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

Ectopic Expression of Arabidopsis Glycosyltransferase UGT85A5 Enhances Salt Stress Tolerance in Tobacco

  • Yan-Guo Sun ,

    *sunyanguo1987@126.com (YGS); bkhou@sdu.edu.cn (BKH)

    Affiliation The Key Lab of Plant Cell Engineering and Germplasm Innovation, Education Ministry of China; School of Life Science, Shandong University, Jinan, Shandong, P. R. China

  • Bo Wang,

    Affiliation The Key Lab of Plant Cell Engineering and Germplasm Innovation, Education Ministry of China; School of Life Science, Shandong University, Jinan, Shandong, P. R. China

  • Shang-Hui Jin,

    Affiliation The Key Lab of Plant Cell Engineering and Germplasm Innovation, Education Ministry of China; School of Life Science, Shandong University, Jinan, Shandong, P. R. China

  • Xiao-Xia Qu,

    Affiliation The Key Lab of Plant Cell Engineering and Germplasm Innovation, Education Ministry of China; School of Life Science, Shandong University, Jinan, Shandong, P. R. China

  • Yan-Jie Li,

    Affiliation The Key Lab of Plant Cell Engineering and Germplasm Innovation, Education Ministry of China; School of Life Science, Shandong University, Jinan, Shandong, P. R. China

  • Bing-Kai Hou

    *sunyanguo1987@126.com (YGS); bkhou@sdu.edu.cn (BKH)

    Affiliation The Key Lab of Plant Cell Engineering and Germplasm Innovation, Education Ministry of China; School of Life Science, Shandong University, Jinan, Shandong, P. R. China

Abstract

Abiotic stresses greatly influence plant growth and productivity. While glycosyltransferases are widely distributed in plant kingdom, their biological roles in response to abiotic stresses are largely unknown. In this study, a novel Arabidopsis glycosyltransferase gene UGT85A5 was identified as significantly induced by salt stress. Ectopic expression of UGT85A5 in tobacco enhanced the salt stress tolerance in the transgenic plants. There were higher seed germination rates, better plant growth and less chlorophyll loss in transgenic lines compared to wild type plants under salt stress. This enhanced tolerance of salt stress was correlated with increased accumulations of proline and soluble sugars, but with decreases in malondialdehyde accumulation and Na+/K+ ratio in UGT85A5-expressing tobacco. Furthermore, during salt stress, expression of several carbohydrate metabolism-related genes including those for sucrose synthase, sucrose-phosphate synthase, hexose transporter and a group2 LEA protein were obviously upregulated in UGT85A5-expressing transgenic plants compared with wild type controls. Thus, these findings suggest a specific protective role of this glycosyltransferase against salt stress and provide a genetic engineering strategy to improve salt tolerance of crops.

Introduction

Salinity is a major abiotic stress that impacts on plant growth and productivity. Almost all aspects of plant activities are directly or indirectly affected by salt stress, which is generally defined as the presence of excessive amounts of soluble salts that hinders or affects the normal functions needed for plant growth [1]. Ionic and osmotic stresses are considered to be two main effects caused by salt stress on plants [2]. Previous studies indicated that salt stress-stimulated physiological imbalance increases the level of reactive oxygen species (ROS) in plant cells, leading to oxidative stress [3][5]. To cope with these environmental challenges, plants have evolved elaborate mechanisms to trigger stress responses at both molecular and morphological levels. Up to now, salt stress responses have been revealed to be mediated by many pathways. For instance, the SOS pathway genes SOS1, SOS2 and SOS3 play significant roles in resisting salt stress by regulating ionic homeostasis and Na+ tolerance [6][8]. A similar effect was also observed for the AtNHX1 gene, which encodes a vacuolar Na+/H+ antiporter and contributes to the transportation of Na+ [9]. Mitogen-activated-protein kinase (MAPK) cascades are proposed to function in salt stress tolerance by regulating osmolyte biosynthesis [10], [11]. It is documented that overexpression of a tobacco ethylene receptor NTHK1 in Arabidopsis resulted in altered salinity sensitivity, indicating that ethylene signaling is also required in salt stress response [12]. In addition, a number of other genes such as late-embryogenesis-abundant (LEA) proteins, RD29A [13] and CBF/DREB transcription factor ect, which are strongly induced by salt stress, also act as critical regulators in mediating plant salt stress adaptation [14], [15].

Glycosylation occurs in a wide range of biological processes in plants. Glycosyltransferases (GTs) are enzymes responsible for glycosylation of plant compounds, which catalyze the transfer of sugar moieties from activated donor molecules to specific acceptor molecules, for example sugars, lipids, proteins, nucleic acids, antibiotics and other small molecules, with either retention or inversion of configuration at the anomeric center [16]. Until now, GTs have been classified into 94 distinct families based on their gene products or by sequence comparison with carbohydrate active enzymes (CAZy) (http://www.cazy.org/) [17], of which Family 1 is the largest and is closely related to plant functions. GTs of Family 1 characteristically contain a carboxy-terminal consensus sequence termed ‘plant secondary product glycosyltransferase box’ (PSPG box) and use uridine 5′-diphospho sugars as the sugar donors, and are thus named UDP-sugar glycosyltransferases (UGTs) [18][20].

Comprehensive analysis of the Arabidopsis genome revealed multigene GT families consisting of 119 putative UGT genes [21]. In plants, UGTs are generally localized in the cytosol, and are involved in the biosynthesis of plant natural products such as flavonoids, phenylpropanoids, terpenoids and steroids, and in the regulation of plant hormones [22]. It is now recognized that UGT action, by adding a sugar moiety to the acceptors, can change chemical properties and bioactivity of plant molecules and enable their access to membrane transporter systems [18], [19]. Therefore, UGTs might play important roles in maintaining cell homeostasis, regulating plant growth and improving their defense responses to stressful environments [18], [19]. For instance, several UGTs have been shown to glycosylate phytohormones in vitro and participate in phytohormone homeostasis in vivo [23][27]. Several publications indicate that knockout of ugt73b3 and ugt73b5, or overexpression of UGT74F2 and UGT76B1, significantly altered plant resistance to pathogen infection, implying novel roles of UGTs in plant defense responses [28][30]. In addition, UGTs have also been identified to participate in regulating plant adaptation to abiotic stresses. A recent study showed that overexpression of UGT74E2, encoding an indole-3-butyric acid (IBA) GT, greatly affected IBA homeostasis and the expression of various downstream genes, including transcription factors, glycosidases or genes involved in hormone signaling (e.g. DIN2, P450, RD20, ORG1 and PCC1) leading to increased tolerance to salinity and drought stress in Arabidopsis thaliana [31]. UGT73B2 loss-of-function endowed the mutant with oxidative stress tolerance [32]. Family 1 are the largest GT family that has various potential substrates from plant secondary metabolism. The widespread occurrence, diversity and complexity of the glycosides throughout the plant kingdom indicate that Family 1 GTs might have broad functionality. Although some UGTs have been chemically characterized, very little is known about their actual biological functions and molecular mechanisms, particularly during plant growth, development and response to external environments.

In this study, a novel Arabidopsis GT gene UGT85A5, a group G member of Family 1, was identified as involved in regulating plant salt stress tolerance. The expression of UGT85A5 was strongly induced by NaCl. Ectopic expression of UGT85A5 in tobacco resulted in enhanced salt tolerance in transgenic compared to wild type (WT) plants. The data presented here demonstrate, for the first time, that UGT85A5 plays a significant role in enhancing plant salt stress tolerance.

Results

Induction of the Arabidopsis UGT85A5 in response to salt stress

Two-week-old Arabidopsis seedlings were treated with 150 mM NaCl for different time courses to determine the expression pattern of UGT85A5 in response to salt stress. RT-PCR analysis revealed that, compared with untreated controls, the transcript level of UGT85A5 was remarkably upregulated under NaCl treatment (Figure 1). It was induced at 3 h, peaked at 6 h and remained at a very high level at 24 h. These results indicated that UGT85A5 expression was salt inducible and that the gene was a direct or indirect target of a pathway that controlled salt-responsive gene expression.

thumbnail
Figure 1. Expression patterns of UGT85A5 in response to salt stress.

Two-week-old Arabidopsis seedlings were treated with 150 mM NaCl as salt treatment and water as control. RT-PCR analysis was performed with UGT85A5 specific primers using the RNA isolated from the Arabidopsis plants. The Tubulin was used as an internal control.

https://doi.org/10.1371/journal.pone.0059924.g001

Generation of UGT85A5 transgenic tobacco plants

To investigate the physiological function of UGT85A5 in response to salt stress, transgenic tobacco that ectopically expressed UGT85A5 was generated. A total of seven independent T0 transgenic lines were obtained by kanamycin selection and PCR identification. The expected UGT85A5 fragment was amplified in the genomic DNA of transgenic lines but not in the non-transformed plants (Figure 2A). To assess the expression level of the UGT85A5 gene in transgenic tobacco plants, 4-week-old homozygous transgenic seedlings of T2 generation were subjected to reverse transcription (RT)-PCR analysis. Of these, two lines (T1 and T3) showed the highest expression of UGT85A5 and were chosen for further analysis (Figure 2B).

thumbnail
Figure 2. Identification of UGT85A5 transgenic tobacco plants.

(A) PCR analyses of T0 transgenic tobacco lines. (B) RT-PCR analyses of UGT85A5 expression in homologous transgenic lines of T2 generation. The expression level of the Actin gene was used as an internal control. WT: wild type plants. T1 and T3: homologous transgenic tobacco lines of T2 generation.

https://doi.org/10.1371/journal.pone.0059924.g002

Salt tolerance analyses of UGT85A5 transgenic plants

To investigate whether constitutive expression of UGT85A5 in tobacco improved salt stress tolerance, both WT and transgenic plants were challenged with NaCl treatment during both seed germination and vegetative stages.

Seeds of WT and transgenic lines (T1 and T3) were surface sterilized and germinated on Murashige and Skoog (MS) medium supplemented with 0, 100, 150 and 200 mM NaCl, respectively. Overexpression of UGT85A5 in tobacco significantly promoted germination in untreated conditions (Figure 3A), suggesting that a regulatory network involved in seed germination influenced by UGT85A5 action was conserved in tobacco. In the presence of 100 mM NaCl, although there was no difference in the final germination rate of all seeds, those of T1 and T3 germinated faster than WT: >90% of T1 and T3 seeds germinated after 10 d, but only 70% for WT (Figure 3B). Additionally, under harsher salt stress conditions, the transgenic lines showed remarkably higher germination rate. About 80% transgenic seeds under 150 mM NaCl and 35% under 200 mM NaCl germinated after 14 d; however, only 40 and 10% of WT seeds germinated, respectively (Figure 3C and 3D).

thumbnail
Figure 3. Seed germination under salt stress.

Seeds of wild-type (WT) and UGT85A5 transgenic lines (T1 and T3) were germinated on MS medium supplemented with different concentrations of NaCl. (A) Control; (B) 100 mM NaCl; (C) 150 mM NaCl and (D) 200 mM NaCl. The values are means of three independent replicates (each with 150 seeds).

https://doi.org/10.1371/journal.pone.0059924.g003

We next examined whether overexpressing UGT85A5 in tobacco influenced the growth of seedlings under salt treatment. Seeds of WT and transgenic plants were germinated and grown on MS medium for 3 weeks, and then the seedlings were transferred onto MS medium containing 0, 100 and 200 mM NaCl and grown for another 4 weeks. Both T1 and T3 transgenic plants grew better compared with WT (Figure 4A), and total fresh weight (mg plant–1) of the transgenic lines under salt stress conditions was significantly (P<0.05) higher than that of WT plants (Figure 4B).

thumbnail
Figure 4. Effects of UGT85A5 overexpression on salt stress tolerance.

(A) Phenotype of WT and transgenic lines growing on MS medium supplemented with different concentrations of NaCl. (B) Fresh weight of WT and transgenic seedlings. (C) Phenotype of WT and transgenic plants watered with NaCl solution (from 100 mM to 300 mM) for 4 weeks.

https://doi.org/10.1371/journal.pone.0059924.g004

Salt stress tolerance was also investigated for large plants growing in a greenhouse. Six-week-old WT and transgenic plants grown in the same pot were watered with NaCl solution (from 100 to 300 mM as described in Methods section) for 4 weeks. The WT seedlings showed severe chlorosis, stunted phenotypes and ultimately died, whereas the transgenic seedlings continued to grow (Figure 4C) – demonstrating that ectopic expression of UGT85A5 greatly enhanced the salt stress tolerance of these transgenic plants.

Leaf disk assay of UGT85A5 transgenic tobacco plants

A leaf disk assay is a reliable index of the damage to the photosynthesis apparatus under stress. Leaf disks of WT and transgenic tobacco plants were subjected to water solutions containing 0, 100, 200 and 300 mM NaCl respectively to evaluate their tolerance to salt stress. After treatment for 3 d, a yellow phenotype was observed in WT leaf disks, whereas transgenic disks remained green (Figure 5A). Additionally, further chlorophyll measurement showed more dramatic chlorophyll loss in the WT leaf disks than the transgenic lines under salt stress. There was no difference in chlorophyll contents between WT and transgenic plants without treatment (Figure 5B). However, in the presence of 100 mM NaCl, there was 20% chlorophyll loss in WT leaf disks, but only 10% in transgenic ones – the discrepancy was greater with increased NaCl concentrations (Figure 5B). These results suggested that expression of UGT85A5 in tobacco greatly enhanced its photosynthetic abilities under salt stress.

thumbnail
Figure 5. Evaluation of salt stress tolerance of the UGT85A5 transgenic lines by leaf disc assay.

(A) Representative picture that shows phenotype differences of leaf discs between WT and transgenic lines. (B) Chlorophyll contents of leaf discs from WT and transgenic plants treated with various concentrations of NaCl.

https://doi.org/10.1371/journal.pone.0059924.g005

Evaluation of soluble osmolyte content, lipid peroxidation and Na+/K+ ratio in UGT85A5 transgenic plants

Different chemical compounds, such as proline, soluble sugars, sugar alcohols and quaternary ammonium compounds, are known to accumulate in response to abiotic stresses [33], [34]. Therefore, to gain further insight into the physiological relevance of UGT85A5, we analyzed changes of these biochemical markers in transgenic and WT plants under salt stress conditions.

Under normal conditions, a similar proline level was detected between WT and transgenic lines. Under salt stress, however, there were 17- and 16-fold higher proline levels in T1 and T3 transgenic plants, respectively. There was only a 12-fold increase of proline level in WT plants (Figure 6A).

thumbnail
Figure 6. Changes in soluble osmolyte content, lipid peroxidation and Na+/K+ ratio of the WT and transgenic lines under salt stress.

Leaves from plants treated with salt stress were collected and subjected to the determination. (A) Proline content; (B) Soluble sugars content; (C) MDA content; (D) Na+/K+ ratio.

https://doi.org/10.1371/journal.pone.0059924.g006

A similar result was obtained in content of soluble sugars. There were equal soluble sugar levels in untreated WT and transgenic lines. After treatment with 300 mM NaCl, in contrast to a 20% increase in WT plants, the soluble sugar contents significantly increased by 40 and 38% in T1 and T3 plants, respectively (Figure 6B).

To measure lipid peroxidation, malondialdehyde (MDA) content was assessed in 300 mM NaCl stressed WT and transgenic plants – about 40% more MDA accumulated in WT than in transgenic lines (Figure 6C).

Moreover, the Na+/K+ ratio was also examined to see whether it was altered in UGT85A5 transgenic plants. There were similar Na+/K+ ratios in both the transgenic and WT plants under normal conditions (Figure 6D). However, this ratio increased 18-fold in WT plants, in contrast to 12-fold in transgenic plants under salt stress conditions.

The significant level changes in the abovementioned components in the transgenic lines provided further evidence that ectopically expressing UGT85A5 in tobacco greatly protected the plants from damage caused by salt stress, which in turn enhanced salt stress tolerance in the transgenic tobacco.

Expression analysis of stress-related genes in UGT85A5 transgenic plants

To investigate how UGT85A5 increased salt stress tolerance, the expression of some stress-related genes were analyzed. When treated with high salinity, the steady-state mRNA levels of carbohydrate metabolism-related genes encoding sucrose synthase (SS) [35], sucrose-phosphate synthase (SPS) [36], hexose transporter (HT) [37] and a group2 LEA protein (ERD10C) [38] in transgenic lines were upregulated 2- to 4-fold of that in WT plants (Figure 7). However, the transcript levels of these genes were similar between transgenic lines and WT plants under normal conditions. In summary, these results showed that increased salt resistance of UGT85A5 overexpressing plants correlated with elevated expression levels of a subset of genes with known roles in stress responses.

thumbnail
Figure 7. Expression analysis of stress-related genes in UGT85A5 transgenic plants by qRT-PCR.

The transcript levels of (A) SS, (B) SPS, (C) HT and (D) ERD10C in WT and transgenic tobacco plants were analyzed under both normal and salt treatment conditions.

https://doi.org/10.1371/journal.pone.0059924.g007

Discussion

Salt stress tolerance in plants is a complex trait involving multiple physiological and biochemical mechanisms. To endow the plants with greater flexibility and plasticity under stress conditions, it is essential to make dramatic modifications to a variety of small molecules, either intrinsic or extrinsic, including hormones, metabolites and xenobiotics [39], [40]. In particular, glycosylation catalyzed by a superfamily of GTs plays significant roles in modulating the solubility, stability, bioavailability and bioactivity of various small molecules [18], 19,41. Recent studies have demonstrated that GTs play important roles in modulating plant tolerance to abiotic stresses [31], [32]. In this study, for the first time, we demonstrated that ectopic expression of the Arabidopsis GT gene UGT85A5 enhanced the adaptation of the transgenic tobacco plants under salt stress. A BLAST search at NCBI (http://www.ncbi.nlm.nih.gov/Blast) showed that UGT85A5 has two homologs in tobacco: NTGT5a (GenBank AB176523.1) and NTGT5b (GenBank AB176524.1). They share 62% sequence identity to UGT85A5 on a protein level. However, there is no report concerning the physiological role of NTGT5a and NTGT5b. UGT85A5 is a gene of Arabidopsis, and since its expression is salt inducible and its ectopic expression can increase tobacco tolerance to salt stress, we speculate that UGT85A5 may be involved in a regulation pathway of Arabidopsis in response to environmental stress. Our research on the effects of UGT85A5 in Arabidopsis using its overexpresssors and mutants is now in progress. Tobacco is not only a crop plant but also material for easy genetic transformation – thus, we chose tobacco as the initial transgenic plants for the present study. Our data suggest that UGT85A5 has promising utility in improving salinity tolerance in crops, at least in tobacco.

In the context of photosynthesis, chlorophyll is an essential factor as a driving force for plant growth and biomass production. The ability to maintain photosynthesis rate under environmental stresses is fundamental for the maintenance of plant growth and development [42]. The leaf disk assay in this study showed less chlorophyll loss in transgenic than in WT plants under salt stress conditions, indicating higher photosynthetic capacity in UGT85A5 transgenic than in WT plants.

Organic osmolytes such as proline and soluble sugars work in a variety of ways, such as in protection of cellular structures, detoxification of enzymes and scavenging of ROS alone or in combination with other defense-related enzyme systems [43], [44]. These compounds also confer integrity to the membranes and keep the photosynthetic system functioning [44], which is in agreement with our data showing higher chlorophyll contents in transgenic than in WT plants. The plasmalemma is known to play critical roles in perceiving and transferring signals; however, its fluidity can easily be affected by abiotic stresses. Compatible solutes are essential in protecting the membranes from various damages [44] and can help to overcome the osmotic stress of salinity. Therefore, better growth of transgenic lines under salt stress might be partly attributed to increased levels of osmolytes. Additionally, UGT85A5 transgenic plants exhibited less lipid peroxidation, as revealed by low MDA levels of these plants, demonstrating that the transgenic plants suffered less oxidative damage than WT plants under salt stress conditions. Ion measurements showed lower Na+/K+ ratios in leaves of UGT85A5 transgenic compared to WT plants under salt stress, demonstrating that the transgenic plants suffered lower ion toxicity, which is a strong indicator of salt stress tolerance. UGT85A5 may influence ion accumulation or transportation by catalyzing the glycosylation of some molecules, leading to regulation of some ion transporters.

In the present study, the higher expression level of SS, SPS and HT in transgenic plants was consistent with the elevated osmolyte concentration under salt conditions, resulting in improved ability to adjust in response to salt stress. The protein encoded by ERD10C is thought to protect macromolecules and membranes under stress conditions [38], [45], which is consistent with our finding of higher expression of ERD10C in transgenic than WT plants under salt stress. All these results suggested that ectopic expression of UGT85A5 positively altered the expression of these stress-related genes, leading to enhanced tolerance to salt stress in transgenic lines. UGT85A5 has previously been identified as being induced by a combination of drought and heat stress [46]. We analyzed the UGT85A5 promoter and found that it has many binding sites of transcription factors responding to environment signals, such as ABRE binding site motif (ABF binding site), G-box promoter motif and the ACGT elements (Figure S1), which have been reported to be involved in responses to dehydration and ABA treatment [47][49]. Interestingly, the UGT85A5 promoter has a short sequence (CACGTGGC) which contains three motifs in the same site, i.e. ABRE binding site motif, G-box promoter motif and ACGT element (Figure S1). The characteristics of the UGT85A5 promoter suggest that UGT85A5 may be a target of regulatory networks that control salt, drought, cold, light and ABA responsiveness.

In conclusion, our results demonstrated that UGT85A5 was responsive to salt stress, and ectopic expression of UGT85A5 strongly enhanced salt tolerance in transgenic plants. UGT85A5 encodes a putative GT belonging to the group G of Family 1 UGTs [41]. However, almost nothing is really known about its enzyme activity and its substrates. UGT85A1, encoding a cytokinin GT, is the only gene of group G whose substrate has been identified so far [50]. However, UGT85A5 did not display any activity toward cytokinins in our research (unpublished data). Further investigations concerning UGT85A5 activity and its substrates might be helpful in understanding the stress-tolerance mechanism with which it is involved.

Materials and Methods

Plant materials and growth conditions

WT Arabidopsis thaliana Columbia ecotype (Col-0) and tobacco (Nicotiana tabacum Linn, Wisconsin 38) were used in this study. Seedlings of Arabidopsis plants and tobacco were grown on MS medium in a culture room at 23±2°C under a 14/10 h light/dark photoperiod, and a light intensity of 60 µmol m–2 s–1. Plants were grown in flowerpots in soil–perlite mixtures in a greenhouse at 24±1°C under a 16/8 h light/dark photoperiod and a light intensity of 100 µmol m–2 s–1.

Vector construction and generation of transgenic plants

The full-length cDNA of the UGT85A5 gene (AT1G22370) was amplified using the following primers (UGT85A5-F: 5′-GCGTCTAGAATGGCGTCTATGCTGTTC-3′ and UGT85A5-R: 5′-GCGGAGCTCCTACTCCCCTAAAAGAACCT-3′) and cloned into the pBI121 vector driven by the cauliflower mosaic virus (CaMV) 35 S promoter. The recombinant plasmid was transferred into Agrobacterium (strain LBA4404) for tobacco transformation using the leaf disk method [51], and the transformants were screened by kanamycin (100 mg l–1). The genomic DNA of WT and T0 tobacco plants was used to amplify the target gene UGT85A5 by PCR using the primers mentioned above (i.e. UGT85A5-F and UGT85A5-R). For the RT-PCR analysis, total RNA of WT and transgenic tobacco plants (homozygous T2 generation) was extracted and reverse transcribed into cDNA. The resulting cDNA was used for the amplification of the target gene using gene-specific primers (KY-LP: 5′-GGCTTCTTGGCTTATCTACA-3′ and KY-RP: 5′-ACTCCTCAACCTCCTCCCTC-3′). The Actin gene was amplified using the following primers (ACT-F: 5′-ATGCCCTCCCACATGCTATTC-3′ and ACT-R: 5′-AACATGGTAGAGCCACCACTG-3′) as an internal reference.

Salt stress assay in Arabidopsis and transgenic tobacco lines

Two-week-old Arabidopsis seedlings grown on MS solid medium were used to determine the expression pattern of UGT85A5 under salt stress. Parts of the seedlings were treated with 150 mM NaCl, while other seedlings were treated with water as controls. Plants were sampled at 0, 1, 3, 6, 12 and 24 h after each treatment. Total RNA was extracted for RT-PCR analysis using the gene-specific primers (KY-LP and KY-RP). The Tubulin gene was amplified using the following primers (TUB-F: 5′-CTCAAGAGGTTCTCAGCAGTA-3′ and TUB-R: 5′-TCACCTTCTTCATCCGCAGTT-3′) as an internal reference.

To determine the effects of salt stress on seed germination, seeds of WT and transgenic tobacco plants (homozygous T2 generation used for all experiments in this study) were surface sterilized and sown on MS solid medium supplemented with different concentrations of NaCl (0, 100, 150 and 200 mM). The plates were placed under greenhouse conditions for 14 d. The number of germinated seeds was counted daily after sowing.

For evaluation of salt tolerance at seedling stage, 3-week-old seedlings of the WT and transgenic tobacco lines grown on MS medium were transferred to MS medium supplemented with various concentrations of NaCl (0, 100 and 200 mM) and were allowed to grow for 4 weeks. The seedling growth was observed and the fresh weight determined (mg·plant–1).

To evaluate salinity tolerance under greenhouse conditions, 1-month-old seedlings of WT and transgenic plants grown on MS medium were transferred to pots containing a mixture of garden soil and sand (3 1) and grown in a greenhouse for two more weeks before exposure to salt stress treatments. Then each line was irrigated with NaCl solution every 3 d. The NaCl concentration used initially was 100 mM and then stepped up in 50-mM increments until the final concentration of 300 mM was achieved. The phenotype of plants was photographed after 30 d of salt treatment.

For the leaf disk floating assay, leaf disks with a diameter of 1 cm (five disks per treatment) were prepared from leaves of identical development stage of both WT and transgenic plants and floated on solutions of different NaCl concentrations (0, 100, 200 and 300 mM) for 72 h. The phenotype of leaf disks was photographed and the total chlorophyll content in each sample was measured after salt treatment. The chlorophyll content [Chls a+b] was calculated after extraction in aqueous 80% acetone using the following formulae (in g·l–1): [Chls a+b] = (20.21 × A645)+(8.02 × A663) [52]. A663 and A645 represent absorbance values read at 663 and 645 nm wavelengths, respectively.

Determination of proline, soluble sugar, MDA and ion content

Three-week-old seedlings of WT and transgenic lines grown on MS medium were transferred to soil for additional growth duration of 3 weeks. Selected WT and transgenic plants of the same age and size were either transferred to pots to impose salt stress (300 mM NaCl solution) on them or to pots without salt treatment as experimental controls. After 7 d of treatment, leaves were collected for physiological assays. Proline accumulation in tobacco plants following salt treatment was measured according to the method of Shan et al. [53]. Soluble sugars were essentially extracted using the procedure of Li et al. [54]. MDA was assayed for indirect evaluation of lipid peroxidation using thiobarbituric acid, as described by Hodges et al. [55].

To determine the ion contents, 3-week-old seedlings of the WT and transgenic lines were transferred to MS medium containing different concentration of NaCl (0 and 200 mM) and were allowed to grow for 1 month. Then leaves from WT and transgenic plants were collected and dried at 80°C for 48 h. The dried materials were extracted with 1 N HNO3 as described by Storey [56]. The Na+ and K+ contents were measured in the filtrate using atomic absorption spectrophotometry.

Total RNA extraction and qRT-PCR analysis

To study the expression level of stress-related genes, 4-week-old seedlings were treated with 300 mM NaCl for 6 h as the salt treatment, while seedlings treated with water were the controls. Total RNA was extracted using Trizol reagent (TaKaRa) and treated with RNase-free DNase I (TaKaRa). Reverse transcription reactions were performed for 15 min at 37°C using the PrimeScript RT reagent kit (TaKaRa). The resulting cDNA was used for qRT-PCR analysis. The qRT-PCR was performed using the SYBR Premix Ex Taq II kit (TaKaRa) according to the manufacturer's instructions and run on LightCycler480 (Roche) real-time PCR systems. Thermocycling conditions were 95°C for 30 s followed by 40 cycles of 95°C for 5 s, 60°C for 20 s, with a melting curve detected at 95°C for 15 s and 65°C for 15 s. The relative transcript level was normalized with qActin gene. All primers used in qRT-PCR were as follows: (1) qACT-F: 5′-GCAGGCATTCACGAAACGACTTA-3′ and qACT-R: 5′-AGCCAAAATAGAACCTCCAATCCAA-3′; (2) SS-F: 5′-TGGTTGGAGGAGACCGAAGG-3′ and SS-R: 5′-GAGTTCGCCATTCCTCACGC-3′; (3) SPS-F: 5′-GAATTCAGGCGCTTCGTTGTCA-3′ and SPS-R: 5′-ACCCCTAGTTTCTCCAGTGA-3′; (4) HT-F: 5′-TTGCCCGAGACGAAGAATGT-3′ and HT-R: 5′-GCGCAGAAGGATCAAATCCA-3′; and (5) ERD10C-F: 5′-GGAAGAAGAGAAGGCGGGTGA-3′ and ERD10C-R: 5′-GGTCTTTGAGTGATATCCTGGTA-3′.

Statistical analysis

Experiments were repeated at least three times. P-values were determined by Student's t-test (the quantification of stress-related genes expression) or by one-way ANOVA using the protected least-significant difference (LSD) tests (the quantification of phenotypic differences). Data are expressed as mean values ± SE.

Supporting Information

Figure S1.

The structure of the promoter region of the UGT85A5 gene. Promoter nucleotide numbers begin at the 5′ end of the UGT85A5 mRNA (see gene annotation in TAIR). Sequences homologous to ABRE binding site motif (C/TACGTGGC), G-box promoter motif (CACGTG), GATA promoter motif (A/TGATAG/A), ACGT element (ACGT) are underlined. Note that some GATA promoter motifs are presented on the complementary strand.

https://doi.org/10.1371/journal.pone.0059924.s001

(TIF)

Author Contributions

Conceived and designed the experiments: BKH. Performed the experiments: YGS BW SHJ. Analyzed the data: YGS XXQ YJL. Contributed reagents/materials/analysis tools: BW SHJ XXQ. Wrote the paper: BKH YJL YGS.

References

  1. 1. Szabolcs I (1994) Prospect for soil salinity for the 21st century. Paper presented at the 15th World Congress of Soil Science.pp. 123–141.Acapulco,Mexico,July 10–16.
  2. 2. Munns R, Termaat A (1986) Whole-plant responses to salinity. Aust J Plant Physiol 13: 143–160.
  3. 3. Apel K, Hirt H (2004) Reactive oxygen species: Metabolism, oxidative stress, and signal transduction. Annu Rev Plant Biol 55: 373–399.
  4. 4. Mittler R, Vanderauwere S, Gollery M, van Breusegem F (2004) Reactive oxygen gene network of plants. Trends Plant Sci 9: 490–498.
  5. 5. Papadakis AK, Roubelakis-Angelakis KA (2005) Polyamines inhibit NADPH oxidase-mediated superoxides generation and putrescine prevents programmed cell death syndrome induced by the polyamine oxidase generated hydrogen peroxide. Planta 220: 826–837.
  6. 6. Wu SJ, Lei D, Zhu JK (1996) SOS1, a genetic locus essential for salt tolerance and potassium acquisition. Plant Cell 8: 617–627.
  7. 7. Liu J, Zhu JK (1997) An Arabidopsis mutant that requires increased calcium for potassium nutrition and salt tolerance. Proc Nat Acad Sci USA 94: 14960–14964.
  8. 8. Zhu JK, Liu J, Xiong L (1998) Genetic analysis of salt tolerance in Arabidopsis thaliana: evidence of a critical role for potassium nutrition. Plant Cell 10: 1181–1192.
  9. 9. Apse MP, Aharon GS, Snedden WA, Blumwald E (1999) Salt tolerance conferred by overexpression of a vacuolar Na+/H+ antiport in Arabidopsis. Science 285: 1256–1258.
  10. 10. Zhu JK (2002) Salt and drought stress signal transduction in plants. Annu Rev Plant Biol 53: 247–273.
  11. 11. Rodriguez MC, Petersen M, Mundy J (2010) Mitogen-activated plants. Annu Rev Plant Biol 61: 621–649.
  12. 12. CaoWH, Liu J, He XJ, Mu RL, Zhou HL, et al. (2007) Modulation of ethylene responses affects plant salt-stress responses. Plant Physiol 143: 707–719.
  13. 13. Shinozaki K, Yamaguchi-Shinozaki K (2000) Molecular response to dehydration and low temperature: Differences and cross-talk between two stress signaling pathways. Curr Opin Plant Biol 3: 217–223.
  14. 14. Liu Q, Sakuma Y, Abe H, Kasuga M, Miura S, et al. (1998) Two transcription factors, DREB1 and DREB2, with an EREBP/AP2 DNA binding domain, separate two cellular signal transduction pathways in drought- and low temperature-responsive gene expression, respectively, in Arabidopsis. Plant Cell 10: 1391–1406.
  15. 15. Thomashow MF (1999) Plant cold acclimation: Freezing tolerance genes and regulatory mechanisms. Annu Rev Plant Physiol Plant Mol Biol 50: 571–599.
  16. 16. Lairson LL, Henrissat B, Davies GJ, Withers SG (2008) Glycosyltransferases: structures, functions, and mechanisms. Annu Rev Biochem 77: 521–555.
  17. 17. Vaistij FE, Lim EK, Edwards R, Bowles DJ (2009) Glycosylation of Secondary Metabolites and Xenobiotics. In Plant-derived Natural Products. Osbourn AEand Lanzotti VEdited by . pp. 209–228.Springer US,New York.
  18. 18. Jones P, Vogt T (2001) Glycosyltransferases in secondary plant metabolism: tranquilizers and stimulant controllers. Planta 213: 164–174.
  19. 19. Bowles D, Isayenkova J, Lim EK, Poppenberger B (2005) Glycosyltransferases: managers of small molecules. Curr Opin Plant Biol 8: 254–263.
  20. 20. Wang J, Hou BK (2009) Glycosyltransferases: key players involved in the modification of plant secondary metabolites. Front Biol China 4: 39–46.
  21. 21. Li Y, Baldauf S, Lim EK, Bowles DJ (2001) Phylogenetic analysis of the UDP-glycosyltransferase multigene family of Arabidopsis thaliana. J Biol Chem 276: 4338–4343.
  22. 22. Bowles D, Lim EK, Poppenberger B, Vaistij FE (2006) Glycosyltransferases of lipophilic small molecules. Annu Rev Plant Biol 57: 567–597.
  23. 23. Jackson RG, Lim EK, Li Y, Kowalczyk M, Sandberg G, et al. (2001) Identification and biochemical characterization of an Arabidopsis indole-3-acetic acid glucosyltransferase. J Biol Chem 276: 4350–4356.
  24. 24. Jackson RG, Kowalczyk M, Li Y, Higgins G, Ross J, et al. (2002) Over-expression of an Arabidopsis gene encoding a glucosyltransferase of indole-3-acetic acid: phenotypic characterization of transgenic lines. Plant J 32: 573–583.
  25. 25. Lim EK, Doucet CJ, Hou B, Jackson RG, Abrams SR, et al. (2005) Resolution of (+)-abscisic acid using an Arabidopsis glycosyltransferase. Tetrahedron Asymmetry 16: 143–147.
  26. 26. Poppenberger B, Fujioka S, Soeno K, George GL, Vaistij FE, et al. (2005) The UGT73C5 of Arabidopsis thaliana glucosylates brassinosteroids. Proc Nat Acad Sci USA 102: 15253–15258.
  27. 27. Wang J, Ma XM, Kojima M, Sakakibara H, Hou BK (2011) N-Glucosyltransferase UGT76C2 is involved in cytokinin homeostasis and cytokinin response in Arabidopsis thaliana. Plant Cell Physiol 52: 2200–2213.
  28. 28. Langlois-Meurinne M, Gachon CMM, Saindrenan P (2005) Pathogen-responsive expression of glycosyltransferase genes UGT73B3 and UGT73B5 is necessary for resistance to Pseudomonas syringae pv tomato in Arabidopsis. Plant Physiol 139: 1890–1901.
  29. 29. Song JT, Koo YJ, Seo HS, Kim MC, Choi YD, et al. (2008) Overexpression of AtSGT1, an Arabidopsis salicylic acid glucosyltransferase, leads to increased susceptibility to Pseudomonas syringae. Phytochem 69: 1128–1134.
  30. 30. von Saint Paul V, Zhang W, Kanawati B, Geist B, Faus-Keßler T, et al. (2011) The Arabidopsis glucosyltransferase UGT76B1 conjugates isoleucic acid and modulates plant defense and senescence. Plant Cell 23: 4124–4145.
  31. 31. Tognetti VB, Aken OA, Morreel K, Vandenbroucke K, Cotte B, et al. (2010) Perturbation of indole-3-butyric acid homeostasis by the UDP-Glucosyltransferase UGT74E2 modulates Arabidopsis architecture and water stress tolerance. Plant Cell 22: 2660–2679.
  32. 32. Kim IA, Heo JO, Chang KS, Lee SA, Lee MH, et al. (2010) Overexpression and inactivation of UGT73B2 modulate tolerance to oxidative stress in Arabidopsis. J Plant Biol 53: 233–239.
  33. 33. Parvanova D, Ivanov S, Konstantinova T, Karanov E, Atanassov A, et al. (2004) Transgenic tobacco plants accumulating osmolytes show reduced oxidative damage under freezing stress. Plant Physiol Biochem 42: 57–63.
  34. 34. Bartels D, Sunkar R (2005) Drought and salt tolerance in plants. Crit Rev Plant Sci 24: 23–58.
  35. 35. Dejardin A, Sokolov LN, Kleczkowski LA (1999) Sugar/osmoticum levels modulate differential abscisic acid-independent expression of two stress-responsive sucrose synthase genes in Arabidopsis. Biochem J 344: 503–509.
  36. 36. Bornke F (2005) The variable C-terminus of 14-3-β proteins mediates isoform-specific interaction with sucrose-phosphate synthase in the yeast two-hybrid system. J Plant Physiol 162: 161–168.
  37. 37. Warren AP, Pasternak CA (1989) Common pathway for the induction of hexose transport by insulin and stress. J Cell Physiol 138: 323–328.
  38. 38. Kasuga M, Miura S, Shinozaki K, Yamaguchi-Shinozaki K (2004) A combination of the Arabidopsis DREB1A gene and stress-inducible rd29A promoter improved drought- and low-temperature stress tolerance in tobacco by gene transfer. Plant Cell Physiol 45: 346–350.
  39. 39. Dixon RA, Paiva NL (1995) Stress-induced phenylpropanoid metabolism. Plant Cell 7: 1085–1097.
  40. 40. Hemm MR, Rider SD, Ogas J, Murry DJ, Chapple C (2004) Light induces phenylpropanoid metabolism in Arabidopsis roots. Plant J 38: 765–778.
  41. 41. Ross J, Li Y, Lim EK, Bowles DJ (2001) Higher plant glycosyltransferases. Genome Biol 2: 3004.1–3004.6.
  42. 42. Lawlor DW (1995) The effects of water deficit in photosynthesis. In Environment and Plant Metabolism: Flexibility and Acclimation.Edited by Smirnoff N.pp.129–160.BIOS Scientific Publishers, Oxford.
  43. 43. Verma DPS (1999) Osmotic stress tolerance in plants: role of praline and sulfur metabolisms. In Molecular Responses to Cold, Drought, Heat and Salt Stress in Higher Plants. Edited by Shinozaki K.& Yamaguchi-Shinozaki K. pp. 153–168.R.G. Landes Company,Austin,TX.
  44. 44. Xiong L, Schumaker KS, Zhu JK (2002) Cell signaling during cold, drought, and salt stress. Plant Cell 14: S165–S183.
  45. 45. Grover A, Kapoor A, Lakshmi OS, Agarwal S, Sahi CH, et al. (2001) Understanding molecular alphabets of the plant abiotic stress responses. Curr Science 80: 206–221.
  46. 46. Rizhsky L, Liang HJ, Shuman J, Shulaev V, Davletova S, et al. (2004) When Defense Pathways Collide. The Response of Arabidopsis to a Combination of Drought and Heat Stress. Plant Physiol 134: 1683–1696.
  47. 47. Nakashima K, Fujita Y, Katsura K, Maruyama K, Narusaka Y, et al. (2006) Transcriptional regulation of ABI3- and ABA-responsive genes including RD29B and RD29A in seeds, germinating embryos, and seedlings of Arabidopsis. Plant Mol Biol 60: 51–68.
  48. 48. Nakashima K, Ito Y, Yamaguchi-Shinozaki K (2009) Transcriptional regulatory networks in response to abiotic stresses in Arabidopsis and grasses. Plant Physiol 149: 88–95.
  49. 49. Simpson SD, Nakashima K, Narusaka Y, Seki M, Shinozak K, et al. (2003) Two different novel cis-acting elements of erd1, a clpA homologous Arabidopsis gene function in induction by dehydration stress and dark-induced senescence. Plant J 33: 259–270.
  50. 50. Hou B, Lim EK, Higgins GS, Bowles DJ (2004) N-glucosylation of cytokinins by glycosyltransferases of Arabidopsis thaliana. J Biol Chem 279: 47822–47832.
  51. 51. Horsch RB, Fry JE, Hoffmann NL, Eichholtz D, Rogers SG, et al. (1985) A simple and general-method for transferring genes into plants. Science 227: 1229–1231.
  52. 52. Arnon DI (1949) Copper enzymes in isolated chloroplasts. Plant Physiol 24: 1–15.
  53. 53. Shan DP, Huang JG, Yang YT, Guo YH, Wu CA, et al. (2007) Cotton GhDREB1 increases plant tolerance to low temperature and is negatively regulated by gibberellic acid. New Phytol 176: 70–81.
  54. 54. Li W, Maoyin L, Zhang W, Welti R, Wang X (2004) The plasma membrane-bound phospholipase Dδ enhances freezing tolerance in Arabidopsis thaliana. Nat Biotechnol 22: 427–433.
  55. 55. Hodges DM, DeLong JM, Forney CF, Prange RK (1999) Improving the thiobarbituric acid-reactive-substances assay for estimating lipid peroxidation in plant tissues containing anthocyanin and other interfering compounds. Planta 207: 604–611.
  56. 56. Storey R (1995) Salt tolerance, ion relations and the effects of root medium on the response of Citrus to salinity. Aust J Plant Physiol 22: 101–114.