Skip to main content

Systematic analysis of Heat Shock Protein 70 (HSP70) gene family in radish and potential roles in stress tolerance

Abstract

The 70 kD heat shock proteins (HSP70s) represent a class of molecular chaperones that are widely distributed in all kingdoms of life, which play important biological roles in plant growth, development, and stress resistance. However, this family has not been systematically characterized in radish (Raphanus sativus L.). In this study, we identified 34 RsHSP70 genes unevenly distributed within nine chromosomes of R. sativus. Phylogenetic and multiple sequence alignment analyses classified the RsHSP70 proteins into six distinct groups (Group A–F). The characteristics of gene structures, motif distributions, and corresponding cellular compartments were more similar in closely linked groups. Duplication analysis revealed that segmental duplication was the major driving force for the expansion of RsHSP70s in radish, particularly in Group C. Synteny analysis identified eight paralogs (Rs-Rs) in the radish genome and 19 orthologs (Rs-At) between radish and Arabidopsis, and 23 orthologs (Rs-Br) between radish and Chinese cabbage. RNA-seq analysis showed that the expression change of some RsHSP70s were related to responses to heat, drought, cadmium, chilling, and salt stresses and Plasmodiophora brassicae infection, and the expression patterns of these RsHSP70s were significantly different among 14 tissues. Furthermore, we targeted a candidate gene, RsHSP70–23, the product of which is localized in the cytoplasm and involved in the responses to certain abiotic stresses and P. brassicae infection. These findings provide a reference for further molecular studies to improve yield and stress tolerance of radish.

Peer Review reports

Background

Heat shock proteins (HSPs) are conserved stress-responsive proteins induced by adverse environmental conditions, which can be divided into six major subfamilies according to their molecular masses, i.e., HSP40, HSP60, HSP70, HSP90, HSP100, and small HSPs (sHSPs) [1]. HSP70 is the most widely studied type of HSP and is characterized by the N-terminal ATPase domain (NBD), substrate binding domain (SBD), and a variable C-terminal lid region domain [2]. Based on the subcellular localization, HSP70s in plants have been classified into four major subfamilies: those present in the cell nucleus/cytoplasm (EEVD motif), endoplasmic reticulum (HDEL motif), plastids (PEGDVIDADFTDSK motif), and mitochondria (PEAEYEEAKK motif) [3].

As molecular chaperones, the most important biological function of HSP70s is linked to acquired thermotolerance under high-temperature stress, and function as negative feedback regulators of heat shock transcription factor (HSF) activity [4,5,6,7]. In Arabidopsis, cpHsc70–1(At4g24280) is not only essential for normal plant growth but is also important for root growth from heat-stressed seeds [8]. AtHSP70–15 affects not only the growth phenotype and leaf morphology but also the response to heat stress [9]. Mutation of the rice chloroplast OsHsp70CP1 causes chloroplast developmental defects under high-temperature conditions [10]. Plastid HSP70-2 (cpHsp70-2) involved with the temperature-dependent chalkiness of rice grains [11]. HSP70s have also been studied under high temperature stress in a variety of vegetables, such as pepper [12], pumpkin [13], potato [14], cucumber [15] and tomato [16]. In addition, HSP70 proteins in plants also improve tolerance to low-temperature, high-salinity, drought, light, flooding, and heavy metal stress in Arabidopsis, maize, rice, barley, wheat, soybean, tobacco, poplar, sugarcane, and chlamydomonas [16]. Furthermore, some HSP70s in plants are also involved in microbial pathogenesis, particularly viral infections [17]. Cytoplasmic HSP70s enhance the infection of Nicotiana benthamiana by tobacco mosaic virus, potato virus x, cucumber mosaic virus, and watermelon mosaic virus [18]. HSP70-depleted plants show increased susceptibility to Pseudomonas syringae [19], and silencing of cytosolic HSP70a in pepper results in enhanced susceptibility to Xanthomonas campestris infection [20]. Plants show enhanced resistance to infection by rice stripe virus (RSV) [21] through the accumulation of endoplasmic reticulum (ER)-resident HSP70s (also called binding proteins, BiPs).

Radish (Raphanus sativus L., 2n = 18) is an important commercial root vegetable crop that belongs to the Brassicaceae family and is sensitive to heat [22], salt [23] or heavy metal stress [24,25,26]. Although the HSP70 gene family is associated with different abiotic stress responses in many plant species, the genome-wide identification and functional characterization of the HSP70 family in radish have not been reported previously. In this study, 34 HSP70 family members were identified in the radish genome and divided into different classes, and the phylogenetic relationships, chromosome arrangements, gene structures, conserved motifs, and expression profiles of RsHSP70 genes in distinct tissues or in response to various environmental stresses were also systematically analyzed. Our results provide a biological reference for elucidating the functions of HSP70 genes in radish, and will be useful for the selection of candidate genes for genetic engineering in R. sativus breeding.

Materials and methods

Identification of HSP70 genes

The whole-genome sequences of radish were obtained from the BRAD database (http://brassicadb.org/brad/) [27]. A Hidden Markov Model (HMM) profile containing HSP70 (Pfam: PF00012) was downloaded from the InterPro database (https://www.ebi.ac.uk/interpro/) [28], and used to confirm the candidate HSP70s in the radish genome with HMMER v3.2.1 (http://hmmer.org/) with an E cutoff value of 0.01. The HSP70 protein sequences of Arabidopsis were obtained from the TAIR database [29]. In addition, all AtHSP70s protein sequences were searched against radish protein sequences using BLASTP with an E cutoff value of 10–10. Subsequently, all of the putative radish HSP70 sequences were merged and further verified using the online SMART database (http://smart.embl-heidelberg.de/) [30]. The HSP70 protein sequences of rice, cabbage, rapeseed, and Chinese cabbage were obtained from a previous report [31]. The physicochemical parameters of each HSP70 protein were calculated using the online ExPASy server (https://www.expasy.org/resources/compute-pi-mw) [32], and the subcellular localization was analyzed using WoLF PSORT (http://www.genscript.com/psort/wolf_psort.html) [33].

Sequence alignments and phylogenetic analysis

The full protein sequences of HSP70 from radish, Arabidopsis, cabbage, rapeseed, Chinese cabbage, and rice were aligned using ClustalW with the default parameters implemented in MEGA 7.0 (v7.0.21), which were used to construct the neighbor-joining (NJ) phylogenetic tree with specific parameters (model: Poisson model; gap: pairwise deletions; bootstrap replicates: 1000) [34]. The phylogenetic tree was visualized with EVOLVIEW (http://www.evolgenius.info/evolview/).

Analyses of gene structure, and cis-regulatory elements

For gene structure analysis, the exon–intron organizations of individual RsHSP70 genes were determined using the online GSDS program (http://gsds.cbi.pku.edu.cn/) [35]. The Multiple Em for Motif Elicitation (MEME v5.1.1,http://meme-suite.org/) was used to analyze the conserved motifs with the following parameters: number of repetitions: any; maximum number of motifs: 20; optimum motif widths: 10–100 amino acid residues [36]. The exon–intron organizations and conserved motifs were visualized with TBtools (v1.108) [37]. The online PlantCARE database (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/) was used to determine the putative cis-acting regulatory elements, as described previously [38].

Chromosomal location and identification of orthologous and paralogous genes

The radish genome was searched for all RsHSP70 gene loci and the corresponding location was visualized with Circos software (v 0.69) [39]. Duplicated RsHSP70 gene pairs were identified using MCScanX with the default settings [40]. The nonsynonymous replacement (Ka) and synonymous replacement (Ks) rates were estimated using Ka/Ks Calculator 2.0 implemented in TBtools (v1.108) [37], The formula T = Ks/2λ (λ = 6.96 × 10–9) was used to estimate the divergence time of duplication events [41]. OrthoMCL(v6.18) (http://orthomcl.org/orthomcl/app) was used to investigate potential orthologous and paralogous HSP70 genes in radish, Arabidopsis, and Chinese cabbage with the standard settings, and visualized with Circos (v 0.69) [39].

Expression analysis of RsHSP70 genes in various tissues or under different stresses

To examine the potential functions of RsHSP70 genes in different tissues of radish, the raw RNA-seq data of 14 tissues were acquired from a previous report [42] (Table S 1). These data represented five tissues (leaf, root tip, cortex, cambium, and xylem) and five stages (7, 14, 20, 40, and 60 days after sowing, DAS). To investigate whether RsHSP70 genes play important roles in various stress responses in radish, data from seedlings exposed to heat, drought, cadmium (Cd), chilling, salt stress were obtained from the NCBI SRA database (SUB13505845) and P.brassicae infection were obtained from the CNCB database (https://ngdc.cncb.ac.cn/) with the number CRA004024. The expression level of each gene presented in transcripts fragments per kilobase of exon model per million mapped fragments (FPKM) was reanalyzed as described previously [43]. Differentially expressed genes (DEGs) were selected with cutoff values of |log2fold-change|> 1 and p < 0.05. Heat maps were generated in R (v3.6.3) with log2-transformed FPKM values after the addition of a pseudocount of 0.01.

Plant material and treatment

Two inbred radish lines with contrasting resistance to P. brassicae, WR1150 and 1116 T, (resistant and susceptible, respectively) obtained from the Institute of Vegetables and Flowers, Chongqing Academy of Agricultural Sciences, China, were used. The seeds of the two lines were surface-sterilized, sown in trays containing a 3:1 mixture of nutrient soil and sand, and grown under conditions of 25 °C/20 °C (day/night) with a 16 h photoperiod. The pathogen used in this study was obtained from the natural clubroot nursery at Wulong, Chongqing, China, where P. brassicae race 4 is the dominant clubroot pathogen. P. brassicae suspensions were prepared as described previously [44], and inoculate we prepared by diluting the suspension to a resting spore concentration of 1 × 108/mL. Ten similar seedlings with 20-day leaves were individually injected at the bottom of the stem with 5 mL spore suspension. Total RNA was isolated from roots collected at 0, 4, 7, 14, 21, and 28 days post inoculation (dpi) of P. brassicae for qRT-PCR analysis.

To analyze abiotic stress, five similar seedlings of 1116 T with 20-day leaves were treated with CdCl2·2.5H2O (200 mg/L), mannitol (300 mM), or NaCl (200 mM) for 48 h [23, 24], or by chilling (4 °C) or heating (40 °C) for 24 h [22]. Leaves were collected at each time point and used to isolate total RNA.

RNA isolation and qRT-PCR analysis

Total RNA was isolated from 20-day leaves using an RNAprep pure Plant Kit (Catalog no.dp432; Tiangen Biotech Co., Ltd., Beijing, China), and quantified using the NanoDrop ND-1000 (Termo Scientifc, Waltham, MA, USA).The first-stand cDNA was synthesized using a HiScript III 1st Strand cDNA Synthesis Kit (+ gDNA wiper) (Catalog no.R312–01; Vazyme Biotech Co., Ltd., Nanjing, China). ChamQ Universal SYBR qPCR Master Mix (Catalog no.Q711–02; Vazyme Biotech) was used for qRT-PCR on a BIO-RADCFX96 Real Time System (Bio-Rad Laboratories, Hercules, CA, USA). Three independent biological replicates were used for each sample. The transcript levels were calculated using the ΔΔCT method with normalization relative to the level of RsActin [45]. All gene-specific primers used in this study are presented in Table S 2.

Determination of subcellular localization

For transient expression, the coding sequence of RsHSP70–23 was inserted into pAN580 to generate GFP fusion proteins. The plasmids pAN580-GFP-35Spro:GFP and pAN580-GFP-35Spro:RsHSP70–23-GFP were transformed into Arabidopsis protoplasts. The GFP signals were detected by confocal laser microscopy (Carl Zeiss, Oberkochen, Germany).

Yeast constructs, tolerance assay and growth curve

The coding sequence of RsHSP70–23 was inserted into pRS-416-GFP, and the recombinant vector and the empty pRS-416-GFP vector (control) were transformed into the wild-type strain JRY472 and allowed to grow on SD-Ura plates. Positive recombinant transformants cultured in SD-Ura medium were diluted until OD600 = 0.1. Then the cell culture was diluted ten fold and treated with 75 µM Cd, 1 M NaCl, and 2 M mannitol, and incubated at 30 °C for 3 days. Cold and heat stress were applied at 4 °C and 37 °C for 2 days before transfer to 30 °C for 1 or more days [46]. Then, the phenotypes of the yeast cells were photographed, and the experiment was repeated three times. RsHSP70-23 overexpressing yeast cells were grown in liquid SD-Ura medium at 4℃, 30℃ and 37℃, respectively, and with 1 M NaCl were also grown at 30℃, and were diluted until OD600 = 0.1, OD600 is recorded every 2 h to prepare cell growth curve [46].

Results

Identification of RsHSP70 genes in radish genome

Through a combination of BLASTP and HMM profile analysis, 34 putative RsHSP70 genes were identified in the “Xiangyabai” genome and renamed according to their relative linear order on each chromosome (Table 1). Among them, the lengths of predicted RsHSP70 protein sequences varied from 112 (RsHSP70–32) to 1335 (RsHSP70–4) amino acids with molecular weights (MWs) ranging from 12.21 to 144.26 kDa. The theoretical isoelectric points (pIs) of most predicted RsHSP70 proteins were < 7, with the exceptions of RsHSP70–4, RsHSP70–29, RsHSP70–30, and RsHSP70–32. In addition, WoLF PSORT online analysis predicted that within the 34 RsHSP70 proteins, 19 radish HSP70 proteins were localized to the cytosol/nucleus, 7 to the chloroplast, 4 to the mitochondria, and 4 to the ER (Table 1).

Table 1 The general information and sequence characterization of 34 RsHSP70 genes

Six groups defined among the HSP70 genes of six species

To investigate the phylogenetic relationships of these HSP70s, the full-length amino acid sequences of RsHSP70s (34), BnHSP70s (Brassica napus, 47), BrHSP70s (Brassica rapa, 29), BoHSP70s (Brassica oleracea, 20), OsHSP70s (Oryza sativa, 32), and AtHSP70s (Arabidopsis thaliana, 18) were used to construct a neighbor-joining (NJ) phylogenetic tree; they clustered into six distinct groups (Groups A–F) (Fig. 1). Among them, Group A was the most abundant subfamily containing 58 members, consisting of 13 RsHSP70 members that were predicted to be localized in the cytoplasm/nucleus (Table 1). Groups B, C, and D consisted of four, five, and four members from radish, which were predicted to be located in the ER, chloroplast, and mitochondria, respectively. These results suggest that the closely related HSP70s are usually located in the same subcellular structure. In total, 38 members belonging to the Hsp110/SSE subfamily were classified into Group F, which also contained 7 RsHSP70 members. Group E only had one RsHSP70, which was suggested to be a truncated gene based on comparative analysis with its Arabidopsis counterpart [47]. Each group contained proteins from both monocotyledonous and dicotyledonous plants, indicating that the main characteristics of HSP70 proteins in plants emerged before the separation of these two lineages.

Fig. 1
figure 1

Phylogenetic relationships of radish, rapeseed, Chinese cabbage, cabbage, rice, and Arabidopsis HSP70 proteins. The tree was divided into six subgroups, marked by different color backgrounds

Characterization of RsHSP70 proteins and distributions of conserved motifs and gene structures

Sequence alignments revealed that most HSP70 family proteins in radish included three distinct domains. The highly conserved NBD domain possessed three HSP70 signature sequences, the SBD domain was also conserved, while the C-terminal domain was highly variable (Figure S1). However, RsHSP70–14 and RsHSP70–27 did not include the NBD domain, and RsHSP70–32 and RsHSP70–33 lacked the C-terminal domain. All 34 RsHSP70s clustered into six groups in phylogenetic analysis (Fig. 2A). The majority (11/13, 84.6%) of RsHSP70s in Group A possessed a conserved retention signal “EEVD” sequence at the C-terminus, and were localized in the cytoplasm. Half (2/4) of those in Group B were localized in the ER and 60% (3/5) of those in Group C were localized in the chloroplasts, and possessed the conserved C-terminal signature sequences “HDEL” and “DVIDADFTDSK,” respectively. In addition, the three RsHSP70s in Group D possessed the conserved signature sequence “PEAEYEEAKK” in the C-terminus, and were localized in the mitochondria (Figure S1).

Fig. 2
figure 2

Phylogenetic relationship, gene structures and conserved motifs of RsHSP70s. A Phylogenetic tree of 34 RsHSP70s proteins. B Conserved motifs of RsHSP70s proteins. Each colored box represents a specific motif in the protein identified using the MEME motif search tool. C Exons and introns were indicated by green rectangles and red lines respectively

Twenty conserved motifs were identified in radish HSP70 proteins using the MEME motif search tool (Fig. 2B and Table S 3). Motifs 2 (24/34), 3 (26/34), and 12/10 (22/34) were found in the RsHSP70 family, corresponding to signature sequences. Motifs 1 (23/34), 5 (29/34), and 7 (25/34) were included in the SBD domain. Similar compositions, orders, and numbers of motifs were found in the groups localized to the mitochondria and chloroplasts. However, some subfamilies also included several specific motifs. For example, motifs 13 and 14 were exclusively present in the HSP110/SSE subfamily, which lacked motif 1; motif 19 was only found in those localized to the cytosol; and motifs 16 and 20 were absent in those localized to the cytosol or ER.

The online GSDS tool was used to identify the exon–intron organization in the coding sequences shared among the RsHSP70s (Fig. 2C). The number of exons in RsHSP70s varied from 1 to 14. Most closely related members that clustered together shared similar exon–intron organization and exon length. For example, the two members of the HSP110/SSE subfamily have 13 introns and 14 exons, and are nearly 2700 bp in length, while 84.6% of cytosolic HSP70s had no introns. These results indicate a diversity of exon–intron organization in the radish HSP70 family.

Gene duplication and synteny analysis of RsHSP70s

Chromosome location analysis showed that in addition to RsHSP70–34 the remaining 33 RsHSP70 genes were irregularly distributed on the nine chromosomes in radish (Fig. 3). Chromosome 8 had the greatest number of HSP70 genes (n = 6), followed by five genes on each of chromosomes 2 and 6, four genes on chromosomes 4, three genes on each of chromosomes 1, 7 and 8, and only two on each of chromosomes 3 and 9.

Fig. 3
figure 3

Chromosomal localization and gene duplication events of RsHSP70 genes. Respective chromosome numbers are indicated at the top of each bar. Tandem duplicated genes are marked on a grey background. Segmental duplicated genes are shown by black line

To investigate the forces driving gene family expansion, tandem and segmental duplication patterns were examined. Among the 34 RsHSP70 genes, 35.29% (12/34) arose from duplication events, including one pair derived from tandem duplication and 11 genes derived from segmental duplications (Fig. 3). The tandem duplicated genes were RsHSP70–4 and RsHSP70–5 located on chromosome 2. The 11 segmental duplicated RsHSP70 genes were classified into five groups. One group consisted of three genes, RsHSP70–4, RsHSP70–11, and RsHSP70–27, located on chromosomes 2, 4, and 8, respectively. The other four groups consisted of pairs of genes: RsHSP70–3 and RsHSP70–14, RsHSP70–6 and RsHSP70–12, RsHSP70–7 and RsHSP70–9, and RsHSP70–21 and RsHSP70–31. It was noteworthy that RsHSP70–4 underwent tandem duplication before segmental duplication (Table 2). The Ka/Ks ratio varied from 0.014 to 0.388, indicating that these genes were subject to purifying selection. The segmental duplication occurred at 2.49–7.58 million years ago (Mya), while tandem duplication occurred at 3.32 Mya.

Table 2 Ka-Ks calculation for each pair of HSP70 in radish

Synteny analysis showed the presence of 8 paralogs (Rs-Rs) in the radish genome, 19 orthologs (Rs-At) between radish and Arabidopsis, and 23 orthologs (Rs-Br) between radish and Chinese cabbage (Fig. 4 and Table S 4). Between radish and Arabidopsis, AT4G24280 (cpHSC70–1) has three orthologous genes (RsHSP70–4, RsHSP70–11, and RsHSP70–27), AT1G16030 (HSP70B), AT4G16660, and AT5G09590 (mtHSC70–2) have two orthologous genes, and six AtHSP70s had only one orthologous gene, e.g., AT5G42020(BIP2)/RsHSP70–23, AT1G09080(BIP3)/RsHSP70–16, and AT2G32120(HSP70T-2)/RsHSP70–1, suggesting that there have been gene losses in many orthologous groups. It is noteworthy that 13 RsHSP70 genes (RsHSP70–1/-4/-6/-7/-9/-11/-12/-16/-21/-22/-27/-29/-31) were shared between the orthologs identified in Rs-At and Rs-Br. These results provide a theoretical foundation for further exploration of RsHSP70 gene functions.

Fig. 4
figure 4

The syntenic relationship of HSP70 genes between radish and Arabidopsis, radish and Chinese cabbage. Lines in red indicate radish paralogous genes pairs, both blue line and green line indicate orthologous gene pairs

Expression patterns of radish HSP70s in various tissues

Based on RNA-seq data, we compared the expression profiles of all RsHSP70 genes in 14 tissues (Fig. 5A and Table S 5). The expression patterns were divided into three major groups. In subgroup Rs-A, all of the genes maintained a relatively high transcription level in all tissues examined, particularly RsHSP70–2, RsHSP70–25, and RsHSP70–34 (TPM > 100). Nine RsHSP70s (RsHSP70–1, -4, -16, -19, -21, -23, -27, -32, and -33) in subgroup Rs-B showed low or no expression in all of the tissues examined (TPM < 5). Twelve genes in subgroup Rs-C showed differential expression between different tissues or developmental stages, while some genes maintained higher expression in specific organs, such as RsHSP70–5 showing selective high-level expression in leaves, and RsHSP70–31 showing higher expression in 20-day leaves compared to all other tissues examined. Most RsHSP70 genes showed higher expression levels in 20-day leaves and roots (Fig. 5B), suggesting that they may play vital roles in seedling development.

Fig. 5
figure 5

Expression profiles of RsHSP70 genes in different tissues. A The normalized expression levels of the hierarchical clustering of 34 RsHSP70 genes in 14 tissues. The relative expression levels corresponding to the log2-transformed TPM values after the addition of a pseudocount of 0.01 are shown. The scale represents the relative signal intensity of the TPM values. B The expression values of all the RsHSP70 genes in each tissue

Interestingly, three pairs of paralogs (RsHSP70–6/RsHSP70–12, RsHSP70–7/RsHSP70–9, and RsHSP70–21/RsHSP70–31) showed different expression levels in different tissues, suggesting that duplicated genes may acquire new functions and increase adaptation to stressful environmental conditions. Moreover, five pairs of paralogs (RsHSP70–2/RsHSP70–34, RsHSP70–3/RsHSP70–14, RsHSP70–10/RsHSP70–17, RsHSP70–29/RsHSP70–30, and RsHSP70–32/RsHSP70–33) exhibited similar expression levels, possibly indicating subfunctionalization in the adaptation to abiotic stresses (Fig. 5A).

Expression profiles of radish HSP70 genes under abiotic stress

RNA-seq analyses showed that 25 genes were differentially expressed after exposure to heat, Cd, drought, high salinity, and chilling stresses (Fig. 6A and Table S 6). Seven RsHSP70s (RsHSP70–3, RsHSP70–4, RsHSP70–8, RsHSP70–14, RsHSP70–20, RsHSP70–21, and RsHSP70–31) were significantly upregulated following treatment with Cd and NaCl. The expression levels of three RsHSP70 genes changed only under conditions of drought stress (RsHSP70–5, RsHSP70–26, and RsHSP70–34). Under low-temperature conditions, the expression levels of RsHSP70–3, RsHSP70–14, RsHSP70–23, and RsHSP70–28 were significantly downregulated. Sixteen RsHSP70 genes were consistently upregulated under conditions of heat stress, including RsHSP70–1, RsHSP70–4, RsHSP70–8, RsHSP70–14, RsHSP70–16, RsHSP70–19, RsHSP70–21, RsHSP70–23, and RsHSP70–31 (4- to eightfold increases). We predicted the presence of several cis-elements that were responsive to environmental stressors (drought, low temperature, stress defense mechanism, and wounding) and hormones (abscisic acid [ABA], auxin [indole-3-acetic acid, IAA], gibberellin [GA3], methyl jasmonate [MeJA], and salicylic acid [SA]) in the promoter regions of RsHSP70s (Table S 7). Using the PlantPAN3 online database, we found that 31 RsHSP70s contained the heat shock factor (HSF1)-binding element (HSE) consisting of two main subunits (5′NGAAN3′ and 5′NTTCN3′)(Table S 8). These results suggest that most RsHSP70 genes are induced and/or repressed by abiotic stress. To assess the accuracy of the RNA-seq data, nine RsHSP70 genes from different groups were examined by qRT-PCR analysis under conditions of heat, salt, cold, drought, and Cd stress. The gene expression patterns obtained by qRT-PCR were generally consistent with the results of RNA-seq analyses (Fig. 6B).

Fig. 6
figure 6

Expression profiles of RsHSP70 genes in response to abiotic stress treatments. A Heatmap of differential expression of RsHSP70 genes in response to heat, drought, cd, salt and chilling stress. The color scale of heatmap is based on the log2-transformed FPKM values after the addition of a pseudocount of 0.01 between experimental and control tissues are shown. B Expression profiles of selected genes under heat, drought, cd, salt and chilling stresses from RNA-seq data and qPCR analysis. Correlation (R.2 = 0.7383–0.9295, P < 0.05) between RNA-seq data and qPCR results. Each bar shows the mean ± SE of the triplicate assay. Different letters indicate significant differences among treatments (P < 0.05, ANOVA)

Potential roles of HSP70 genes in radish root with P. brassicae infection

Most HSP70 genes in Arabidopsis are upregulated in response to biotic stressors [17]. To explore the roles of RsHSP70s genes responsive to biotic stress, we compared the expression profiles of all RsHSP70 genes in the roots at 0, 7, and 28 dpi with P. brassicae (Fig. 7). In all, 11 RsHSP70 genes were either up- or downregulated and showed significant differential expression in response to P. brassicae infection between the clubroot-resistant (RB) and clubroot-susceptible (SB) lines. Only RsHSP70–5, RsHSP70–23, and RsHSP70–33 tended to show continuous upregulation at all time points examined. It is noteworthy that the level of RsHSP70–23 expression was 6.83-fold higher in RB than SB roots at 28 dpi (Fig. 7). qRT-PCR analysis showed that the level of RsHSP70–23 transcript increased sharply in the clubroot-resistant strain (WR1150) compared to the clubroot-susceptible strain (1116 T) after 4 dpi, and remained at a high level from 7 to 28 dpi. These results suggest that RsHSP70–23 may participate in the response to P. brassicae infection in radish.

Fig. 7
figure 7

Expression profiles of RsHSP70 genes in response to P.brassicae infection. A Heatmap of differential expression of RsHSP70 genes in response to P.brassicae infection in RB and SB at 0, 7 and 28 dpi. The color scale of heatmap is based on the log2Foldchange values between RB and SB are shown. B RT-PCR analysis of selected genes under P. brassicae infection in WR1150 and 1116 T at 0, 4, 7, 14, 21 and 28 dpi. Error bars indicate the SD for three independent replicates. * and ** indicate a signifcant diference between WR1150 and 1116 T at P < 0.05 and P < 0.01 levels, respectively (two tailed T-test)

Subcellular localization of RsHSP70–23

Large-scale synteny analysis revealed that RsHSP70–23 was orthologous to AT5G42020 (BIP2), which is localized to the ER (Fig. 4). WoLF PSORT analysis showed that RsHSP70–23 was also localized to the ER (Table 1), but did not possess the conserved “HDEL” sequence in the C-terminus (Figure S1). To determine the actual subcellular localization of RsHSP70–23 in vivo, GFP-tagged RsHSP70–23 was transiently expressed in Arabidopsis protoplasts. Free GFP was evenly distributed in the cytoplasm as expected, and RsHSP70–23-GFP fusion protein was also observed in the cytoplasm (Fig. 8).

Fig. 8
figure 8

The subcellular localization of RsHSP70-23. Bars = 10 μm

RsHSP70–23 overexpression in response to abiotic stresses in yeast

To examine whether expression of RsHSP70–23 responds to abiotic stress, RsHSP70–23-overexpressing yeast cells were exposed to 75 µM Cd, 1 M NaCl, 2 M mannitol, chilling (4 °C), and heat (37 °C) stresses (Fig. 9). There were no differences between cells carrying the RsHSP70–23 overexpression vector or empty vector (EV) control under conditions of mannitol or Cd stress. However, the RsHSP70–23-overexpressing cells were sensitive to cold stress compared to EV. Under conditions of NaCl and heat stress, all cells overexpressing RsHSP70–23 grew faster than those carrying EV. In addition, we conducted growth curves of the yeast cells under heat, chilling and salt stresses (Fig. 10). The growth rate of the RsHSP70–23-overexpressing cells showed no difference compared with that of EV under normal conditions, and under salt and heat stress, the optical density was higher than that of EV, whereas under chilling stress, the optical density was decreased compared with that of EV. These results indicated that overexpression of RsHSP70-23 can alleviate the adverse effects of high temperature and salinity stress on yeast cell growth.

Fig. 9
figure 9

Yeast wild type (EV) strain and RsHSP70-23 overexpressing cell were exposed to abiotic stresses (75 µM-Cd, 1 M-NaCl, 2 M-Mannitol, and cold and heat stresses. Triangles represent the tenfold serial dilutions (the starting OD600 is 0.1)

Fig. 10
figure 10

Growth curves of the RsHSP70-23 gene overexpressing yeast cells and EV (empty vector (yeast WT)) under normal, heat, salt and chilling stress. Cell density was monitored after 12, 14, 16, 18, 20, 22 and 24 h after the treatment. The error bar represents the deviation of three independent replications

Discussion

HSP70 proteins have been characterized as molecular chaperones expressed in a variety of plants, and most HSP70 family genes play important roles in the regulation of plant growth, development, and defense [16]. In the present study, 34 potential HSP70 genes were identified based on the R. sativus genome sequence, and were divided into six major subfamilies (Fig. 1 and Table 1). Within the same subfamilies, the most closely related RsHSP70 members shared more similarity in terms of motif composition, exon–intron organization, and the corresponding cellular compartments (Fig. 2). Notably, the number of HSP70 members varied among the species tested regardless of genome size and chromosome number (Fig. 1). Radish has twice as many HSP70s as Arabidopsis, which is inconsistent with a whole genome triplication (WGT) event that occurred in the Brassicaceae family [48], indicating that some RsHSP70 genes were lost after expansion. Segmental duplication was the main mechanism of RsHSP70 gene expansion, accounting for 32.35% of cases (Fig. 3). Segmental duplication occurred 2.49–7.58 Mya (Table 2), indicating that the duplication event of all pairs occurred after the divergence of Arabidopsis and Brassica at about 14.5–20.4 Mya [48], and also after divergence from cabbage and Chinese cabbage at approximately 7.1–10.4 Mya [49]. In addition, approximately 57.15% of segmentally duplicated genes had a Ka/Ks value < 0.1 (Table 2), indicating strong purifying selection, and these gene pairs may have become conserved and their functions tended to be constrained [31].

After duplication, genes may be maintained through subfunctionalization and/or neofunctionalization at the expression or sequence level. Alternatively, duplicated copies may accumulate deleterious mutations and become nonfunctional pseudogenes [50]. In this study, we found some sister pairs with changes in their exon–intron structures and numbers; RsHSP70–27 contained 2 exons, while its paralogs RsHSP70–4 and RsHSP70–11 had 11 and 4 exons, respectively (Fig. 2), indicating that exons were lost during evolution, similar to reports in potato and soybean [12, 51]. Large-scale synteny analysis showed that AT4G24280 had three orthologous genes in radish (RsHSP70–4, RsHSP70–11, and RsHSP70–27) (Fig. 4). RsHSP70–27 was not expressed in any of the tissues and/or stress conditions examined, suggesting that it is undergoing pseudogenization. However, RsHSP70–4 and RsHSP70–11 exhibited distinct expression patterns between both tissues and various abiotic stress conditions, indicating neofunctionalization (Figs. 5 and 6 and Tables S 5 and S 6). Moreover, there were 13 common syntenic genes between radish, Chinese cabbage, and Arabidopsis, thus providing a valuable reference for further understanding of the biological functions of these homologous genes in radish.

HSP70 genes are key components in plant development and in responses to a wide range of abiotic stresses [16]. We found that most RsHSP70s exhibited diverse expression profiles in different tissues, particularly in the 20-day leaves and 20-day roots of radish plants (Fig. 5 and Table S 5), suggesting that HSP70 family genes may play important roles in radish seedling development [47]. Under conditions of abiotic stress, transcription factors bind to the cis-regulatory elements of stress-responsive gene promoters and specifically initiate transcription of the corresponding genes [51]. In potato, most StHSP70s respond to various abiotic stresses (salt, drought, heat, and cold) and hormone treatments (ABA, IAA, GA3, and SA) [12]. More than half of the HSP70 genes are responsive to ABA, drought, and salt stresses in rice, Arabidopsis, and moss [52]. ABA (ABRE), MeJA (CGTCA-motif and TGACG-motif), SA (TCA element), drought (MBS), cold (LTR), and heat (HSE) response elements have been observed in the promoter regions of PyyHSP70 genes [53]. Here, a variety of hormone and stress response elements were found in the promoter regions of RsHSP70 family genes (Table S 7), where the numbers of HSEs were significantly greater than those of other elements (Table S 8). The two main subunits (5′NGAAN3′ and 5′NTTCN3′) of HSE are recognized by HSF1 [54]. As molecular chaperones, the most important biological function of HSP70s is related to acquired thermotolerance under heat stress, and their expression functions as a negative feedback regulator of heat shock transcription factor (HSF) activity [4,5,6,7]. Among the five stress conditions examined here, heat stress induced the greatest number of HSP70 genes in radish (Fig. 6A). These findings suggest that the RsHSP70 genes may respond to multiple hormones and abiotic stresses, particularly heat stress, in radish plants.

Cytosolic and ER-resident HSP70s also play essential regulatory roles in the innate immune response in plant cells [17]. AtBIP2 is localized to the ER and upregulated in response to Sclerotinia sclerotiorum, P. syringae, and FLG-22. In phylogenetic analysis, RsHSP70–23 clustered with AtBIP2 of Group B (Fig. 1), and large-scale synteny analysis also showed that RsHSP70–23 was orthologous with AtBIP2 (Fig. 4 and Table S 4). The expression levels of RsHSP70–23 in response to P. brassicae infection were significantly higher in clubroot-resistant than clubroot-susceptible lines (Fig. 7). In addition, BiP genes are induced by multiple abiotic stressors [17, 52]. We found that RsHSP70–23 showed a lower rate of expression in all tissues examined (Fig. 5) but was significantly upregulated under conditions of salt and heat stress (Figs. 6, 9 and 10). Surprisingly, analysis of the subcellular localization of RsHSP70–23-GFP fusion protein indicated exclusive cytoplasmic localization in Arabidopsis protoplasts (Fig. 8). These results suggest that the functions of RsHSP70–23 proteins may have been conserved after the divergence of radish and Arabidopsis but also exhibit unique functions through changes in localization during adaptation to changes in the environment.

Conclusions

In summary, 34 RsHSP70 genes were identified in the radish genome. Their physiochemical properties, phylogenetic relationships, gene organization, gene structures, chromosome distribution, and gene duplication were analyzed, and their expression patterns were characterized to understand their critical functions. These genes may play crucial roles in the growth, development, and stress responses of radish. In addition, RsHSP70–23 was localized to the cytoplasm and was involved in responses to certain abiotic stressors and P. brassicae infection. This comprehensive characterization of the RsHSP70 gene family will facilitate analysis of HSP70-gene mediated molecular mechanisms of stress responses in root vegetable crops.

Availability of data and materials

All the data obtained in the current study have been presented in this article.

References

  1. Wang WX, Vinocur B, Shoseyov O, Altman A. Role of plant heat-shock proteins and molecular chaperones in the abiotic stress response. Trends Plant Sci. 2004;9(5):244–52.

    Article  PubMed  CAS  Google Scholar 

  2. Dragovic Z, Broadley SA, Shomura Y, Bracher A, Hartl FU. Molecular chaperones of the Hsp110 family act as nucleotide exchange factors of Hsp70s. Embo J. 2006;25(11):2519–28.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Karlin S, Brocchieri L. Heat shock protein 70 family: multiple sequence comparisons, function, and evolution. J Mol Evol. 1998;47(5):565–77.

    Article  PubMed  CAS  Google Scholar 

  4. Lee JH, Schöffl F. An Hsp70 antisense gene affects the expression of HSP70/HSC70, the regulation of HSF, and the acquisition of thermotolerance in transgenic Arabidopsis thaliana. Mol Gen Genet. 1996;252(1–2):11–9.

    PubMed  CAS  Google Scholar 

  5. Qi YC, Wang HJ, Zou Y, Liu C, Liu YQ, Wang Y, Zhang W. Over-expression of mitochondrial heat shock protein 70 suppresses programmed cell death in rice. FEBS Lett. 2011;585(1):231–9.

    Article  PubMed  CAS  Google Scholar 

  6. Kim BH, Schöffl F. Interaction between Arabidopsis heat shock transcription factor 1 and 70 kDa heat shock proteins. J Exp Bot. 2002;53(367):371–5.

    Article  PubMed  CAS  Google Scholar 

  7. Yang KZ, Xia C, Liu XL, Dou XY, Wang W, Chen LQ, Zhang XQ, Xie LF, He L, Ma X, Ye D. A mutation in Thermosensitive Male Sterile 1, encoding a heat shock protein with DnaJ and PDI domains, leads to thermosensitive gametophytic male sterility in Arabidopsis. Plant J. 2009;57(5):870–82.

    Article  PubMed  CAS  Google Scholar 

  8. Su PH, Li HM. Arabidopsis stromal 70-kD heat shock proteins are essential for plant development and important for thermotolerance of germinating seeds. Plant Physiol. 2008;146(3):1231–41.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Jungkunz I, Link K, Vogel F, Voll LM, Sonnewald S, Sonnewald U. AtHsp70-15-deficient Arabidopsis plants are characterized by reduced growth, a constitutive cytosolic protein response and enhanced resistance to TuMV. Plant J. 2011;66(6):983–95.

    Article  PubMed  CAS  Google Scholar 

  10. Kim SR, An G. Rice chloroplast-localized heat shock protein 70, OsHsp70CP1, is essential for chloroplast development under high-temperature conditions. J Plant Physiol. 2013;170(9):854–63.

    Article  PubMed  CAS  Google Scholar 

  11. Tabassum R, Dosaka T, Ichida H, Morita R, Ding Y, Abe T, Katsube-Tanaka T. FLOURY ENDOSPERM11–2 encodes plastid HSP70–2 involved with the temperature-dependent chalkiness of rice (Oryza sativa L.) grains. Plant J. 2020;103(2):604–16.

    Article  PubMed  CAS  Google Scholar 

  12. Guo M, Lu JP, Zhai YF, Chai WG, Gong ZH, Lu MH. Genome-wide analysis, expression profile of heat shock factor gene family (CaHsfs) and characterisation of CaHsfA2 in pepper (Capsicum annuum L). BMC Plant Biology. 2015;15(1):1–20.

    Article  Google Scholar 

  13. Davoudi M, Chen JF, Lou QF. Genome-wide identification and expression analysis of heat shock protein 70 (HSP70) gene family in pumpkin (Cucurbita moschata) rootstock under drought stress suggested the potential role of these chaperones in stress tolerance. Int J Mol Sci. 2022;23:1422–0067 ((Electronic)):1918).

    Article  Google Scholar 

  14. Liu J, Pang X, Cheng Y, Yin YH, Zhang Q, Su WB, Hu B, Guo QW, Ha S, Zhang JP, Wan HJ. The Hsp70 gene family in solanum tuberosum: genome-wide identification, phylogeny, and expression patterns. Sci Rep. 2018;8(1):16628.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Li H, Liu SS, Yi CY, Wang F, Zhou J, Xia XJ, Shi K, Zhou YH, Yu JQ. Hydrogen peroxide mediates abscisic acid-induced HSP70 accumulation and heat tolerance in grafted cucumber plants. Plant Cell Environ. 2014;37(12):2768–80.

    Article  PubMed  CAS  Google Scholar 

  16. Ul Haq S, Khan A, Ali M, Khattak AM, Gai WX, Zhang HX, Wei AM, Gong ZH. Heat shock proteins: dynamic biomolecules to counter plant biotic and abiotic stresses. Int J Mol Sci. 2019;20(21):5321.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Berka M, Kopecká R, Berková V, Brzobohatý B, Černý M. Regulation of heat shock proteins 70 and their role in plant immunity. J Exp Bot. 2022;73(7):1894–909.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Chen ZR, Zhou T, Wu XH, Hong YG, Fan ZF, Li HF. Influence of cytoplasmic heat shock protein 70 on viral infection of Nicotiana benthamiana. Mol Plant Pathol. 2008;9(6):809–17.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  19. Jelenska J, van Hal JA, Greenberg JT. Pseudomonas syringae hijacks plant stress chaperone machinery for virulence Proc Natl Acad Sci U S A. 2010;107(29):13177–82.

    Article  PubMed  CAS  Google Scholar 

  20. Kim NH, Hwang BK. Pepper heat shock protein 70a interacts with the type III effector AvrBsT and triggers plant cell death and immunity. Plant Physiol. 2015;167(2):307–32.

    Article  PubMed  CAS  Google Scholar 

  21. Jiang SS, Lu YW, Li KF, Lin L, Zheng HY, Yan F, Chen JP. Heat shock protein 70 is necessary for rice stripe virus infection in plants. Mol Plant Pathol. 2015;15(9):907–17.

    Article  CAS  Google Scholar 

  22. Wang RH, Mei Y, Xu L, Zhu XW, Wang Y, Guo J, Liu LW. Genome-wide characterization of differentially expressed genes provides insights into regulatory network of heat stress response in radish ( Raphanus sativus L.). Funct Integr Genomics. 2018;18(2):225–39.

    Article  PubMed  CAS  Google Scholar 

  23. Sun XC, Xu L, Wang Y, Luo XB, Zhu XW, Kinuthia KB, Nie SS, Feng HY, Li C, Liu LW. Transcriptome-based gene expression profiling identifies differentially expressed genes critical for salt stress response in radish (Raphanus sativus L). Plant Cell Rep. 2016;35(2):329–46.

    Article  PubMed  CAS  Google Scholar 

  24. Xu L, Wang Y, Liu W, Wang J, Zhu XW, Zhang KY, Yu RG, Wang RH, Xie Y, Zhang W, Gong YQ, Liu LW. De novo sequencing of root transcriptome reveals complex cadmium-responsive regulatory networks in radish (Raphanus sativus L.). Plant Sci. 2015;236(313):323.

    Google Scholar 

  25. Xie Y, Shan Y, Wang Y, Liang X, Zhu XW, Yang JL, Feng HY, Yu RG, Karanja B, Gong YQ, Liu LW. Transcriptome-based gene profiling provides novel insights into the characteristics of radish root response to Cr stress with next-generation sequencing. Front Plant Sci. 2015;6:202.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Wang Y, Xu L, Chen YL, Shen H, Gong YQ, Limera C, Liu LW. Transcriptome profiling of radish (Raphanus sativus L.) root and identification of genes involved in response to lead (Pb) stress with next generation sequencing. PLoS ONE. 2013;8(6):e66539.

  27. Shirasawa K, Hirakawa H, Fukino N, Kitashiba H, Isobe S. Genome sequence and analysis of a Japanese radish (Raphanus sativus) cultivar named 'Sakurajima Daikon' possessing giant root. DNA Res. 2020;27(2):1–6.

  28. Paysan-Lafosse T, Blum M, Chuguransky S, Grego T, Pinto BL, Salazar GA, Bileschi ML, Bork P, Bridge A, Colwell L, Gough J, Haft DH, Letunić I, Marchler-Bauer A, Mi H, Natale DA, Orengo CA, Pandurangan AP, Rivoire C, Sigrist CJA, Sillitoe I, Thanki N, Thomas PD, Tosatto SCE, Wu CH, Bateman A. InterPro in 2022. Nucleic Acids Res. 2023;51(D1):D418–27.

    Article  PubMed  CAS  Google Scholar 

  29. Berardini TZ, Reiser L, Li DH, Mezheritsky Y, Muller R, Strait E, Huala E. The arabidopsis information resource: Making and mining the “gold standard” annotated reference plant genome. genesis. 2015;53(8):474–85.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Letunic I, Doerks T, Bork P. SMART: recent updates, new developments and status in 2015. Nucleic Acids Res. 2015;43(D1):D257–60.

    Article  PubMed  CAS  Google Scholar 

  31. Liang ZW, Li MD, Liu ZY, Wang JB. Genome-wide identification and characterization of the Hsp70 gene family in allopolyploid rapeseed (Brassica napus L.) compared with its diploid progenitors. PeerJ. 2019;7:e7511. https://0-doi-org.brum.beds.ac.uk/10.7717/peerj.7511.

  32. Gasteiger E, Hoogland C, Ivanyi I, Appel RD, Bairoch A. ExPASy: the proteomics server for in-depth protein knowledge and analysis. Nucleic Acids Res. 2003;31(13):3784–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Horton P, Park KJ, Obayashi T, Fujita N, Harada H, Adams-Collier CJ, Nakai K. WoLF PSORT: Protein Localization Prediction Software. Nucleic Acids Res. 2006;35:W585–7.

    Article  Google Scholar 

  34. Kumar S, Stecher G, Tamura K. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol. 2016;33(7):1870–4.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Hu Jin, Jin JP, Guo AY, Zhang H, Luo J, Gao G. GSDS 2.0: an upgraded gene feature visualization server. Bioinformatics. 2015;31(8):1296–7.

    Article  PubMed  Google Scholar 

  36. Bailey TL, Williams N, Misleh C, Li WW. MEME: discovering and analyzing DNA and protein sequence motifs. Nucleic Acids Res. 2006;34:W369–73.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Chen CJ, Chen H, Zhang Y, Thomas HR, Frank MH, He YH, Xia R. TBtools. An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol Plant. 2020;13(8):1194–202.

    Article  PubMed  CAS  Google Scholar 

  38. Song HF, Fu XX, Li J, Niu TZ, Shen J, Wang X, Li YL, Hou QW, Liu AK. Phylogenetic analysis and expression profiles of jasmonate ZIM-domain gene family provide insight into abiotic stress resistance in sunflower. Front Plant Sci. 2022;13:1010404.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Krzywinski M, Schein J, Birol I, Connors J, Gascoyne R, Horsman D, Jones SJ, Marra MA. Circos: an information aesthetic for comparative genomics. Genome Res. 2009;19(9):1639–45.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Wang YP, Tang HB, Debarry JD, Tan X, Li JP, Wang XY, Lee TH, Jin HZ, Marler B, Guo H, Kissinger JC, Paterson AH. MCScanX: a toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012;40(7): e49.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. de Sá Moniz M, Drouin G. Phylogeny and substitution rates of angiosperm actin genes. Mol Biol Evol. 1996;13(9):1198–212.

    Article  Google Scholar 

  42. Mitsui Y, Shimomura M, Komatsu K, Namiki N, Shibata-Hatta M, Imai M, Katayose Y, Mukai Y, Kanamori H, Kurita K, Kagami T, Wakatsuki A, Ohyanagi H, Ikawa H, Minaka N, Nakagawa K, Shiwa Y, Sasaki T. The radish genome and comprehensive gene expression profile of tuberous root formation and development. Sci Rep. 2015;5:10835.

  43. Liu AK, Liu CL, Lei HY, Wang ZJ, Zhang M, Yang XR, Yang G, Ren JH. Phylogenetic analysis and transcriptional profiling of WRKY genes in sunflower (Helianthus annuus L.): Genetic diversity and their responses to different biotic and abiotic stresses. Ind Crops Prod. 2020;148:112268. https://0-doi-org.brum.beds.ac.uk/10.1016/j.indcrop.2020.112268.

  44. Galindo-González L, Manolii V, Hwang SF, Strelkov SE. Response of Brassica napus to Plasmodiophora brassicae Involves Salicylic Acid-Mediated Immunity: An RNA-Seq-Based Study. Front Plant Sci. 2020;11:1025.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Xu YY, Zhu XW, Gong YQ, Xu L, Wang Y, Liu LW. Evaluation of reference genes for gene expression studies in radish (Raphanus sativus L.) using quantitative real-time PCR. Biochem Biophys Res Commun. 2012;424(3):398–403.

  46. Anwar A, Zhang S, Wang LX, Wang FD, He LL, Gao JW. Genome-wide identification and characterization of chinese cabbage S1fa transcription factors and their roles in response to salt stress. Antioxidants (Basel). 2022;11(9):1782.

    Article  PubMed  CAS  Google Scholar 

  47. Sung DY, Vierling E, Guy CL. Comprehensive expression profile analysis of the Arabidopsis Hsp70 gene family. Plant Physiol. 2001;126(2):789–800.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Yang YW, Lai KN, Tai PY, Li WH. Rates of nucleotide substitution in angiosperm mitochondrial DNA sequences and dates of divergence between Brassica and other angiosperm lineages. J Mol Evol. 1999;48(5):597–604.

    Article  PubMed  CAS  Google Scholar 

  49. Xu L, Wang Y, Dong JH, Zhang W, Tang MJ, Zhang WL, Wang K, Chen YL, Zhang XL, He Q, Zhang XY, Wang K, Wang L, Ma YB, Xia K, Liu LW. A chromosome-level genome assembly of radish (Raphanus sativus L.) reveals insights into genome adaptation and differential bolting regulation. Plant Biotechnol J. 2023;21(5):990–1004.

  50. Birchler JA, Yang H. The multiple fates of gene duplications: Deletion, hypofunctionalization, subfunctionalization, neofunctionalization, dosage balance constraints, and neutral variation. Plant Cell. 2022;34(7):2466–74.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Zhang L, Zhao HK, Dong QL, Zhang YY, Wang YM, Li HY, Xing GJ, Li QY, Dong YS. Genome-wide analysis and expression profiling under heat and drought treatments of HSP70 gene family in soybean (Glycine max L.). Front Plant Sci. 2015;6:33650. https://0-doi-org.brum.beds.ac.uk/10.1038/srep33650.

  52. Tang T, Yu AM, Li P, Yang H, Liu GJ, Liu L. Sequence analysis of the Hsp70 family in moss and evaluation of their functions in abiotic stress responses. Sci Rep. 2016;6:33650.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Yu XZ, Mo ZL, Tang XH, Gao T, Mao YX. Genome-wide analysis of HSP70 gene superfamily in Pyropia yezoensis (Bangiales, Rhodophyta): identification, characterization and expression profiles in response to dehydration stress. BMC Plant Biol. 2021;21(1):435.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  54. Rossi A, Trotta E, Brandi R, Brandi R, Arisi I, Coccia M, Santoro MG. AIRAP, a new human heat shock gene regulated by heat shock factor 1. J Biol Chem. 2010;285:13607–15.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Funding

The research was financially supported by the Special Construction Project of National Modern Agricultural Industrial Technology System (CARS-23); the Technology Innovation and Application Development Program of Chongqing (cstc2019jscx-gksbX0143); the application of biotechnology in the precise identification and utilization of superior crop germplasm resources (cqaas2023sjczhx016).

Author information

Authors and Affiliations

Authors

Contributions

Na Zhou conceived and designed the experiments. Xaioxue Pan carried out the experiments. Yang Zheng, Keirong Lei and WeiLin Tao analyzed the data. Xiaoxue Pan and Na Zhou wrote and revised the paper. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Na Zhou.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1:

Figure S1. Multiple sequence alignment of HSP70 proteins.

Additional file 2:

Table S1. The accession number of RNAseq data.

Additional file 3:

Table S2. Primers and corresponding sequences used in the study.

Additional file 4:

Table S3. The MEME motif sequences and lengths of HSP70 gene family proteins in radish.

Additional file 5:

Table S4. The orthologous and paralogous gene pairs of HSP70 proteins among the radish, Arabidopsis and Chinese cabbage.

Additional file 6:

Table S5. The TPM expression values of RsHSP70 genes in various tissues.

Additional file 7:

Table S6. Significantly-regulated DEGs at all the stress.

Additional file 8:

Table S7. Summary of abiotic stresses inducible cis-elements is in the promoter regions of HSP70 family genes in radish.

Additional file 9:

Table S8. The detailed information related to the HSF1-binding motif provided by PlantPAN3 online tool.

Additional file 10:

Table S9. Significantly-regulated DEGs under P. brassicae infection.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pan, X., Zheng, Y., Lei, K. et al. Systematic analysis of Heat Shock Protein 70 (HSP70) gene family in radish and potential roles in stress tolerance. BMC Plant Biol 24, 2 (2024). https://0-doi-org.brum.beds.ac.uk/10.1186/s12870-023-04653-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12870-023-04653-6

Keywords