, ChLT1, ChLT2, TeNT0, TeNT1, TeNT2, TeLT0, TeLT1, and TeLT2.
Figure 3. Hierarchical clustering of the three sets of differentially expressed genes. (A): “Line Response” containing two subsets, a (Ch up) and b (Te up); (B): “Temperature Common Response” having two subsets, a (TCR down) and b (TCR up) and (C): “Temperature Line Specific Response” including four subsets, TLSR a, b, c, and d. The order of the columns from left to right are ChNT0, ChNT1, ChNT2, ChLT0, ChLT1, ChLT2, TeNT0, TeNT1, TeNT2, TeLT0, TeLT1, and TeLT2.
Plants 08 00288 g003
Figure 4. Transcription factors and regulators (TFs) compiled from iTAK data and AGRIS (Arabidopsis Gene Regulatory Information Server) (AtTFDB) in each subset of DEGs.
Figure 4. Transcription factors and regulators (TFs) compiled from iTAK data and AGRIS (Arabidopsis Gene Regulatory Information Server) (AtTFDB) in each subset of DEGs.
Plants 08 00288 g004
Table 1. Summary of the functional annotation.
Table 1. Summary of the functional annotation.
SetSubset Description# Genes5 Most Significant Enriched GO Terms Related to Biological Process5 Most Significant Enriched GO Terms Related to Cellular Localization Kegg Pathways: # Mapped Genes (%), # Represented Pathways, Most Represented Pathways (#)iTak: # Kinases (%), # FamiliesTranscription Factors: # TF (%), # Families, Most Represented Family (#)GO Terms Related to Cold: # Genes
Line response (LR): genes differentially expressed between Ch and Te at T0 and not modulated under LTGenes more expressed in Ch than in Te
(Figure 3A a)
906metabolic process, cellular process, localization, cellular component organization or biogenesis, response to stimulusplastid, chloroplast cytoplasmic part, cell part, cell553 (61.0%), 110, Cysteine and methionine metabolism (13)26 (3.0%), 1250 (5.5%),
26,
AP2-EREBP (6)
-response to cold GO:0009409: 23
-cold acclimation GO:0009631: 4
-cellular response to cold GO:0070417: 2
-total: 29 (3.2%)
Genes more expressed in Te than in Ch(Figure 3A b)1581cellular process, metabolic process, localization, cellular component organization or biogenesis, response to stimuluscell part, cell, intracellular part, intracellular and intracellular organelle616 (39.0%), 106, Ribosome (25)70 (4.4%), 19138 (8.7%)
43,
bHLH (13)
-response to cold GO:0009409: 35
-cellular response to cold GO:0070417: 1
-total: 36 (2.3%)
Temperature Common Response (TCR): Genes responding identically in Ch and Te under LTGenes down expressed during LT
(Figure 3B a)
520metabolic process, cellular process, cellular component organization or biogenesis, localization, response to stimuluschloroplast, plastid, chloroplast part, plastid part, cytoplasm366 (70.4%), 77, Glyoxylate and dicarboxylate metabolism (10)16 (3.1%),
12
25 (4.8%),
16,
bHLH (4)
-response to cold GO:0009409: 14
-Total: 14 (2.7%)
Genes up expressed during LT
(Figure 3B b)
883metabolic process, cellular process, cellular component organization or biogenesis, localization, response to stimulusintracellular, intracellular part, cell part, cell, membrane-bounded organelle388 (43.9%), 88, Ribosome biogenesis (21)38 (4.3%),
22
89 (10.1%),
38,
bHLH (7), C2C2-CO-like (7)
-response to cold GO:0009409: 25
-cold acclimation GO:0009631: 3
-cellular response to cold GO:0070417: 2
-total: 30 (3.4%)
Temperature Line Specific Response (TLSR): genes responding differentially in Ch and Te under LTGenes more expressed in Ch than in Te and down regulated during LT
(Figure 3C a)
253metabolic process, cellular process, localization, cellular component organization or biogenesis, response to stimuluscytoplasmic part, plastid, chloroplast, cytoplasm, intracellular organelle part201 (79.5%), 73, Oxidative phosphorylation (6), Starch and sucrose metabolism (6)7 (2.8%),
6
16 (6.3%),
10,
C3H (3), MYB (3)
-response to cold GO:0009409: 8
-cold acclimation GO:0009631: 1
-total: 9 (3.6%)
Genes more expressed in Ch than in Te and up regulated during LT
(Figure 3C b)
228cellular process, metabolic process, localization, response to stimulus, cellular component organization or biogenesiscell part, cell, intracellular part, cytoplasm, intracellular170 (74.6%), 62, Purine metabolism (5)11 (4.8%), 1021 (9.2%),
15,
AP2-EREBP (3)
-response to cold GO:0009409: 4
-cold acclimation GO:0009631: 1
-cellular response to cold GO:0070417: 1
-total: 6 (2.6%)
Genes less expressed in Ch than in Te and down regulated during LT
(Figure 3C c)
131metabolic process, cellular process, localization, response to stimulus, cellular component organization or biogenesiscytoplasmic part57 (43.5%), 34, several pathways represented by 2 genes2 (1.5%),
2
12 (9.2%),
11, AP2-EREBP (2)
-response to cold GO:0009409: 8
-cold acclimation GO:0009631: 1
-total: 9 (6.9%)
Genes less expressed in Ch than in Te and up regulated during LT
(Figure 3C d)
479metabolic process, cellular process, localization, cellular component organization or biogenesis, response to stimuluscell part, cell, intracellular part, intracellular, nucleus173 (36.1%), 71, mRNA surveillance pathway (7)23 (4.8%),
16
46 (9.6%),
25,
C2H2 (5)
-response to cold GO:0009409: 11
-cold acclimation GO:0009631: 1
-total: 12 (2.5%)
Ch: Champagne; LT: low temperature treatment; Te: Térèse; TF: transcription factor.
Table 2. GO matching of the cold response (i.e., “response to cold”, “cold acclimation”, “cellular response to cold”) in eight subsets of differentially expressed genes. The Ch (Champagne) up and Te (Térèse) up columns correspond to “LR”, followed by two columns of “TCR” down and up and the last four columns referring to the “TLSR” a, b, c, and d, as described in the legend of Figure 3.
Table 2. GO matching of the cold response (i.e., “response to cold”, “cold acclimation”, “cellular response to cold”) in eight subsets of differentially expressed genes. The Ch (Champagne) up and Te (Térèse) up columns correspond to “LR”, followed by two columns of “TCR” down and up and the last four columns referring to the “TLSR” a, b, c, and d, as described in the legend of Figure 3.
Protein/Gene/FunctionClass of FunctionReferenceOrthologous in ATCh UpTe UpTCR DownTCR UpTLSR aTLSR bTLSR cTLSR d
GO:0009409 (response to cold)
Calmodulin-binding transcription activator 2TF, Induction of CBFs[53]AT5G64220.2 1
Plasma-membrane cation-binding protein 1Plasma membrane protein[54]AT4G20260.6 1
Vacuolar H(+)-ATPase subunit E1Hydrogen ion transport[54]AT4G11150.1 1
Adenine nucleotide alpha hydrolases-likeCold shock response[54]AT3G53990.1 2
Calcium-dependent lipid-binding proteinResponse to cold[54]AT4G34150.1 1
SAUR-like auxin-responsive protein familyAuxin metabolism[55]AT4G38840.1 44
Enolase, ENO2Glycolysis/Gluconeogenesis[56]AT2G36530.1 1
MAP kinase kinase kinase1Kinase activity[57]AT4G08500.1 1
Protein HAPLESS 6, Ribophorin IIN-linked glycosylation[58]AT4G21150.3 1
Cold, circadian rhythm, RNA-binding 2, GRP7RNA-binding[59]AT2G21660.255 5
Glycine-rich RNA-binding protein 3, RBG3RNA-binding, transcription[59]AT5G61030.12
Hydrophobic protein RCI2A and LTI6AResponse to cold[60]AT3G05880.1 2
Late embryogenesis abundant protein 46Cryoprotectant[61]AT5G06760.1 1
Galactinol synthase 2, GOLS2Galactose metabolism[62]AT1G56600.1 1
Protein sensitive to freezing 2Glucosidase activity[63]AT3G06510.1 1
Pentatricopeptide repeat-containing proteinRNA modification, binding[64]AT3G22690.2 1
RNA-binding (RRM/RBD/RNP motifs)rRNA-binding[64]AT1G70200.1 1
Outer envelope pore protein 16-1, OEP16-1Amino acid transport, porin[65]AT2G28900.12 2
Glutathione S-transferase F8Oxidoreductase, peroxidase[66]AT2G47730.1 1
Glyoxalase Imethylglyoxal degradation[66]AT1G67280.2 2
Plastid-lipid-associated protein 1, Fibrillin-1aPhotoinhibition[66]AT4G04020.1 1
Phosphoribulokinase, PRKPhotosynthesis, transferase[66]AT1G32060.1 1
NADPH-dependent alkenal/one Oxidoreductase[66]AT1G23740.12 2
Serine hydroxymethyltransferase One-carbon metabolism[66]AT4G37930.1 1
RuBisCo activaseATP- and nucleotide-binding[66]AT2G39730.1 1
Chlorophyll a-b binding protein 4, LHCA4Photosynthesis[67]AT3G47470.1 1 1
NADPH-dependent aldo-keto reductaseOxidation-reduction process[68]AT2G37770.2 2 2
3-hydroxyisobutyryl-CoA hydrolase 1 L-valine degradation[69]AT5G65940.1 1
PLAT domain-containing protein 1, PLAT1Catalase, Peroxidase activity[70]AT4G39730.12 2
Serine/threonine protein kinaseATP binding, transferase[71]AT3G08720.2 1
Phosphoglyceromutase 1, PGAM 1 Glycolysis/Gluconeogenesis[72]AT1G09780.1 1
3-ketoacyl-CoA synthase 1Fatty acid elongation[73]AT2G26250.11 2 2
Acyl-CoA-binding protein 1, ACBP1Fatty Acid Beta-Oxidation[74]AT5G53470.1 1 1
Agamous-like MADS-box protein, SOC1Transcription regulation[75]AT2G45660.1 2
Alcohol dehydrogenase 1Glycolysis/Gluconeogenesis[76]AT1G77120.11
Annexin D8, calcium/phospholipid bindingCalcium binding[77]AT5G12380.1 11 1
BAG family molecular chaperone regulator 4Chaperone binding[78]AT3G51780.11
Calcium-binding proteinCalcium ion binding[79]AT1G02270.11
Heat shock 70 KDa protein 1, HSP70-1ATPase activity, chaperone[79]AT5G02500.1 2
DNA damage-repair/toleration, DRT102Isomerase activity[79]AT3G04880.1 1
Proteasome subunit alpha type-3, PAG1Folding, sorting, degradation[79]AT2G27020.1 1
Hsp 70 kDa protein 1ATPase activity, Chaperone[79]AT5G02500.1 2
Calmodulin-binding receptor-like, kinase 1 Calmodulin binding[80]AT5G58940.1 1
Serine/threonine-protein kinaseATP-binding[81]AT1G01140.3 1
Chaperonin-like RBCX protein 1Protein folding chaperone[82]AT4G04330.1 1
Cinnamoyl-CoA reductase 1Lignin biosynthesis[83]AT1G15950.1 1
Cold regulated protein 27, COR27Cold, circadian rhythm[84]AT5G42900.3 1
Cysteine proteinase inhibitor 6, CYS6Cysteine proteinase inhibitor[85]AT3G12490.2 1
Protein CRYOPHYTE, RH38RNA-binding, hydrolase[86]AT3G53110.1 1
Diacylglycerol kinase 2Glycerolipid metabolism[87]AT5G63770.1 1
E3 ubiquitin-protein ligase HOS1Protein ubiquitination[88]AT2G39810.1 1
Early light-induced protein 1, ChloroplasticPhotosynthesis[89]AT3G22840.11
Ethylene-responsive TF, RAP2-4 and RAP2Transcription factor[90]AT1G78080.1 1 1
Glycine-rich RNA-binding protein RZ1ARNA-binding, transcription[91]AT3G26420.1 1 1
HVA22-like protein a, similarity to TB2/DP1Cold and stress response[92]AT1G74520.11 1
Inositol-1-monophosphataseMyo-inositol biosynthesis[93]AT3G02870.31
Lipid transfer protein EARLI 1Lipid-transfer[94]AT4G12480.1 1
LOW-TEMPERATURE-INDUCED 65, LTI65Response to abscisic acid[95]AT5G52300.2 1
MAP kinase 3ATP binding[96]AT3G45640.11
MYB-related transcription factor CCA1DNA binding[97]AT2G46830.1 2
Phosphoinositide phospholipase C1Hydrolase, lipid metabolism[98]AT5G58670.1 1
Phospholipase D deltaLipid degradation[99]AT4G35790.2 2 2
Protein EARLY FLOWERING 3, ELF3DNA-binding TF activity[100]AT2G25930.1 1
Protein ESKIMO 1, Signal-anchorXylan O-acetyltransferase[101]AT3G55990.1 1
Protein GIGANTEA, GIPhytochrome B signaling[102]AT1G22770.1 1
Protein Senescence-Associated Gene 21Oxidative stresses[103]AT4G02380.12
Raffinose synthase 6Carbohydrate metabolism[104]AT5G20250.4 1
Synaptotagmin-1, SYT1Lipid binding, Ca2+ transport[105]AT2G20990.1 1
Transcription factor GTE10, NPX1ABA signaling pathway[106]AT5G63320.1 1
Tubulin beta-6 chain, TUBB6GTPase activity, Transport[107]AT5G12250.1 1
WRKY DNA-binding protein 33, WRKY33DNA-binding TF activity[108]AT2G38470.1 1
Xyloglucan endotransglucosylase/hydrolase protein 22 Glycosidase, Transferase[109]AT5G57560.1 1
GO:0009631 (cold acclimation)
Cold-regulated 413 thylakoid membrane 1Cellular response to cold[110]AT1G29395.11
3-oxoacyl-[acyl-carrier-protein] synthase IIFatty acid biosynthesis[111]AT1G74960.31
Mediator of RNA polymerase II transcription subunit 32Transcription regulation[112]AT1G11760.11
Calcium/calmodulin-regulated receptor-like kinase 1Calmodulin-binding[113]AT5G54590.2 1
Alpha-glucan water dikinase 1Carbohydrate metabolism[114]AT1G10760.1 1
Transcriptional adapter ADA2b, PRZ1Transcription regulation[115]AT4G16420.1 1
Glycine-rich RNA-binding protein 2Chaperone, RNA-binding[116]AT4G13850.3 1
Acetyl-CoA carboxylase 1Fatty acid metabolism[117]AT1G36160.2 1
C-repeat binding factor 3, DREB1A, CBF3TF, DNA-binding[118]AT4G25480.11
VOZ1, vascular plant one zinc fingerTranscription factor[119]AT1G28520.2 1
GO:0070417 (cellular response to cold)
Pre-mRNA-processing factor 31 homologRNA-binding, splicing[120]AT1G60170.1 1
Bidirectional sugar transporter SWEET17Fructose, sugar transport[121]AT4G15920.1 1
Delta(8)-fatty-acid desaturase1Oxidoreductase[122]AT2G46210.11 1
Glutamate receptor 3.4, GLR3.4Ion transport[123]AT1G05200.2 1
Organic cation/carnitine transporter 3, OCT3Transporter activity[124]AT1G16390.1 1
spliceosome protein-like proteinRNA processing, splicing[125]AT1G54380.11
Total of genes in GO bulk analyses 2936143096912
Total of genes in each subset 9061581520883253228131479
Ratio % 3.202.282.693.403.562.636.872.51

Share and Cite

MDPI and ACS Style

Bahrman, N.; Hascoët, E.; Jaminon, O.; Dépta, F.; Hû, J.-F.; Bouchez, O.; Lejeune-Hénaut, I.; Delbreil, B.; Legrand, S. Identification of Genes Differentially Expressed in Response to Cold in Pisum sativum Using RNA Sequencing Analyses. Plants 2019, 8, 288. https://doi.org/10.3390/plants8080288

AMA Style

Bahrman N, Hascoët E, Jaminon O, Dépta F, Hû J-F, Bouchez O, Lejeune-Hénaut I, Delbreil B, Legrand S. Identification of Genes Differentially Expressed in Response to Cold in Pisum sativum Using RNA Sequencing Analyses. Plants. 2019; 8(8):288. https://doi.org/10.3390/plants8080288

Chicago/Turabian Style

Bahrman, Nasser, Emilie Hascoët, Odile Jaminon, Frédéric Dépta, Jean-François Hû, Olivier Bouchez, Isabelle Lejeune-Hénaut, Bruno Delbreil, and Sylvain Legrand. 2019. "Identification of Genes Differentially Expressed in Response to Cold in Pisum sativum Using RNA Sequencing Analyses" Plants 8, no. 8: 288. https://doi.org/10.3390/plants8080288

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop