Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Transcriptomic Analysis of Cadmium Stress Response in the Heavy Metal Hyperaccumulator Sedum alfredii Hance

  • Jun Gao ,

    Contributed equally to this work with: Jun Gao, Ling Sun

    Affiliation MOE Key Laboratory of Environment Remediation and Ecological Health, College of Natural Resource & Environmental Sciences, Zhejiang University, Hangzhou, China

  • Ling Sun ,

    Contributed equally to this work with: Jun Gao, Ling Sun

    Affiliation State Key Laboratory of Genetic Engineering, Institute of Plant Biology, School of Life Sciences, Fudan University, Shanghai, China

  • Xiaoe Yang ,

    jianxiangliu@fudan.edu.cn (JXL); xyang581@gmail.com (XY)

    Affiliation MOE Key Laboratory of Environment Remediation and Ecological Health, College of Natural Resource & Environmental Sciences, Zhejiang University, Hangzhou, China

  • Jian-Xiang Liu

    jianxiangliu@fudan.edu.cn (JXL); xyang581@gmail.com (XY)

    Affiliation State Key Laboratory of Genetic Engineering, Institute of Plant Biology, School of Life Sciences, Fudan University, Shanghai, China

Abstract

The Sedum alfredii Hance hyperaccumulating ecotype (HE) has the ability to hyperaccumulate cadmium (Cd), as well as zinc (Zn) and lead (Pb) in above-ground tissues. Although many physiological studies have been conducted with these plants, the molecular mechanisms underlying their hyper-tolerance to heavy metals are largely unknown. Here we report on the generation of 9.4 gigabases of adaptor-trimmed raw sequences and the assembly of 57,162 transcript contigs in S. alfredii Hance (HE) shoots by the combination of Roche 454 and Illumina/Solexa deep sequencing technologies. We also have functionally annotated the transcriptome and analyzed the transcriptome changes upon Cd hyperaccumulation in S. alfredii Hance (HE) shoots. There are 110 contigs and 123 contigs that were up-regulated (Fold Change ≧2.0) and down-regulated (Fold Change ≦0.5) by chronic Cd treatment in S. alfredii Hance (HE) at q-value cutoff of 0.005, respectively. Quantitative RT-PCR was employed to compare gene expression patterns between S. alfredii Hance (HE) and non-hyperaccumulating ecotype (NHE). Our results demonstrated that several genes involved in cell wall modification, metal translocation and remobilization were more induced or constitutively expressed at higher levels in HE shoots than that in NHE shoots in response to Cd exposure. Together, our study provides large-scale expressed sequence information and genome-wide transcriptome profiling of Cd responses in S. alfredii Hance (HE) shoots.

Introduction

The non-essential heavy metal cadmium (Cd) in soils is very mobile and readily enters plant tissues, the food chain and drinking water. The major concern about Cd pollution is its potential threat to human health with the high risk of causing cancer including lung, bladder, renal, prostate, and breast cancer [1], [2]. Cd interferes with cell proliferation, differentiation, DNA replication and repair, as well as protein synthesis and folding [2]. Cd is also toxic to plant tissues, adversely affecting the photosynthetic apparatus, carbohydrate metabolism and nitrate absorption [3]. Tolerant plants differ in the ways in which they contend with Cd. “Excluders” tolerate heavy metals by limiting the entry and root-to-shoot translocation of trace metals. On the other hand, hyper-accumulating plants can accumulate extremely high levels of heavy metals in their above-ground tissues without exhibiting toxicity symptoms. Hyper-accumulating plants show promise for use in phytomediation, i.e., for cleaning up heavy metal contaminated sites and reducing the bioavailability of heavy metals in the environment [4]. However, the use of hyperaccumulators for this purpose has been limited due to their low biomass, lack of metal selectivity or poor agronomic practice. Understanding the molecular mechanisms of hyperaccumulation may help in enhancing the performance of hyperaccumulators for phytoremediation [5].

So far over 450 heavy metal accumulators have been identified, most of which (75%) are nickel (Ni) tolerant [6]. Cd hyperaccumulators have been defined as plants that accumulate and tolerate of at least 100 mg kg−1 dry weight (DW) in shoots of plants [6]. Cd hyperaccumulation occurs in four species of Brassicaceae and Crassulaceae family. Thlaspi caerulescens, also known as Noccaea caerulescens, and Arabidopsis halleri are two of the best studied hyperaccumulating species in Brassicaceae family, and the availability of Arabidopsis thaliana genome sequence information has helped in identify genes associated with heavy metal tolerance in its close relatives, T. caerulescens and A. halleri [7], [8]. A. halleri has been found to have a high constitutive level of Zn/Cd transporter HMA4 (P-type ATPase) expression apparently due to a modification in cis-regulatory sequences and gene copy number expansion [9]. In other species, heavy metal hyper-tolerance is mainly associated with vacuolar sequestration and metal chelation, both of which are heavy metal inducible [10]. Recently, the constitutive, high-level expression of a tonoplast-localized transporter HMA3 (ATPase) was reported to be associated with specific Cd sequestration into vacuoles in T. caerulescens leaves [11].

Sedum alfredii Hance hyperaccumulating ecotype (HE) is the only hyperaccumulating species in Crassulaceae family identified to date, originally found as a Zn hyperaccumulator in a soil-Pb/Zn-rich region in China [12]. It is not only a Zn/Cd cohyperaccumulator, but also highly tolerant to copper (Cu) and Pb toxicity [13], [14], [15]. A number of physiological studies have been carried out to understand the hyper-tolerance and hyperaccumulation mechanisms in S. alfredii Hance (HE) [16], [17], [18], [19]. In contrast to T. caerulescens and A. halleri, the molecular basis for hyper-tolerance in S. alfredii Hance (HE) is not known, due in part to our lack knowledge about the genome of this organism.

The development of high-throughput deep sequencing technology has enabled the large-scale RNA-seq of dynamic transcriptomes, even without fully sequenced reference genomes [20], [21]. In the present study, we have employed RNA-seq to explore the transcriptome of S. alfredii Hance (HE) and to identify transcriptional changes in response to high Cd accumulation in shoots. We also compared the expression levels of selected genes in shoots of Sedum alfredii Hance (HE) and the non-hyperaccumulating ecotype (NHE). We observed that several genes involved in cell wall modification and metal translocation were more induced by Cd or constitutively expressed at higher levels under normal growth condition in HE than that in NHE. Our data will help in furthering our understanding of the molecular mechanisms of heavy metal hyper-tolerance in plants.

Results

Investigation the Transcriptome of S. alfredii Hance (HE) Shoots

S. alfredii Hance (HE) can accumulate more than 5,000 mg kg−1 DW of Cd in shoots without any observable symptoms of toxicity when grown hydroponically (Figure 1A–B). S. alfredii Hance (HE) shoots were used in most of these studies since Cd content in shoots was much higher than that in roots (Figure 1B), and high-quality RNA was better obtained from shoots than roots in this species. Initially, RNAs from Cd treated (Cd) or untreated control (Cont) samples were pooled and sequenced on the Roche’s 454 sequencing platform. Pooling of the sequencing runs of Cd and Cont samples resulted in 135,894 qualified reads with an average length of 287 bases (b) after removal of adaptor sequences (Figure S1A), which were equal to 39.1 millionbases (Mb) in total. Subsequently, Illumina/Solexa cDNA libraries were constructed and sequenced from Cd and Cont samples with three biological replicates, respectively. In total, around 9.4 gigabases (Gb) of adaptor-trimmed raw sequences representing 47.5 million qualified reads with an average of 99 b were obtained from Illumina/Solexa sequencing (Table S1). The 454 reads were firstly assembled with Newbler to generate 4,193 EST clusters (isotigs) containing 83,333 reads (61.3% of total), which provided a reference framework for the later assembly of Illumina/Solexa sequences. First, 47,504,847 Illumina/Solexa reads (22.9% of total) were mapped to 454-derived isotigs; the rest of the Illumina/Solexa reads were separately assembled using Velvet/Oases [22] together with the 454 singleton reads, yielding 119,731 EST clusters (contigs). The aforementioned isotigs and contigs were finally re-assembled in CAP3 [23] to generate the transcriptome of S. alfredii Hance (HE) shoots which contains 57,162 contigs with lengths longer than 200 b (Figure S1B).

thumbnail
Figure 1. Cd hyperaccumulation in S. alfredii Hance (HE).

(A) Growth of HE after exposure to 0 µM (Cont) or 100 µM Cd2+ (Cd) for 8 days. (B) Cd content (mg kg−1 dry weight) in shoots and roots of HE. Bars depict SE (n = 3).

https://doi.org/10.1371/journal.pone.0064643.g001

Functional Annotation of the S. alfredii Hance (HE) Transcriptome

Gene prediction was carried out with the GETORF program using the EMBOSS package [24] and 57,115 protein-coding contigs were identified. All the predicted proteins were compared with the non redundant protein sequence databases in Swiss-Prot and GeneBank using BlastP searches. A total of 28,578 contigs were functionally annotated with an E-value cut-off of 1E-3. Gene Ontology (GO) analysis was subsequently performed with GoPipe [25]. In total, 14,046 contigs were matched to 75,775 GO terms (Figure 2A–C). The largest cellular component for S. alfredii Hance (HE) proteome represented intracellular components (35.9%, Figure 2A). The majority of biological processes identified were involved in metabolic processes (25.3%, Figure 2B) and macromolecule metabolic processes (15.8%, Figure 2B). Most of the molecular functions were associated with binding (25.0%, Figure 2C) and catalytic activities (20.3%, Figure 2C). The largest group in KEGG Orthology (KO, Figure 3) was related to signal transduction mechanisms (11.7%), and the smallest group was cell motility (0.1%). In total, 4,380 of the contigs (7.7%) were associated with posttranslational modification, protein turnover and chaperones while 3458 contigs (6.1%) were related to transcription. Totally 649 contigs encoding putative transcription factors were identified; among them, zinc finger proteins (16.3%) and MYB (16.2%) were the most abundant followed by bHLH (12.5%) and bZIP transcription factors (8.3%). Pathway enrichment of the transcriptome was also conducted according to GO annotations for KEGG. 5,023 contigs were found to be involved in 37 different pathways (Figure S2). When compared to protein databases from six sequenced model plant species (Chlamydomonas reinhardtii, Arabidopsis thaliana, Oryza sativa, Zea mays, Medicago truncatula and Vitis vinifera), the S. alfredii Hance (HE) proteome was most similar to the V. vinifera proteome, followed by the M. truncatula proteome (Figure 4A–B), both of which are in dicotyledous plants in the same Rosidae subclass as S. alfredii Hance. Given a BlastP E-value cutoff of 1E-30, 47.7% of V. vinifera proteins were found to be homologous to 39.4% of S. alfredii Hance proteins/contigs, whereas 16.9% of M. truncatula proteins were similar to 33.0% of S. alfredii Hance proteins/contigs. The green alga C. reinhardtii contigs had the lowest overall similarity to the S. alfredii Hance contigs (Figure 4B). Microsatellites, also known as simple sequence repeats (SSRs), are widely recognized as useful co-dominant, locus-specific markers for DNA fingerprinting, genome mapping and phylogenetic analysis [26]. Among the 57,162 contigs in S. alfredii Hance (HE), 6,176 perfect SSRs and 3,019 imperfect SSRs were found (Dataset S1). Many species in Crassulaceae family are CAM plants, in which water use efficiency is increased and malic acid level is elevated [27]. Phosphoenolpyruvate carboxylase (PEPC) is a key enzyme in CAM and C4 metabolism and has been used as a marker for phylogenetic classification among C3, C4 and CAM plants [28]. Twenty contigs encoding PEPC homologs were found in the S. alfredii Hance (HE) transcriptome (Table S3) and phylogenetic analysis revealed that S. alfredii Hance (HE) had both C3- and CAM-type PEPCs. Among the S. alfredii Hance PEPCs, one contig (Sa_Contig08207) was most closely related to CAM-type PEPC in K. blossfeldiana, which is also in the Crassulaceae family (Figures S3S4). These results indicate that S. alfredii Hance (HE) are CAM plants, and our deep sequencing data may be useful for CAM studies in the future.

thumbnail
Figure 2. Gene Ontology (GO) analysis of S. alfredii Hance (HE) contigs.

GO distribution of (A) cellular component, (B) biological process and (C) molecular function.

https://doi.org/10.1371/journal.pone.0064643.g002

thumbnail
Figure 3. Proportion of S. alfredii Hance (HE) contigs in different functional categories.

KEGG (Kyoto Encyclopedia of Genes and Genomes) Orthology Groups (KOG) analysis was applied to the assembled contigs.

https://doi.org/10.1371/journal.pone.0064643.g003

thumbnail
Figure 4. Homologous protein comparisons of S. alfredii Hance (HE) to other sequenced model plants.

(A) Distribution of S. alfredii homologous proteins identified in other plants using several BlastP E-value cutoffs. (B) Comparison of the number of homologs identified between S. alfredii and other plant proteomes with BlastP E-value cutoff of 1E-30. The number of contigs in each species is displayed in the same color as the species name.

https://doi.org/10.1371/journal.pone.0064643.g004

Comparative Transcriptome Analysis in S. alfredii Hance (HE) Shoots

Long exposure of S. alfredii Hance (HE) to Cd was employed in the present study to investigate the molecular mechanisms of Cd hyper-tolerance to minimize the interference of acute phase responses. The expression level of each contig was calculated and normalized to RPKM (Reads Per Kilobase of exon model per Million total reads in sample) according to standard protocols [29]. Compared with the Cont, totally 110 contigs were upregulated and 123 contigs were down-regulated by Cd treatment (FC≧2.0 and FC≦0.5, respectively) at a q-value cutoff of 0.005 (Table S2). The upregulated contigs were grouped into five categories according to their functional annotations (Table 1). One category included 20 contigs, which were related to phenylpropanoid biosynthesis, cell wall deposition and modification. Another category included 4 contigs for heavy metal efflux and 7 contigs related to metal ligand synthesis and metal-ligand transport. There were 6 contigs (2 peroxiredoxins and 4 peroxidases) in a 3rd category, which were considered to be involved in reactive oxygen species (ROS) detoxification. Another 6 contigs in the 4th category represented four types of transcription factors. The remaining 70 contigs with unknown functions or with functions involved in other processes were grouped into the 5th category and listed in Table S2. We focused on the up-regulated contigs in later studies although the down-regulated contigs shown in Table S2 might also be important.

thumbnail
Table 1. Expression levels of Cd up-regulated contigs in HE shoots.

https://doi.org/10.1371/journal.pone.0064643.t001

To evaluate the validity of Illumina/Solexa data, 15 Cd-induced genes were randomly selected from Table 1 and their expression levels in Cont and Cd were examined by quantitative RT-PCR (qRT-PCR). As expected, the expression pattern of those contigs obtained from qRT-PCR was very highly correlated to the Illumina/Solexa sequencing results with a correlation coefficient of 0.9235 (Figure S5).

Comparison of Gene Expression in Shoots of Two Contrasting S. alfredii Hance Ecotypes

Previously, the physiological properties of the non-hyperaccumulating ecotype (NHE) was compared with the hyperaccumulating ecotype (HE) when grown on media supplemented with different concentrations of Cd [33]. In the current study, the HE and NHE plants were cultivated hydroponically and supplied with subtoxic Cd concentration for the same period [8], [30]. No morphological difference was observed after Cd exposure. qRT-PCR was employed to compare the gene expression levels between HE and NHE plants in response to prolonged Cd exposure. Several genes were selected based on the current RNA-seq results and three previously reported metal-tolerance-associated genes were also included (Figure 5A). Among them, four contigs (Sa_Contig14694, laccase 1–1; Sa_Contig01582, laccase 1–2; Sa_Contig13861, cinnamate 4-hydroxylase; Sa_Contig03765, metal transporter NRAMP2) were more induced in HE than that in NHE (P<0.05); two contigs (Sa_Contig43609, SAM synthetase; Sa_Contig12362, cellulose synthase) were up-regulated in HE while down-regulated in NHE. In contrast, three contigs (Sa_Contig06552, metal-nicotianamine transporter YSL3; Sa_Contig30461, metal transporter NRAMP3; Sa_Contig11685, P-type metal ATPase HMA4) were more induced in NHE than that in HE (P<0.05). Eight other contigs (Sa_Contig12607, 1-Cis peroxiredoxin; Sa_Contig00501, tyrosine aminotransferase TAT; Sa_Contig15461, endo-1,3-beta-glucanase; Sa_Contig09892, probable peptide transporter; Sa_Contig14529, metal transporter NRAMP4; Sa_Contig08243, ABC transporter B family member 19; Sa_Contig08963, Oligopeptide transporter 3; Sa_Contig10290, zinc transporter ZIP1) were up-regulated by Cd with no significant difference between two ecotypes (P<0.05). Interestingly, one contig (Sa_Contig47062, metal tolerance protein MTP3) was down-regulated in HE while up-regulated in NHE. In order to know whether Cd hyper-tolerance in HE is correlated to constitutive higher expression of certain genes, the expression levels of 18 aforementioned contigs were compared between HE and NHE under normal growth condition (Figure 5B). Except 1-Cis peroxiredoxin (Sa_Contig12607), metal-nicotianamine transporter YSL3 (Sa_Contig06552) and cinnamate 4-hydroxylase (Sa_Contig13861), of which expression levels were higher in NHE than that in HE (P<0.05), nine other contigs (Sa_Contig01582, laccase 1–2; Sa_Contig00501, tyrosine aminotransferase TAT; Sa_Contig47062, metal tolerance protein MTP3; Sa_Contig43609, SAM synthetase; Sa_Contig15461, endo-1,3-beta-glucanase; Sa_Contig14529, metal transporter NRAMP4; Sa_Contig10290, zinc transporter ZIP1; Sa_Contig11685, P-type metal ATPase HMA4; Sa_Contig12362, cellulose synthase) related to metal transport and cell wall modification process were constitutively expressed at higher levels in HE than that in NHE (P<0.05). The rest of the six contigs (Sa_Contig14694, laccase 1–1; Sa_Contig03765, metal transporter NRAMP2; Sa_Contig30461, metal transporter NRAMP3; Sa_Contig09892, probable peptide transporter; Sa_Contig08243, ABC transporter B family member 19; Sa_Contig08963, Oligopeptide transporter 3) expressed at similar level in both ecotypes under normal growth condition (P<0.05).

thumbnail
Figure 5. Comparison of gene expression in S. alfredii Hance hyper-accumulation ecotype (HE) and non-hyperaccumulation ecotype (NHE).

(A) Transcriptional responses of two ecotypes to Cd exposure. Fold change of expression is the relative gene expression level in Cd treatment divided by the value in the control. (B) Relative gene expression levels of two ecotypes under normal growth condition. The relative gene expression is the expression level normalized to a constitutively expressed gene Actin. P-type metal ATPase HMA4: Sa_Contig11685; Zinc transporter ZIP1: Sa_Contig10290; Metal tolerance protein MTP3: Sa_Contig47062. The contig numbers for other genes are listed in Table 1. P values are indicated as follows: *P<0.05; **P<0.01; ***P<0.001. Bars depict SE (n = 3).

https://doi.org/10.1371/journal.pone.0064643.g005

Discussion

Heavy metal hyperacculuators have extreme lifestyles in which high levels of toxic elements are accumulated in these plants without obvious toxicity symptoms. The Crassulaceae family plant S. alfredii Hance (HE) is the only identified Zn and Cd hyperaccumulating plant to date that does not belong to Brassicaceae family [6].The major strategy hyperaccumulators have used is to protect themselves by compartmentalization of metal ions in the cell. Several contigs related to cell wall deposition and modifications were identified as Cd responsive genes in S. alfredii Hance (HE) shoots in the current study. Lignin is a complex phenylpropanoid polymer deposited in the secondary cell walls [31]. The formation of phenylpropanoids starts from phenylalanine and involves three important enzymes: phenylalanine ammonia-lyase, cinnamate 4-hydroxylase, and 4-coumaroyl CoA-Ligase. Cinnamyol alcohol dehydrogenase is considered as the key enzyme in linking the central metabolite 4-coumaroyl CoA to lignin production. Tyrosine aminotransferase catalyzes the conversion of tyrosine to 4-hydroxyphenylpyruvic acid, which is an intermediate in the metabolism of phenylalanine [32]. The ATP-binding cassette (ABC) subfamily B transporters (ABCB) and multi-copper-containing glycoprotein laccases have also been demonstrated to be associated with cell wall lignifications [33], [34]. Contigs encoding the above-mentioned six enzymes and one ABCB gene were up-regulated by Cd in HE shoots (Table 1). Lignin accumulates between cellulose, hemicellulose and pectin components in the cell wall. Cellulose synthase (CESA) forms large membrane complexs and is responsible for the cellulose formation. Xyloglucan is the major hemicellulose in cell walls and crosslinks microfibrils. Xyloglucan endotransglucosylase (XTH) has the ability to breakdown xyloglucan strands that are not tightly linked to cellulose and to integrate new xyloglucans into the cell walls. Beside XTH, endo-1,3(4)-beta-glucanases in glycoside hydrolase family also have the potential to function in cell wall loosening [35]. Contigs encoding CESA, XTH and endo-1,3-beta-glucanase were also up-regulated by Cd in HE shoots (Table 1). Furthermore, genes encoding laccase, cinnamate 4-hydroxylase, and cellulose synthase were not only more induced but also expressed at higher basal level in HE than that in NHE (Figure 5). Although genes encoding endo-1,3(4)-beta-glucanase and tyrosine aminotransferase were both up-regulated by Cd with similar fold change in HE and NHE, they were constitutively expressed at higher levels in HE than that in NHE (Figure 5). These results indicate that Cd tolerance in HE shoots is probably associated with cell wall formation/modification. In a previous physiological study, we found that the highest amount of Cd was associated with cell wall fraction in HE shoots [36].

Vacuolar sequestration in leaves is a common mechanism for heavy metal compartmentalization in the hyperacculators A. halleri and T. caerulescens [6], [11]. Cation-diffusion facilitator genes (also known as MTPs) or P-type ATPase genes (e.g. AtHMA3) are important for vacuolar sequestration [6], however none were found to be up-regulated by Cd in HE shoots in the current study. On the contrary, metal tolerance protein MTP3 was down-regulated in HE but up-regulated in NHE. Two NRAMP (nature resistance associated with microphage) family genes (NRAMP2 and NRAMP3) functioning in vacuolar efflux of heavy metals [6] were significantly up-regulated by Cd in both HE and NHE. Although basal expression levels of NRAMP2 and NRAMP3 were not significantly different (P<0.05) between HE and NHE, NRAMP2 was more highly induced in HE while NRAMP3 was more induced in NHE. Another NRAMP gene NRAMP4 was also identified in our experiments, of which basal expression level was much higher in HE than that in NHE. Nevertheless, NRAMP3 expression level in HE was similar to the one in NHE while NRAMP2 and NRAMP4 expression levels were much higher in HE than that in NHE under Cd treated condition in our experiments. The P-type metal ATPase, HMA4, is involved in cytosolic metal efflux and xylem loading/unloading [9]. In the current study, one heavy metal transporter HMA4 was more up-regulated by Cd in NHE than that in HE. However, HMA4 was more expressed in HE than that in NHE (10.2 fold higher in Cont and 2.7 fold higher in Cd, respectively). We also examined the expression of one ZIP family gene metal transporter ZIP1. There was no significant difference in terms of up-regulation ratio between HE and NHE, but ZIP1 expression level was again much higher in HE than that in NHE (98.8 fold higher in Cont and 97.7 fold higher in Cd, respectively). Root-to-shoot translocation of Cd in S. alfredii Hance (HE) was found to be enhanced in our previous physiological study [16]. It would be interesting to know in the future whether the high expression level of HMA4 and ZIP1 in HE could contribute to such active translocation process.

In summary, we have generated a large collection of annotated transcript contigs from the multi-heavy-metal accumulator S. alfredii Hance (HE) shoots, which should provide more opportunities for studying heavy metal hyperaccumlation. Our shoot transcriptome analysis has revealed that several genes related to cell wall modification, metal translocation and remobilization are highly induced in S. alfredii Hance (HE). Subsequent qRT-PCR results have demonstrated that there are significant differences in the gene expression pattern between HE and NHE shoots under both normal growth condition and Cd-treated conditions. The sequencing data presented here will also aid in elucidating the functional roles of the constitutively expressed genes in Cd hyper-tolerance in S. alfredii Hance (HE).

Materials and Methods

Plant Growth and Treatment

The hyperaccumulating ecotype S. alfredii Hance (HE) was obtained from an old Pb/Zn mine area in Zhejiang Province in China and the non-hyperaccumulating ecotype S. alfredii Hance (NHE) was obtained from a tea plantation of Hangzhou in Zhejiang Province. The plants described here did not involve endangered or protected species. No specific permissions were required for collection of samples in these locations. Plants were vegetatively propagated to ensure homogeneity and cultured hydroponically in basal nutrient solution [14] supplied with or without 100 µM CdCl2 for HE and 5 µM CdCl2 for NHE for 8 days. Samples (three plants per sample) were separated into shoots and roots for RNA sequencing or qRT-PCR or elemental analysis. There were three biological replicates in each experiment. Cd concentrations were measured routinely by Inductively Coupled Plasma Mass Spectroscopy (ICP-MS) (Agilent, USA). All the data was statistically analyzed using student’s t-test or ANOVA.

Sequence Generation, Mapping and Assembly

For deep sequencing, total RNA was extracted from HE shoots with a RNAout kit (TANDZ, China) and quality checked with an Agilent 2100 Bioanalyzer (Agilent, USA). mRNA was purified with Micropoly(A)PuristTM mRNA purification kit (Ambion, USA) following the DNase I digestion (Ambion, USA). First-strand cDNA was synthesized with superscript III reverse transcriptase (Invitrogen, USA) and GsuI-oligo dT. mRNA was biotinylated after being oxidized at its 5′ end with NaIO4 (Sigma, USA) and removed with Dynal M280 magnetic streptavidin beads (Invitrogen, USA) after alkaline digestion. Adaptors were added to the 5′-end of first-strand cDNA with DNA ligase (TaKaRa, Japan) and second-strand cDNA was synthesized with Ex Taq polymerase (TaKaRa, Japan). polyA tails and 5′ end adaptors were removed with GsuI digestion. cDNA was purified with Ampure beads (Agencourt, USA) after sonication (Fisher, USA). 454 and Illumina/Solexa libraries were constructed with GS DNA Library Preparation kit (Roche Applied Science, USA) and TruSeqTM DNA Sample Prep Kit – Set A (Illumine, USA), respectively. Deep sequencing was performed with Roche’s 454 Genome Sequencer FLX and Illumina/Solexa Genome Analyzer II platforms in the Chinese National Human Genome Center at Shanghai. Sequences from 454 sequencing were assembled with the Newbler Assembler [37]. Illumina/Solexa reads were mapped to the 454-derived contigs with bowtie [29], and the remaining unmapped reads were separately assembled using Velvet/Oases [22]. All the contigs were finally assembled in CAP3 [23]. Default settings were used in the above-mentioned programs. To minimize the alternative splicing effects on assembly, the longest contig/isotig was selected for each isogroup, or the best match was selected when potential paralogs were presented. Microsatellites (SSRs) were obtained in SciRoKo (v3.4) with misa (perfect) and mmvp (mismatch allowed) modes [38]. Enrichment of KEGG pathways for a given gene list was calculated using a classical hypergeometric distribution statistical comparison of a query gene list against a reference gene list. For homologous protein comparison, sequence similarity between S. alfredii Hance (HE) contigs and other sequenced plants were obtained by BlastP using different E-value cutoffs [39].

Gene Expression Analysis, Quantitative RT-PCR and Phylogenic Analysis

Low quality reads were detected and removed with the FASTQ Quality Trimmer (−t 5, −l 50) in FASTX-Toolkit after reads containing undetermined bases (N) were discarded, the clean Illumina/Solexa reads in each sample were mapped to the assembled contigs and normalized to RPKM [29]. When reads were mapped to multiple contigs, only one match was randomly selected. Gene expression levels was determined with the MARS (MA-plot-based method with Random Sampling) model with DEGseq package [40]. The FDR was controlled using q-value (q = 0.005) to identify significant differences between treatments [40]. For qRT-PCR analysis, total RNA from three biological replicates was isolated (Qiagen, Germany) from shoots of HE and NHE, and reverse transcribed using the Supertranscript III RT kit (Invitrogen, USA) according to the manufacturer’s instructions. For primer design (Table S4), blast searches were conducted against the Arabidopsis nucleotide database with S. alfredii Hance (HE) contig sequences, and primers were selected from sequences in the sequence-conserved regions. Melting curves were monitored to ensure good amplification in both ecotypes. qRT-PCR was performed with the Mastercycler® ep realplex2 instrument (Eppendorf, Germany) with SsoFastTMEvaGreen® Supermix kit (Bio-Rad, USA). Expression levels were calculated relative to actin using a comparative threshold cycle method (CT method) [41]. C3-, C4- and CAM-type PEPC protein sequences were aligned together with the S. alfredii PEPC contigs in ClustalX and maximum parsimony tree was constructed in MEGA (version 5.05) with standard settings. The accession number of plant PEPCs is listed in Table S5.

Data Deposition

The RNA-seq results reported in this paper have been deposited into EMBL-Bank under the accession number: E-MTAB-1011; ArrayExpress under the accession number: E-MTAB-934. The assembled contigs were assigned with the accession number from HE717106 to HE774267.

Supporting Information

Figure S1.

Sequence size distribution. (A) Length distribution of sequences from Roche’s 454 sequencing. (B) Length distribution of assembled contigs from Illumina/Solexa sequencing.

https://doi.org/10.1371/journal.pone.0064643.s001

(PDF)

Figure S2.

Pathway analysis of S. alfredii Hance (HE) contigs.

https://doi.org/10.1371/journal.pone.0064643.s002

(PDF)

Figure S3.

Phylogenetic analysis of PEPC genes. Phylogenetic hypothesis was derived from maximum parsimony analysis of the C-terminal fragments of PEPCs from S. alfredii (Sa_) and Arabidopsis thaliana (At_), Zea mays (Zm_), Sorghum vulgare (Sv_), Saccharum officinarum (So_), Mesembryanthemum crystallinum (Mc_), Kalanchoë blossfeldiana (Kb_), Clusia venosa (Cv_), Clusia rosea (Cr_), Clusia schomburgkiana (Cs_), Clusia hilariana (Ch_) and Clusia aripoensis (Ca_). The tree was rooted with Saccharum officinarum. Accession numbers are listed in Table S5.

https://doi.org/10.1371/journal.pone.0064643.s003

(PDF)

Figure S4.

Partial protein sequence alignment of CAM-type PEPCs. Accession numbers are listed in Table S5. Identical amino acids are marked with stars (*) under the lines, colons and dots indicate different amino acids with strong and weak similarity, respectively.

https://doi.org/10.1371/journal.pone.0064643.s004

(PDF)

Figure S5.

Validation of RNA-seq results by qRT-PCR. The gene expression level of each contig in the hyperaccumulation ecotype (HE) was quantified and normalized to the actin control. The fold change of expression is the gene expression level in Cd treatment normalized to that in CK. Annotation of each contig is listed in Table 1. Bars depict SE (n = 3).

https://doi.org/10.1371/journal.pone.0064643.s005

(PDF)

Table S1.

Overview of the Illumina/Solexa sequencing.

https://doi.org/10.1371/journal.pone.0064643.s006

(DOC)

Table S2.

Contigs differentially regulated by Cd treatment.

https://doi.org/10.1371/journal.pone.0064643.s007

(DOC)

Table S3.

List of PEPCs in S. alfredii Hance.

https://doi.org/10.1371/journal.pone.0064643.s008

(DOC)

Table S4.

List of primers used for qRT-PCR.

https://doi.org/10.1371/journal.pone.0064643.s009

(DOC)

Table S5.

Accession numbers of plant PEPCs for sequence alignment.

https://doi.org/10.1371/journal.pone.0064643.s010

(DOC)

Dataset S1.

Simple sequence repeats found in the study.

https://doi.org/10.1371/journal.pone.0064643.s011

(XLS)

Acknowledgments

We would like to thank Dr. Patrick Schnable in Iowa State University for useful discussions prior to the start of this project and Dr. Stephen H. Howell in Iowa State University for proofreading the paper.

Author Contributions

Conceived and designed the experiments: XY JG JXL. Performed the experiments: JG LS. Analyzed the data: JG LS JXL. Wrote the paper: XY JXL.

References

  1. 1. Nawrot T, Plusquin M, Hogervorst J, Roels HA, Celis H, et al. (2006) Environmental exposure to cadmium and risk of cancer: a prospective population-based study. Lancet Oncol 7: 119–126.
  2. 2. Satarug S, Garrett SH, Sens MA, Sens DA (2010) Cadmium, environmental exposure, and health outcomes. Environ Health Perspect 118: 182–190.
  3. 3. DalCorso G, Farinati S, Furini A (2010) Regulatory networks of cadmium stress in plants. Plant signal behav 5: 663–667.
  4. 4. Pilon-Smits E (2005) Phytoremediation. Annu Rev Plant Biol 56: 15–39.
  5. 5. Pilon-Smits E, Pilon M (2002) Phytoremediation of metals using transgenic plants. Crit Rev Plant Sci 21: 439–456.
  6. 6. Verbruggen N, Hermans C, Schat H (2009) Molecular mechanisms of metal hyperaccumulation in plants. New Phytol 181: 759–776.
  7. 7. Roosens NHCJ, Willems G, Saumitou-Laprade P (2008) Using Arabidopsis to explore zinc tolerance and hyperaccumulation. Trends Plant Sci 13: 208–215.
  8. 8. Talke IN, Hanikenne M, Kraemer U (2006) Zinc-dependent global transcriptional control, transcriptional deregulation, and higher gene copy number for genes in metal homeostasis of the hyperaccumulator Arabidopsis halleri. Plant Physiol 142: 148–167.
  9. 9. Hanikenne M, Talke IN, Haydon MJ, Lanz C, Nolte A, et al. (2008) Evolution of metal hyperaccumulation required cis-regulatory changes and triplication of HMA4. Nature 453: 391–344.
  10. 10. Verbruggen N, Hermans C, Schat H (2009) Mechanisms to cope with arsenic or cadmium excess in plants. Curr Opin Plant Biol 12: 364–372.
  11. 11. Ueno D, Milner MJ, Yamaji N, Yokosho K, Koyama E, et al. (2011) Elevated expression of TcHMA3 plays a key role in the extreme Cd tolerance in a Cd-hyperaccumulating ecotype of Thlaspi caerulescens. Plant J 66: 852–862.
  12. 12. Yang XE, Long XX, Ni WZ, Fu CX (2002) Sedum alfredii H: A new Zn hyperaccumulating plant first found in China. Chin Sci Bull 47: 1634–1637.
  13. 13. Yang XE, Long XX, Ye HB, He ZL, Calvert DV, et al. (2004) Cadmium tolerance and hyperaccumulation in a new Zn-hyperaccumulating plant species (Sedum alfredii Hance). Plant Soil 259: 181–189.
  14. 14. Tian SK, Lu LL, Yang XE, Labavitch JM, Huang YY, et al. (2009) Stem and leaf sequestration of zinc at the cellular level in the hyperaccumulator Sedum alfredii. New Phytol 182: 116–126.
  15. 15. Tian SK, Lu LL, Yang XE, Webb SM, Du Y, et al. (2010) Spatial imaging and speciation of lead in the accumulator plant Sedum alfredii by Microscopically Focused Synchrotron X-ray investigation. Environ Sci Technol 44: 5920–5926.
  16. 16. Lu LL, Tian SK, Yang XE, Wang XC, Brown P, et al. (2008) Enhanced root-to-shoot translocation of cadmium in the hyperaccumulating ecotype of Sedum alfredii. J Exp Bot 59: 3203–3213.
  17. 17. Li TQ, X.E Y, Yang JY, He ZL (2006) Zn accumulation and subcellular distribution in the Zn hyperaccumulator Sedum alfredii Hance. Pedosphere 16: 616–623.
  18. 18. Yang XE, Li TQ, Yang JC, He ZL, Lu LL, et al. (2006) Zinc compartmentation in root, transport into xylem, and absorption into leaf cells in the hyperaccumulating species of Sedum alfredii Hance. Planta 224: 185–195.
  19. 19. Tian SK, Lu LL, Labavitch J, Yang XE, He ZL, et al. (2011) Cellular sequestration of cadmium in the hyperaccumulator plant species Sedum alfredii. Plant Physiol 157: 1914–1925.
  20. 20. Wang Z, Gerstein M, Snyder M (2009) RNA-Seq: a revolutionary tool for transcriptomics. Nat Rev Genet 10: 57–63.
  21. 21. Ozsolak F, Milos PM (2011) RNA sequencing: advances, challenges and opportunities. Nat Rev Genet 12: 87–98.
  22. 22. Zerbino DR, Birney E (2008) Velvet: Algorithms for de novo short read assembly using de Bruijn graphs. Genome Res 18: 821–829.
  23. 23. Huang XQ, Madan A (1999) CAP3: A DNA sequence assembly program. Genome Res 9: 868–877.
  24. 24. Rice P, Longden I, Bleasby A (2000) EMBOSS: The European molecular biology open software suite. Trends Genet 16: 276–277.
  25. 25. Chen ZZ, Xue CH, Zhu S, Zhou FF, Ling XFB, et al. (2005) GoPipe: Streamlined Gene Ontology annotation for batch anonymous sequences with statistics. Prog Biochem Biophy 32: 187–190.
  26. 26. Gupta PK, Balyan IS, Sharma PC, Ramesh B (1996) Microsatellites in plants: A new class of molecular markers. Curr Sci 70: 45–54.
  27. 27. Borland AM, Zambrano VAB, Ceusters J, Shorrock K (2011) The photosynthetic plasticity of crassulacean acid metabolism: an evolutionary innovation for sustainable productivity in a changing world. New Phytol 191: 619–633.
  28. 28. Mallona I, Egea-Cortines M, Weiss J (2011) Conserved and divergent rhythms of Crassulacean Acid Metabolism-related and core clock gene expression in the Cactus Opuntia ficus-indica. Plant Physiol 156: 1978–1989.
  29. 29. Mortazavi A, Williams BA, McCue K, Schaeffer L, Wold B (2008) Mapping and quantifying mammalian transcriptomes by RNA-Seq. Nat Methods 5: 621–628.
  30. 30. Xiong YH YX, Ye ZQ, He ZL (2004) Characteristics of cadmium uptake and accumulation by two contrasting ecotypes of Sedum alfredii Hance. J Environ Sci Health A Tox Hazard Subst Environ Eng 39: 2925–2940.
  31. 31. Vogt T (2010) Phenylpropanoid biosynthesis. Mol Plant 3: 2–20.
  32. 32. Davin LB, Jourdes M, Patten AM, Kim KW, Vassao DG, et al. (2008) Dissection of lignin macromolecular configuration and assembly: Comparison to related biochemical processes in allyl/propenyl phenol and lignan biosynthesis. Nat Prod Rep 25: 1015–1090.
  33. 33. Kaneda M, Schuetz M, Lin BSP, Chanis C, Hamberger B, et al. (2011) ABC transporters coordinately expressed during lignification of Arabidopsis stems include a set of ABCBs associated with auxin transport. J Exp Bot 62: 2063–2077.
  34. 34. Cai X, Davis EJ, Ballif J, Liang M, Bushman E, et al. (2006) Mutant identification and characterization of the laccase gene family in Arabidopsis. J Exp Bot 57: 2563–2569.
  35. 35. Cosgrove DJ (2005) Growth of the plant cell wall. Nat Rev Mol Cell Biol 6: 850–861.
  36. 36. Ni TH, Wei YZ (2003) Subcellular distribution of cadmium in mining ecotype Sedum alfredii. Acta Bot Sin 45: 925–928.
  37. 37. Margulies M, Egholm M, Altman WE, Attiya S, Bader JS, et al. (2005) Genome sequencing in microfabricated high-density picolitre reactors. Nature 437: 376–380.
  38. 38. Kofler R, Schloetterer C, Lelley T (2007) SciRoKo: a new tool for whole genome microsatellite search and investigation. Bioinformatics 23: 1683–1685.
  39. 39. Xie F, Sun G, Stiller JW, Zhang B (2011) Genome-Wide Functional Analysis of the Cotton Transcriptome by Creating an Integrated EST Database. Plos One 6.
  40. 40. Wang L, Feng Z, Wang X, Wang X, Zhang X (2010) DEGseq: an R package for identifying differentially expressed genes from RNA-seq data. Bioinformatics 26: 136–138.
  41. 41. Liu JX, Srivastava R, Che P, Howell SH (2007) An endoplasmic reticulum stress response in Arabidopsis is mediated by proteolytic processing and nuclear relocation of a membrane-associated transcription factor, bZIP28. Plant Cell 19: 4111–4119.