Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Mol Cell Proteomics. 2016 Jan; 15(1): 266–288.
Published online 2015 Sep 25. doi: 10.1074/mcp.M115.051961
PMCID: PMC4762511
PMID: 26407991

Mechanisms of Soybean Roots' Tolerances to Salinity Revealed by Proteomic and Phosphoproteomic Comparisons Between Two Cultivars* An external file that holds a picture, illustration, etc.
Object name is sbox.jpg

Erxu Pi,¶¶§§ Liqun Qu,¶¶ Jianwen Hu,§¶¶ Yingying Huang,¶¶ Lijuan Qiu, Hongfei Lu, Bo Jiang,** Cong Liu, Tingting Peng, Ying Zhao, Huizhong Wang, Sau-Na Tsai,‡‡ Saiming Ngai,‡‡§§ and Liqun Du§§

Associated Data

Supplementary Materials

Abstract

Understanding molecular mechanisms underlying plant salinity tolerance provides valuable knowledgebase for effective crop improvement through genetic engineering. Current proteomic technologies, which support reliable and high-throughput analyses, have been broadly used for exploring sophisticated molecular networks in plants. In the current study, we compared phosphoproteomic and proteomic changes in roots of different soybean seedlings of a salt-tolerant cultivar (Wenfeng07) and a salt-sensitive cultivar (Union85140) induced by salt stress. The root samples of Wenfeng07 and Union85140 at three-trifoliate stage were collected at 0 h, 0.5 h, 1 h, 4 h, 12 h, 24 h, and 48 h after been treated with 150 mm NaCl. LC-MS/MS based phosphoproteomic analysis of these samples identified a total of 2692 phosphoproteins and 5509 phosphorylation sites. Of these, 2344 phosphoproteins containing 3744 phosphorylation sites were quantitatively analyzed. Our results showed that 1163 phosphorylation sites were differentially phosphorylated in the two compared cultivars. Among them, 10 MYB/MYB transcription factor like proteins were identified with fluctuating phosphorylation modifications at different time points, indicating that their crucial roles in regulating flavonol accumulation might be mediated by phosphorylated modifications. In addition, the protein expression profiles of these two cultivars were compared using LC MS/MS based shotgun proteomic analysis, and expression pattern of all the 89 differentially expressed proteins were independently confirmed by qRT-PCR. Interestingly, the enzymes involved in chalcone metabolic pathway exhibited positive correlations with salt tolerance. We confirmed the functional relevance of chalcone synthase, chalcone isomerase, and cytochrome P450 monooxygenase genes using soybean composites and Arabidopsis thaliana mutants, and found that their salt tolerance were positively regulated by chalcone synthase, but was negatively regulated by chalcone isomerase and cytochrome P450 monooxygenase. A novel salt tolerance pathway involving chalcone metabolism, mostly mediated by phosphorylated MYB transcription factors, was proposed based on our findings. (The mass spectrometry raw data are available via ProteomeXchange with identifier PXD002856).

Cultivated soybean (Glycine max (L.) Merrill) is one of the most important legume crops (1, 2), and is estimated to contributes to 30% of edible vegetable oil and 69% of protein-rich food or feed supplements worldwide (3). However, the yield of soybean is significantly reduced under environmental stresses such as salinity especially during the early vegetative growth stage (3, 4). Soil salinity is estimated to affect at least 20% of the irrigated land worldwide (5, 6) and could affect 50% of the cultivated land by year 2050 (7).

High salinity causes oxidative stress and ionic imbalance in plant cells, and further inhibits the growth and development of the whole plant (6, 8, 9). Elimination of excessive reactive oxygen species (ROS)1 via glutathione-ascorbate cycle and maintaining tolerable salt levels inside the plant cells through exportation or compartmentalization are generally accepted as two major strategies used by plants to survive salinity stress (10). Plants have evolved a series of adaptive mechanisms to sense and respond to salinity cues and these include active involvements of multiple phosphorylation cascades, such as salt overly sensitive (SOS) pathway, phosphatidic acid (PA)-mediated activation of calcium-dependent protein kinase (CDPK), abscisic acid (ABA)-regulated activation of mitogen-activated protein kinase (MAPK) cascades (1114). Phosphorylation of specific signaling components are known to be initiated at critical time points after plants been subjected to the salt stresses (15) and they coordinate specific metabolic processes, cell-wall porosity and lateral root initiation to help plants adapt to salt stresses (10, 13, 16).

Recently, major high throughput strategies including transcriptomic, proteomic, and metabolomic approaches, have been used to dissect the responses of soybean root to salinity stress (1721). However, most of these studies were focused on relatively late responses to salinity (e.g. over 48 h after Na+ treatment), earlier signal events minutes after the treatments were apparently ignored. Signaling events through protein phosphorylation are well known to play critical roles mediating appropriate physiological responses in determining the salt-tolerant capability of different soybean species. Many techniques have recently been developed for the specific enrichment of phosphopeptides; these includes immobilized metal affinity chromatography (22), strong cation-exchange chromatography (23, 24), and TiO2 affinity chromatography (25). The TiO2 affinity chromatography has been generally accepted as one of the most effective approaches in enrichment of phosphopeptides (26).

Glycine max cultivar Union85140 and Glycine soja cultivar Wenfeng07 are salinity sensitive- and tolerant-cultivar, respectively; their drastic difference in salt tolerance enable us to explore the critical proteins contributing the salt tolerance in cultivated soybeans (27, 28). In the present research, we compared the proteomes and phosphoproteomes of these two soybean species at different time points after salinity treatment. Technologies including TiO2 affinity chromatography, 2-DE MS/MS, and LC-MS/MS were used to generate the row proteome and phosphoproteome data; large-scale bioinformatic analyses including gene ontology (GO) enrichment and phosphorylation motif enrichment were conducted to identified interested targets; functional characterization of selected target genes using gain-of-function composites in soybean and loss-of-function mutant of their homologs in Arabidopsis were conducted to confirm their role in regulating plant tolerance to salt stresses. Our results reveal that normal chalcone metabolism plays a potential role in regulating plant responses to salt stresses in soybean and provide new insights into the mechanism contributing to the difference in salt tolerance of these two soybean cultivars.

EXPERIMENTAL PROCEDURES

Plant Materials and Stress Treatments

Seeds of Glycine max cultivar Union85140 (a salt sensitive species) and Glycine soja cultivar Wenfeng07 (a salt tolerant species) were kindly provided by Prof. Lijuan Qiu from the Chinese Academy of Agricultural Sciences. The seeds were surface sterilized with 5% NaClO for 5 min and rinsed three times with sterile distilled H2O. Seeds were germinated in wet filter paper at room temperature (about 22–25 °C) with 40–60% humidity. The seedlings were transferred to 1/4 fold Hoagland's solution. Seedlings at three-trifoliate stage were treated with 150 mm NaCl for 0 h, 0.5 h, 1 h, 4 h, 12 h, 24 h, and 48 h before the root samples were collected for analyses. All the samples were stored at −80 °C until use.

Protein Extraction

Total proteins from roots was extracted as described by Lv et al (29) with minor modifications. Briefly, about 4 g of root tissue for each sample was ground into fine powder in liquid nitrogen. The powder was thoroughly suspended in 45 ml of precooled TCA/Acetone (v:v = 1: 9); the homogenate was settled for overnight and then centrifuged at 14,000 × g for 15 min. The pellet was washed three times with acetone and the residual acetone was removed by vacuum. All the above experiments were carried out at 4 °C. 50 mg white powder was resuspended in 800 μl SDT lysis buffer (4% SDS, 100 mm Tris-HCl, 1 mm DTT, 1 mm PMSF, pH7.6, including one-fold PhosSTOP phosphatase inhibitor mixture from Roche), and boiled for 15 min in water bath, and followed by 100 s of sonication. After centrifugation at 14,000 × g for 15 min at 4 °C, the protein in supernatant was quantified via BCA (bicinchoninic acid) method (30).

Protein Digestion with Prior Filter Aided Sample Preparation

Approximately 1.5 mg aliquot of dissolved protein for each sample was processed by the filter aided sample preparation method to remove SDS in the samples (31). Briefly, dithiothreitol (DTT) was added to the protein solution to reach 100 mm, and then boiled for 5 min. 25 μl aliquot of each sample was mixed with 200 μl UA buffer (8 m Urea, 150 mm Tris-HCl pH 8.0), loaded into a Microcon filtration devices (Millipore, with a MWCO of 10 kd), and centrifuged at 14,000 × g for 15 min; 200 μl of fresh UA buffer was added to dilute the concentrate in the device and centrifuged again. The volume of concentrate was brought to 100 μl with UA buffer supplemented with 50 mm iodoacetamide (IAA) and the sample was shaken at 600 rpm for 1 min. After 30 min incubation at room temperature, the samples were diluted with 40 μl of digestion buffer (contains 5 μg of trypsin). The mixture was shaken at 600 rpm for 1 min, and incubated at 37 °C for 16–18 h. After digestion, the peptide solution was passed through a Microcon filtration device (MWCO 10 kd), and the concentration of the collected peptides was estimated based on their OD at 280 nm (32).

Eight-plex iTRAQ Labeling

For every eight-plex set, a pooled sample was obtained by combing two groups of samples representing seven time points (a control and six salt treatments) from two cultivars (Union 85140 and Wenfeng07). These pooled samples serve as normalizing reference for the peptide content in samples from all the tested eight-plex sets. A 200 μg digested peptides of each sample was subjected to AB Sciex iTRAQ labeling (Fig. 1). The eight-plex iTRAQ labeling was performed according to the manufacturer's instructions. A total of six eight-plex sets of iTRAQ samples were used for the three biological replicates.

An external file that holds a picture, illustration, etc.
Object name is zjw0111551750001.jpg

Sample set of quantitative phosphoproteomic analysis. For each biological replication, two eight-plex iTRAQ sets were used for the seven time points (C, T0.5, T2, T4, T12, T24 and T48). A pool sample, combined equally with all the 14 samples, was included in each eight-plex iTRAQ set for normalization between different sets. *W: Wenfeng07; U: Union 85140; T0∼T48: Plant treated with 150 mm NaCl for 0 h, 0.5 h, 1 h, 4 h, 12 h, 24 h and 48 h.

Phosphopeptide Enrichment

Phosphopeptides were enriched using TiO2 beads as described by Ostasiewicz et al. (33) with minor changes. Labeled peptide solutions were lyophilized and acidified by dissolving into DHB buffer (3% 2, 5-DiHydroxyBenzoic acid, 80% ACN and 0.1% TFA). The 25 μg of TiO2 beads (10 μ in diameter, Sangon Biotech) were added to 50 μl peptide solution and spun down after 2 h incubation at room temperature. The pellets were packed into plastic tips (fit to 10 μl pipette), washed 3 times with 20 μl of wash solution 1 (20% acetic acid, 300 mm octanesulfonic acid sodium salt and 20 mg/ml DHB) then followed by three times with 20 μl wash solution 2 (70% water; 30% ACN). The enriched phosphopeptides were eluted using freshly prepared ABC buffer (50 mm ammonium phosphate, pH 10.5) and lyophilized for MS analysis.

NanoRPLC-MS/MS Analysis of Phosphorylated Peptides

The lyophilized phosphopeptides were subjected to capillary liquid chromatography tandem mass spectrometry using a two dimensional EASY-nLC1000 system coupled to a Q Exactive Hybrid Quadrupole-Orbitrap Mass Spectrometer (Thermo Scientific). In nanoLC separation system, mobile phase A solution contains 2% acetonitrile (ACN) and 0.1% formic acid in water, and mobile phase B solution contains 84% ACN and 0.1% formic acid. The Thermo EASY SC200 trap column (RP-C18, 3 μm, 100 mm × 75 μm) was pre-equilibrated with mobile phase A before peptides loading. The phosphopeptides were initially transferred to the SC001 column (150 μm × 20 mm, RP-C18) using 0.1% formic acid solution. The peptides were then separated via the trap column using a gradient of 0–55% mobile phase B for 220 min with a flow rate of 250 nL/min followed by a 8 min rinse with 100% of mobile phase B. The trap column was re-equilibrated to the initial conditions for 12 min. The MS data of each sample were acquired for 300–1800 m/z at the resolution of 70 k. The 20 most abundant ions from each MS scan were subsequently dissociated by higher energy collisional dissociation (HCD) in alternating data-dependent mode. The HCD generated MS/MS spectra were acquired with a resolution no less than 17,500.

Phosphopeptide Identification and Quantitative Analysis

The raw HCD files were analyzed by Mascot2.2 and Proteome Discoverer1.4 and searched against a peptide database derived from the Glycine max genome sequence (“uniprot_Glycine_74305_20140429.fasta” downloaded from http://www.uniprot.org/on April 29, 2014, which includes 74, 305 nonredundant predicted peptide sequences) (34). The Mascot search parameters were list in Table I. The Proteome Discoverer 1.4 was used for integrating the spectra intensity (> 200) of the eight-plex reporter ions. The quantitative value of phosphopeptides at different treatment time points was normalized using the pooled sample as a reference and converted to log2 value of fold-change. The phosphopeptides pass the cutoff and detected in at least two replicates were used for assessment of significant change in response to NaCl stress. In this research, two statistical approaches were used for significance analysis. The “significance A” value previously described by Cox and Mann (35) was adapted to evaluate the changes between the treated (samples) and untreated (control, T0) root tissues in each biological replicate with each of which includes three technical replicates. A Student's t test was performed using the standard deviation of the pooled sample (standard) between different biological replicates for assessing the global variability of all tested samples (29). The phosphopeptides that passed both Significance A < 0.05 and p value < 0.05 were considered significantly changed (36).

Table I

Parameters of mascot search
Type of searchMS/MS Ion search
EnzymeTrypsin
Mass valuesMonoisotopic
Max missed cleavages2
Fixed modificationsCarbamidomethyl (C), Itraq8plex(N-term), iTRAQ8plex(K)
Variable modificationsOxidation (M)
Peptide mass tolerance± 20 ppm
Fragment mass tolerance0.1 Da
Protein massUnrestricted
DatabaseUniprot Glycine.fasta
Database patternDecoy
Peptide FDR≦ 0.01

Protein Shotgun Identifications by Thermo Scientific LTQ Velos

To construct a comprehensive database of salt responsive proteins in soybean, the LTQ Velos Mass Spectrometer coupled to Zorbax 300SB-C18 peptide traps (Agilent Technologies, Wilmington, DE), was used for protein identifications (37). In which, the analytical column is 0.15 mm × 150 mm (RP-C18) (Column Technology Inc., Fremont, CA). Each sample was analyzed three times and the peptides/proteins identified were combined and listed in supplemental Table S1 and S2.

2-DE Gel Based MALDI-TOF/TOF Mass Spectrometer Analysis for Protein Identification

0.2% (w/v) DTT and 0.5% IPG buffer (Lot No.: 17–6000-87, GE Healthcare Life Sciences, Piscataway, NJ) were added into the 200 μg samples before IEF. Total 250 μl samples containing about 200 μg proteins were applied to the dry IPG strips (13 cm, pH 3–10 nonlinear, GE healthcare). The program of IEF was as followed: rehydration at 20 °C for 12 h, 30 V for 8 h, 150 V for 2 h, 500 V for 0.5 h, 1000 V for 0.5 h, 4500 V for 4000 v·hrs, 8000 V for 66000 v·hrs. Focused strips were first equilibrated by incubating in equilibration buffer (pH 8.8, 2% (w:v) SDS, 6 m urea, 50 mm Tris-HCl, 30% glycerol (v:v) containing 1% DTT (w:v) for 15 min, followed by incubation in the abovementioned equilibration buffer containing 4% (w:v) iodoacetamide (IAA) for also 15 min. The second dimension separation was conducted on the 12% acrylamide SDS-PAGE. The PAGE gels were stained with Coomassie brilliant blue for over 2 h. Then, all these gels were captured by magic scanner with the same contrast and brightness. Sequentially, spots in these gel images were analyzed using ImageMasterTM 2D Platinum 5.0 software (GE Healthcare) and their relative volumes (% Vol) were represented as relative abundances. Each sample had at least two independent replicates and the differentially expressed protein spots' relative volumes were compared with Student's t test analysis (p ≤ 0.05). Spots with significant changes were excised out, and destained with 100 μl destaining solution combined with 25 mm ammonium bicarbonate and 50% (v: v) methanol in Milli-Q water. The gel crystals were dehydrated in 100% acetonitrile and vacuum-dried. Then, gel plugs were rehydrated with 10 μg/μl of trypsin in 25 mm ammonium bicarbonate on ice for 40 min and transferred into 30 °C incubator for 16–18 h digestion. Finally, 80% acetonitrile with 20% trifluoroacetic acid (v:v) was used to extracted digested peptides from the gels. MALDI-TOF/TOF mass spectrometer 4700 Proteomics Analyzer (Applied Biosystems, USA) was applied to identify mass spectrometry of digested peptide. The MS scans were acquired among the mass range from m/z 700 to 3500 Da and the mass errors were less than 50 ppm. The MS precursor ions corresponding to porcine trypsin autolysis products (m/z 805.417, m/z 906.505, m/z 1153.574, m/z 2163.057, and m/z 2273.160) were excluded. All MS and MS/MS spectra were search via the MASCOT search engine against the soybean database (source: http://www.phytozome.net/soybean). The proteins were annotated against Uniprot database. The annotations were confirmed by comparison to the annotation of the top protein hits from the online blast search against the NCBI protein database.

Quantitative RT-PCR Analysis

RNA isolation, mRNA reverse transcription and qRT-PCR methods were performed as described by Wang et al. (38) with mini modifications. The root samples were frozen with liquid N2 and total RNAs were extracted with TRIZOL Reagent (Invitrogen). The genomic DNA was removed with DNase I and cDNA was synthesized using the Plant RNeasy Mini kit (Qiagen) according to the manufacturer's instructions. The primers were generated with NCBI online Primer-BLAST against the G. max genome (39). The soybean actin11 gene was used as a reference for normalization. Quantitative RT-PCR used SYBRTM Premix Ex Taq™ (TaKaRa, Shiga, Japan) and the reaction was conducted on a CFX96 System (Bio-Rad). The gene specific primers are listed in supplemental Table S3.

Bioinformatic Analysis

Peptide motifs were extracted using the motif-X algorithm (40). The width of the generated motifs was set as seven amino acids and serine or threonine was selected as the central amino. Gene oncology (GO) analysis was carried as described by Lv et al. (29). The cis-elements recognized by transcription factor binding were identified using JASPAR software (41, 42).

Scavenging Activity of the Superoxide Anion (O2) Assay

This assay was based on the method of Zhang et al. (43) with slight modifications. Antioxidant enzymes were extracted with 10 ml of 0.05 m phosphate buffer (pH 5.5) from 0.5 g root homogenate. The extract was centrifuged at 12,000 × g (4 °C) for 10 min. 1 ml collected supernatant (crude enzyme extract) was added into 4 ml the reaction buffer, which was consist of 2 ml 0.05 m phosphate buffer, 1 ml 0.05 m guaiacol (substrate, overdose) and 1 ml 2% hydrogen peroxide (H2O2). The increased absorbance at 470 nm due to the enzyme-dependent guaiacol oxidation was recorded every 30 S until the reaction time reached 4 min. The enzyme's radical scavenging activity (RSA) was defined as: RSA = VVt×1w×ΔODΔt (g/min), where w is the weight of fresh root (g), Vt is the volume of crude enzyme used in the reaction mixture (ml), Ve is the total volume of extracted crude enzyme (ml), Δt is the cost time of the reaction (min).

Free Radical Scavenging Activity on ABTS·+

The ABTS cationic radical (ABTS·+) decolorization assay was done by the method of Re et al. (44). ABTS·+ working solution was generated by adding 2.45 mm potassium persulphate (final concentration) to 7 mm ABTS (final concentration). This working solution was incubated in dark at room temperature for 12–16 h until it gave an absorbance of 0.70 ± 0.02 at 734 nm. Ten microliters of extracts were mixed with 1.0 ml of working ABTS·+ solution and incubated at 30 °C for 30 min and the absorbance of reaction mixture was measured at 734 nm. The enzyme's radical scavenging activity was expressed as: RSA = ΔODΔt×1w× Df × M0 (mm/g/min), where ΔOD is the reduced absorbance value, Δt is the reaction time (min), Df is the dilution factor, w is the weight of fresh root (g), M0 is the original ABTS·+ concentration.

Na+ and K+ Ion Content Analyses

Na+ and K+ ion contents were detected followed the methods proposed by Qi et al. (45) using the flame atomic absorption spectrophotometer (Shimadzu AA-6300C). The content was expressed as: milligram ion per gram fresh weight (mg/g FW).

Gain-of-function Test of GmCHS, GmCHI, and GmCPM in Soybean Hairy Root System

The full-length CDSs of GmCHS (Glyma01g43880.1), GmCHI (Glyma04g40030.1), and GmCPM (Glyma07g14460.1) from Wenfeng07 was cloned into the pCAMBIA1301 vector between NcoI and BglII sites downstream of the 35S promoter. The original pCAMBIA1301 vector was used as a negative control. All these constructs were transformed into the salt-sensitive cultivar Union85140 via agrobacterium rhizogenes strain K599 as previously described (3). The composites were treated in 1/2 fold MS medium with 100 mm NaCl or without NaCl. The seedlings were weighted 10 days after salt treatment.

Loss-of-function Test of AtCHS, AtCHI and AtCPM in Arabidopsis thaliana

The seeds of deletion mutants chs, chs/chi, chs/cpm (Seed stock number: CS85, CS8584, CS8592) were got from ABRC (Arabidopsis Biological Resource Center) and germinated on 1/2 Murashige and Skoog (MS) medium. 5 days after germination, the seedlings were transferred to 1/2 MS medium with or without 150 mm NaCl. The photos of plants were taken 10 days after salt treatment.

RESULTS

Salt stress is well known to cause leaf chlorosis by reducing chlorophyll a, b, and total chlorophyll content (46). After NaCl treatment, we found that the relative contents of chlorophyll a, b, and carotenoids in Union85140 decreased more than that in Wenfeng07 (Fig. 2). In addition, the chlorophyll a/b ratio in Union85140 increased less than that in Wenfeng07 at each time points. These results confirmed that Wenfeng07 is significantly more tolerant to salt stress than Union85140 at the physiological level.

An external file that holds a picture, illustration, etc.
Object name is zjw0111551750002.jpg

Different tolerances of Wenfeng07 and Union85140 to salt stress. A, the cultivar Wenfeng07 showed significant stronger tolerance than Union85140. B–E, chlorophyll content analysis. Error bars represent standard error of three biological replicates.

ROS Elimination Capacity and Na+/K+ Content in Roots of the Two Soybean Cultivars

The antioxidant property in plant tissue is generally accepted to correlate with plant tolerance to various abiotic stresses including salinity, and it is usually represented by general radical scavenging capacities of peroxidases (POD), ascorbate peroxidase (APX), glutathione S-transferase (GST), and superoxide dismutase (SOD).

The antioxidant properties in salt-treated root tissue of the two cultivars were analyzed using H2O2 guaiacol, DPPH (2, 2-diphenyl-1-picrylhydrazyl) and ABTS (2, 2′ -azinobis (3-ethylbenzothiazoline 6-sulfonate)) radical scavenging capacity assays as previously described (47). As shown in Fig. 3A and and33B, there was no significant difference (p > 0.05) between Wenfeng07 and Union 85140 in their superoxide scavenging capacities under normal condition (T0). The scavenging capacities of the superoxide anion (SASA) in these two variants increased consistently at the early stage after salt-treatment (from 1.39 ± 0.03 g−1*min−1 in T0 to 4.87 ± 0.12 g−1*min−1 at time point T4). Starting from 4th hr (T4) of salt treatment, SASA values in these two cultivars were found to decline from their climaxes. Interestingly, SASA values in the salt-tolerant Wenfeng07 were shown to be higher than that in the salt-sensitive Union 85140 at the first four sampling times after salt treatment (from T0.5 to T12), but declined quicker and to a much lower level than that in Union 85140 24 h after the treatment. Similar to SASA, ABTS●+ scavenging potentials in the two tested cultivars displayed peak values at T1 after a short increase, then started to decrease in the rest of the stress treatment. Consistent with their salt tolerance, the ABTS●+ scavenging capability of Wenfeng07 were found to be significantly higher than that of Union 85140 (p < 0.05) all the time.

An external file that holds a picture, illustration, etc.
Object name is zjw0111551750003.jpg

Measurement of physiological indices. A and B, analysis of ROS scavenger enzymes' activities. C and D, Na+ and K+ relative content analysis (mg per gram fresh wieght). Error bars represent standard error of three biological replicates.

In addition, the Na+ content and Na+/K+ ratio were compared in the two cultivars. Our results showed that, changes in Na+ content and the Na+/K+ ratio exhibited similar dynamic patterns at different time points in these two cultivars (Fig. 3C and and33D). The salt tolerant wenfeng07 accumulated higher level of Na+ (7.0671 ± 0.5495 mg/g FW) than the salt sensitive Union85140 (1.5189 ± 0.0026 mg/g FW) under control condition. The root Na+/K+ ratio in wenfeng07 (0.1829 ± 0.0143-fold) was also significantly higher than that in Union85140 (0.0350 ± 0.0001-fold). After treatment, two peak values of the root Na+ content were observed at time points T4 and T24 (Fig. 3C).

Protein Expression Profiles Revealed by LC-MS/MS

To obtain a comprehensive observation on soybean responses to salinity and to search for clues to the mechanistic differences resulting drastic difference in their tolerance, LC-MS/MS was used to analyze root samples of the two compared species of the soybean subjected to salt stress as described in the previous section. Results of three biological replicates are included in supplemental Tables S1 and S2, and major discoveries are summarized in Table II. A total of 46410 peptides out of 14702 proteins were identified from Wenfeng07 and 46710 peptides out of 14585 proteins from Union85140 (Table II). Of these, 4464 and 4409 nonredundant proteins were found from Wenfeng07 and Union85140 respectively.

Table II

The differential expressed proteins that were identified by LC-MSMS at different time-points
ControlT0.5T1T4T12T24T48
Wenfeng07Peptides2526320940203610330633943140
Non-redundant peptides1427166019791952179018651747
Non-redundant protein1854205522172146219921152116
Union 85140Peptides2888357539392809350333113330
Non-redundant peptides1611189720481597186818421885
Non-redundant protein1880216023441902209721712031

In total, there were 89 differential expressed proteins been identified by LC-MSMS between these two cultivars (Table III). In detail, 25 and 20 proteins were specifically detected in Wenfeng07 and Union85140 roots, respectively (Table III). Among the 25 proteins specifically presented in Wenfeng07, many of them including MYB transcription factors (TFs), ethylene-responsive transcription factor 6, chalcone synthase, cytochrome P450 monooxygenase CYP51G1, glutamate receptor and a PDR-like ABC-transporter were previously reported to be related with stress responses (23, 25, 29). Although among the 20 proteins specifically detected in Union85140, the auxin pathway related proteins (such as auxin response factor, auxin-induced protein AUX22 and PIN6a), drought stress responsive protein (KS-type dehydrin SLTI629) and many kinases (such as serine/threonine protein kinase and stress-induced receptor-like kinase 2), might contribute to its general response to salinity stress. Additionally, different homologs of a protein family presented differential expressions in the two varieties. For example, for eukaryotic translation initiation factor 3, the subunit F was expressed with higher abundance in Wenfeng07 roots, whereas the M subunit was expressed with higher abundance in Union85140 roots. Similar dynamics were found in hypersensitive induced reaction protein and nodulin proteins. In addition, the ascorbate peroxidase 2, GST 8, pathogenesis-related protein and two superoxide dismutases (I1LKZ3 and I1LR93) showed opposite trends in these two varieties—they were down-regulated in Wenfeng07, but up-regulated in Union85140.

Table III

The differential expressed proteins that were identified (LC-MSMS) between two variants. Note: “-” indicates no detectable signal been found in sample collected at this (these) time point(s); the amount of “+” shows the number of detectable signal(s) been found in sample collected at this (these) time point(s)
Protein descriptionAccession No.Soybean Gene IDsWenfeng07
Union 85140
ControlStressControlStress
14-3-3 proteinQ8LJR3Glyma18g53610.1+
2-hydroxyisoflavanone dehydrataseQ5NUF3Glyma01g45020.1++++++
Alcohol-dehydrogenaseQ9ZT38Glyma04g41990.1+
ascorbate peroxidase 1Q76LA8Glyma11g15680.5+++++++++++++
ascorbate peroxidase 2Q39843Glyma12g07780.3++
Auxin response factorK7KH37Glyma03g41920.2+
Auxin-induced protein AUX22P13088Glyma08g22190.1+
Catalase-1/2P29756Glyma17g38140.1++++++++++++
Catalase-3O48560Glyma14g39810.1+++++++
Chalcone isomerase 4BQ53B71Glyma04g40030.1++
Chalcone synthaseQ6X0M9Glyma05g28610.2+
Chalcone synthase 1P24826Glyma08g11620.1+
Chalcone synthase 2P17957Glyma08g11630.2+
Chalcone synthase 3P19168Glyma08g11635.1+
Chalcone synthase 5P48406Glyma01g43880.1+
Chalcone synthase 7P30081Glyma08g11530.1+++++
Chalcone synthase 9B3F5J6Glyma08g11610.1+
Chalcone synthase CHS4Q6X0N0Glyma08g11520.1+
Chalcone-flavonone isomerase family proteinQ6X0M8Glyma01g22880.1+
Cytochrome P450 monooxygenaseQ2LAJ9Glyma07g14460.1+
Cytochrome P450 monooxygenaseQ2LAL0Glyma09g05440.1+
Ethylene-responsive transcription factor 6C6T283Glyma12g35550.1+
Eukaryotic translation initiation factor 3I1JPD4Glyma03g32950.5+++++
Eukaryotic translation initiation factor 3I1JK05Glyma03g00470.1++++
Eukaryotic translation initiation factor 3I1JQD9Glyma03g36470.1+++++++
Eukaryotic translation initiation factor 3I1M228Glyma13g31200.1++++
Eukaryotic translation initiation factor 3C6TC72Glyma18g03340.1+
Eukaryotic translation initiation factor 3I1KXJ9Glyma08g40110.1++
Eukaryotic translation initiation factor 3C6TL4Glyma12g00510.1+++
Eukaryotic translation initiation factor 3I1L3W4Glyma09g29540.1++
Eukaryotic translation initiation factor 3I1JUS7Glyma04g08570.1+++
Eukaryotic translation initiation factor 3C6TED0Glyma20g22090.1++
Ferritin Fer182I1MYZ9Glyma18g02800.2+
Glutamate receptorI1KFC6Glyma06g46130.1+
Glutathione S-transferase GST 15Q9FQE3Glyma10g33650.1+
Glutathione S-transferase GST 24Q9FQD4Glyma14g03470.1++
Glutathione S-transferase GST 8Q9FQF0Glyma07g16910.1++++++
Glyceraldehyde-3-phosphate dehydrogenaseQ38IX0Glyma04g01750.1++
Heat shock protein 90–2B6EBD6Glyma14g01530.1+++++++++
Histone H2A OSC6SV65Glyma19g42760.1+++++++
Hsp70-Hsp90 organizing protein 1Q43468Glyma17g14660.1++
Hypersensitive induced reaction protein 1G8FVT3Glyma19g02370.1++
Hypersensitive induced reaction protein 3G8FVT2Glyma05g01360.3+++
Isoflavone reductase homolog 2Q9SDZ0Glyma04g01380.1++++++++++++
KS-type dehydrin SLTI629A9XE62Glyma19g29210.1+
Late-embryogenesis abundant protein 1C6TLT7Glyma14g04180.1+++++++++++++
Leucine-rich repeat family protein/protein kinase family proteinC6ZRY3Glyma10g08010.1++
Lysine–tRNA ligaseI1L9B1Glyma10g08040.1+
MATE efflux family proteinI1K9K1Glyma06g09550.1+
Mitochondrial phosphate transporterO80412Glyma19g27380.2+++++++++++++
Mitochondrial Rho GTPaseI1LBC8Glyma10g29580.1+
Mitogen-activated protein kinase 2Q5K6N6Glyma02g15690.2++
MYB transcription factor MYB107Q0PJJ3Glyma08g20270.1+
MYB transcription factor MYB130Q0PJG6Glyma01g40220.1+
MYB transcription factor MYB91Q0PJH2Glyma07g00930.1+
NAK-type protein kinaseC6ZRR4Glyma14g39290.1+
Nodulin 35Q9ZWU0Glyma10g23790.1++
Nodulin-44P04672Glyma13g44100.1+
Pathogenesis-related proteinC6SZ24Glyma17g03340.1+++++++++++++
PDR-like ABC-transporterQ1M2R7Glyma03g32520.1++
Peroxisomal ascorbate peroxidaseB0M196Glyma12g03610.1+++++
Phosphate transporterQ8W198Glyma19g27380.2+++++++++++++
Phosphatidylinositol-specific phospholipaseQ43439Glyma02g42430.1+++
Phosphoinositide-specific phospholipase CQ43443Glyma14g06450.1+++
PIN6aM9WP18Glyma14g27900.1+
Plamsma membrane-associated AAA-Q2HZ34Glyma13g39830.1+++++++
Plasma membrane Ca2+-ATPaseQ9FVE7Glyma06g04900.1+
PR10-like proteinC6T1G1Glyma05g38110.1+++++++
PR-5 proteinB6ZHC0Glyma01g42661.1+++++
Protein kinase Pti1C6TCB9Glyma10g44212.2+++++
Protein ROOT HAIR DEFECTIVE 3I1KGC2Glyma07g01230.1++++++++++++
Pti1 kinase-like proteinC6ZRP9Glyma17g04410.2+
Pto kinase interactorC6ZRX5Glyma02g01150.1+
Putative chalcone isomerase 4Q53B72Glyma06g14820.1+++++++++++++
Putative receptor-like protein kinase 2Q49N12Glyma13g34100.1+++++++
Serine/threonine protein kinaseC6ZRR6Glyma19g40820.1++
Serine/threonine protein kinaseC6ZRT7Glyma20g38980.2+++
Serine/threonine protein kinaseC6TDV2Glyma10g01200.1+
Somatic embryogenesis receptor-like kinase-C6FF61Glyma08g19270.1+
Sterol 24-C methyltransferase 2–1D2D5G3Glyma04g02271.1+
Sterol 24-C methyltransferase 2–2D2D5G4Glyma06g02330.1+
Stress-induced protein SAM22P26987Glyma07g37240.2+++++++
Stress-induced receptor-like kinase 2B2ZNZ2Glyma15g02450.1+
Superoxide dismutaseI1JYA9Glyma04g39930.1+++++
Superoxide dismutaseI1LCI3Glyma10g33710.1++++++++
Superoxide dismutase [Cu-Zn]I1LKZ3Glyma11g19840.3+++++++++++
Superoxide dismutase [Cu-Zn]I1LR93Glyma12g08650.1+++++++++++
Superoxide dismutase [Cu-Zn]I1LTN6Glyma12g30260.1++++++++
Superoxide dismutase [Fe], chloroplasticP28759Glyma20g33880.2+++++++++

Transcriptional Expression Analysis of the Salt Responsive Genes

To explore the changes of abovementioned salt responsive proteins at the transcriptional level, 89 primer pairs of the genes encoding these proteins (supplemental Table S3) were synthesized for transcriptional-level analysis via quantitative RT-PCR. Among the 89 differentially expressed proteins, the transcriptional expression patterns of these genes in the salt treatment group were divided into three groups based on their differences between Wenfeng07 and Union85140 (Fig. 4). The first group (28 genes) had higher expression levels in Wenfeng07 than those in Union85140 at most time points, including genes encoding SOD (Glyma11g19840.2, Glyma12g08650.1, and Glyma04g39930.1), serine/threonine protein kinase (Glyma19g40820.1 and Glyma20g38980.2), MYB transcription factor MYB91 (Glyma07g00930.1), MYB transcription factor MYB107 (Glyma08g20270.1), GST 15 (Glyma10g33650.1), and cytochrome P450 monooxygenase (Glyma07g14460.1). The second group had lower expression levels in Wenfeng07 at most time points, with 39 genes encoding stress-induced receptor-like kinase (Glyma15g02450.1), sterol 24-C methyltransferase (Glyma04g02271.1 and Glyma08g19270.1), Pto kinase interactor (Glyma02g01150.1), protein kinase Pti1 (Glyma10g44212.2), PR10-like protein (Glyma05g38110.1), phosphate transporter (Glyma19g27380.2), MYB transcription factor MYB130 (Glyma01g40220.1), and hypersensitive induced reaction protein (Glyma05g01360.3). The remaining 22 genes were in the third group, among which the gene transcriptional expressions were mostly higher at T1 and T4 in Wenfeng07 but lower at other time points, such as chalcone isomerase (Glyma06g14820.1 and Glyma04g40030.1). Generally, the protein-encoding genes involved in the chalcone metabolism pathway (chalcone synthase, chalcone isomerase, and cytochrome P450 monooxygenase) showed higher expression levels in Wenfeng07. Members of the GmMYB TF family were differentially activated in the two cultivars.

An external file that holds a picture, illustration, etc.
Object name is zjw0111551750004.jpg

Clustering heat map of differentially expressed proteins. Each column represents a time point of NaCl concentration. The color codes represent the average values of three biological replicates. The abbreviations of gene names were listed in supplemental Table S3.

2-DE Mapping and Identification of Differentially Expressed Proteins

The 2-DE MS/MS strategy was applied to visualize and quantitatively analyze the defense-related proteins in the roots of two soybean varieties at times T0 and T4. The results showed that 115 protein spots (including 90 nonredundant proteins) of about 900 reproducible spots, demonstrated significant changes between T4 and T0 and were successfully identified using MALDI-TOF/TOF-MS (Table IV and andV).V). In particular, 46 differentially expressed proteins were identified from Wenfeng07 and 69 proteins from Union85140 (Fig. 5 and supplemental Fig. S1, Table IV and andVV).

Table IV

Differentially expressed proteins in Wenfeng07 under the salinity stress
Spots No.Protein No.Protein descriptionTheoretial MW/pIMatched peptideProtein scoreCI%Changes
675Glyma10g33680.1chaperonin CPN60–2, mitochondrial-like isoform 161393.4/5.7515116100Down
3689Glyma10g39780.7ubiquitin 1117168.2/6.755177100Down
1342Glyma13g41960.1fructokinase-2-like35375.4/5.2919356100Down
1304/1394Glyma04g01380.1isoflavone reductase homolog 233918.7/5.614273100Down/down
1528/1561Glyma09g02800.1Ferrodoxin NADP oxidoreductase42241.3/8.3815223100Down/down
2908/3503/2844Glyma17g03350.1stress-induced protein SAM2216746.6/4.9313325100Down/up
1320gi 62546339PIP2,230750.91/8.269230100Up
1180Glyma01g01180.1malic enzyme/malate dehydrogenase (NADP+)64985.8/5.836100100Up
2170Glyma03g05480.1disease resistance response protein 206-like22015.6/9.88692100Up
1466Glyma03g23890.1NADP-dependent alkenal double bond reductase P1-like37896.5/5.9413214100Up
1819Glyma03g26060.1stellacyanin-like19158.2/5.1436899.97Up
1485Glyma04g16350.2prohibitin-1, mitochondrial-like isoform 130373/7.9317231100Up
1616Glyma04g37120.1elongation factor 1-delta-like24972.7/4.428190100Up
1400Glyma04g40550.2nascent polypeptide-associated complex subunit alpha-like protein 2-like14753.6/5.047236100Up
1702Glyma06g39710.1proteasome subunit alpha type-627366.9/5.5816309100Up
1493Glyma06g47520.1prohibitin-1, mitochondrial-like30300.9/7.9618399100Up
1632Glyma07g33780.1caffeoyl-CoA O-methyltransferase-like28053.4/5.4614295100Up
899Glyma08g02100.2monodehydroascorbate reductase, chloroplastic-like52130.1/8.3612169100Up
1800Glyma08g17810.4proteasome subunit alpha type-2-A-like25562.2/5.5112163100Up
1538Glyma08g24950.1prohibitin-1, mitochondrial-like30462.1/7.9614148100Up
1594Glyma08g40800.1mitochondrial outer membrane protein porin of 36 kDa-like29786.4/7.0714219100Up
2915Glyma09g04530.1ABA-responsive protein ABR1716522.5/4.688156100Up
1916Glyma09g08340.1groes chaperonin, putative26640.2/6.7718310100Up
1264Glyma11g33560.1cytosolic glutamine synthetase GSbeta138966.5/5.4813274100Up
1264Glyma12g00430.1putative quinone-oxidoreductase homolog, chloroplastic-like34810.4/8.2712149100Up
1820Glyma12g31850.3protein usf-like26332.2/5.3810104100Up
2112Glyma13g32300.1flavoprotein wrbA-like21653/6.4310442100Up
285Glyma13g40130.1protein disulfide isomerase-like 1–4-like isoform 162343.4/4.7215165100Up
1920Glyma14g09440.1cysteine proteinase RD21a-like50977.4/5.3711200100Up
1207Glyma14g36850.1fructose-bisphosphate aldolase, cytoplasmic isozyme-like38330/7.1211145100Up
1610Glyma14g40670.2cysteine proteinase 15A-like40216.1/6.828216100Up
1362Glyma15g15200.1glucan endo-1,3-beta-glucosidase, basic isoform-like43758.5/8.7511447100Up
1928Glyma15g19970.120 kDa chaperonin, chloroplastic-like26653.2/7.7910108100Up
370Glyma16g00410.1stromal 70 kDa heat shock-related protein, chloroplastic-like73709.4/5.219230100Up
2264Glyma16g33710.1Kunitz trypsin protease inhibitor-like precursor23640.1/5.176172100Up
1609Glyma17g35720.1cysteine proteinase RD21a-like52082/5.5512413100Up
1649Glyma18g16260.1mitochondrial outer membrane protein porin of 36 kDa-like29814.4/7.8817254100Up
3788Glyma19g42760.1Histone H2A OS14684/10.366329100Down
1996Glyma20g38560.1chalcone flavonone isomerase23250.2/6.2316432100Up
1303/1314Glyma17g02260.1copper amino oxidase75776/6.2114147100Up/down
1379/1784Glyma03g28850.1glucan endo-1,3-beta-glucosidase precursor38088.3/8.7218510100Up/up
1380/1393Glyma05g22180.1peroxidase 73-like35475/9.0312245100Up/up
1539/1941Glyma09g37570.1peroxisomal voltage-dependent anion-selective channel protein29737.6/8.5714351100Up/Up
1757/1810Glyma12g07780.2ascorbate peroxidase 227108.8/5.6515262100Up/Down
1006/1088Glyma12g32160.1peroxidase 39-like35644.1/7.129214100Up/Up
2155/2168Glyma15g41550.1cytosolic phosphoglycerate kinase42408.6/5.9610103100Up/Up

Table V

Differentially expressed proteins in Union85140 under the salinity stress
Spots No.Protein IDProtein descriptionTheoretial MW/pIMatched peptideProtein scoreCI%Changes
796Glyma02g44080.1T-complex protein 1 subunit eta-like60234.2/6.199130100Down
988Glyma03g34830.1enolase-like47628.4/5.4919473100Down
1072Glyma03g38190.2S-adenosylmethionine synthase 1-like isoform 143196.7/5.5713205100Down
1221Glyma04g39380.2actin-7-like41688.9/5.3115214100Down
968Glyma05g24110.1elongation factor 1-alpha-like isoform 149232.7/9.1565481.467Down
910Glyma05g28490.1serine hydroxymethyltransferase 251686.1/6.911152100Down
1217Glyma05g32220.2actin-7-like41711.9/5.3714228100Down
1216Glyma06g15520.2actin-7-like37069.6/5.3889299.995Down
948Glyma07g30210.1methylmalonate-semialdehyde dehydrogenase [acylating], mitochondrial-like57578.5/6.5315139100Down
699Glyma07g33570.1ferredoxin-nitrite reductase, chloroplastic-like65836.6/6.4723271100Down
960Glyma07g36040.1ferric leghemoglobin reductase-2 precursor52968.7/6.914152100Down
1210Glyma08g03120.1biotin carboxylase precursor58770.2/7.2220194100Down
892Glyma08g11490.2serine hydroxymethyltransferase 251733.2/7.5918294100Down
1154Glyma08g17490.1probable inosine-5′-monophosphate dehydrogenase35562.5/7.6811136100Down
3400Glyma08g24760.1ripening related protein17750.8/5.9611229100Down
1486Glyma08g24950.1prohibitin-1, mitochondrial-like30462.1/7.9614145100Down
1930Glyma08g40800.1mitochondrial outer membrane protein porin of 36 kDa-like29786.4/7.0713184100Down
951Glyma10g29600.1seryl-tRNA synthetase-like51333.1/6.03117699.81Down
2302Glyma11g07540.1Transcription factor APFI-like protein29247.9/6.3610119100Down
1334Glyma11g08920.1isocitrate dehydrogenase39315.3/6.4711140100Down
1106Glyma1337s00200.1S-adenosylmethionine synthase-like43027.7/5.6517391100Down
2011Glyma13g01040.2Mitochondrial outer membrane protein porin29738.2/8.6698799.984Down
2222Glyma13g32300.2flavoprotein wrbA-like21112.7/6.0976698.052Down
574Glyma13g41370.1protein TOC75–3, chloroplastic-like87454/7.2922236100Down
1263/2324/2388Glyma13g41960.1fructokinase 235375.4/5.2917272100Down
1032Glyma14g02530.3dihydrolipoyllysine-residue succinyltransferase component of 2-oxoglutarate dehydrogenase complex 2, mitochondrial-like50131.5/9.1710128100Down
3449Glyma15g13140.1actin-depolymerizing factor 2-like10414.2/5.656215100Down
1113Glyma15g21890.2S-adenosylmethionine synthase-like isoform 143025.8/5.522472100Down
3191Glyma15g31520.1ripening related protein21494.8/6.2910148100Down
929Glyma17g04210.1dihydrolipoyl dehydrogenase, mitochondrial-like52854.5/6.913148100Down
531Glyma17g35890.1polyadenylate-binding protein 2-like71880/5.712120100Down
1694Glyma17g37050.1proteasome subunit alpha type-1-A-like isoform 130956.4/5.0712183100Down
1983Glyma18g16260.1mitochondrial outer membrane protein porin of 36 kDa-like29814.4/7.8816264100Down
973Glyma19g37520.1enolase47643.4/5.420507100Down
439Glyma20g19980.1chaperonin CPN60–2, mitochondrial-like isoform 160983.3/6.38128399.965Down
1087Glyma20g38030.126S protease regulatory subunit 6A homolog A-like47425.4/4.9823331100Down
1527/1535/1544/1877/2814Glyma09g37570.1peroxisomal voltage-dependent anion-selective channel protein29737.6/8.5711320100Down/up/down/down/up
1229Glyma02g46380.2pyruvate dehydrogenase E1 component subunit beta, mitochondrial-like38696.8/5.711127100Up
3364Glyma03g05480.1disease resistance response protein 206-like22015.6/9.887122100Up
1402Glyma03g23890.1NADP-dependent alkenal double bond reductase P1-like37896.5/5.9411249100Up
3112Glyma03g38630.1germin-like protein 122832.2/9.065180100Up
675Glyma04g01220.1phosphatidylinositol transfer-like protein III70795.2/8.44125783.799Up
2219/2293Glyma04g01380.1isoflavone reductase homolog 233918.7/5.616347100Up
2436Glyma05g22180.1peroxidase 73-like35475/9.0310229100Up
3481Glyma05g38160.1Protein yrdA, putative27715.5/8.3412157100Up
1151Glyma06g12780.3alcohol dehydrogenase 1-like36891.4/5.7716357100Up
2158Glyma07g33780.1caffeoyl-CoA O-methyltransferase-like28053.4/5.4610140100Up
3103Glyma07g37250.2Stress-induced protein SAM2215524.9/4.748228100Up
3319Glyma08g17810.4proteasome subunit alpha type-2-A-like25562.2/5.5111216100Up
788Glyma09g40690.12,3-bisphosphoglycerate-independent phosphoglycerate mutase60831/5.517182100Up
697Glyma10g41330.2ATP synthase subunit beta, mitochondrial-like58664.8/8.8318479100Up
1438Glyma11g07490.1isoflavone reductase homolog A622-like33978.9/6.1212187100Up
1160Glyma11g33560.1cytosolic glutamine synthetase GSbeta138966.5/5.4811194100Up
1978Glyma11g34380.2tropinone reductase homolog At1g0744029159.8/7.569164100Up
3362Glyma12g31850.3protein usf-like26332.2/5.3867499.731Up
3914Glyma13g32300.1flavoprotein wrbA-like21653/6.438325100Up
2478Glyma14g36850.1fructose-bisphosphate aldolase, cytoplasmic isozyme-like38330/7.1214201100Up
3217Glyma15g04290.1triosephosphate isomerase, cytosolic-like27181.1/5.8716494100Up
1874Glyma15g13550.1peroxidase C3-like isoform 138103.7/8.626108100Up
2301Glyma15g13680.1Ferredoxin–NADP reductase, root isozyme, chloroplastic42164.3/8.5212174100Up
2334Glyma15g15200.1glucan endo-1,3-beta-glucosidase, basic isoform-like43758.5/8.7513468100Up
1936Glyma15g27660.1alpha-amylase/subtilisin inhibitor-like isoform 123521.5/4.779137100Up
2171Glyma17g10880.3malate dehydrogenase, chloroplastic-like43120.4/8.1111185100Up
2109Glyma17g15690.1expansin-like B1-like27650.4/6.37229100Up
3672Glyma20g38560.1chalcone flavonone isomerase23250.2/6.2316456100Up
1218/1388Glyma12g32160.1peroxidase precursor35644.1/7.1213226100Up/up
969/970/1276Glyma09g01270.2fumarylacetoacetase-like40512.1/6.4914197100Up/Up/down
3199/3208Glyma02g40820.1isocitrate dehydrogenase (NADP) (EC 1.1.1.42)46050.5/5.8718156100Up/up/up
1979/1983/2133Glyma06g18110.7Glyceraldehyde-3-phosphate dehydrogenase36662/8.303199100Up/up/up
An external file that holds a picture, illustration, etc.
Object name is zjw0111551750005.jpg

Different expressed proteins in Wenfeng07 and Union85140 identified by 2-D MS/MS under salinity stress (at time points T0 and T4). SDS-PAGE gels were stained with Coomassie Brilliant Blue; A and B, 2-D maps of root proteome of Wenfeng07 at time points T0 and T4, respectively; C and D, 2-D maps of root proteome of Union85140 at time points T0 and T4, respectively.

The results showed that the chalcone flavonone isomerase/chalcone isomerase was also up-regulated in both Wenfeng07 and Union85140, supporting the LC-MS/MS observations. Interestingly, the ascorbate peroxidase 2 (APX2) protein showed two different isoelectric points (Spots 1757 and 1810 in Fig. 5) in Wenfeng07 roots. In addition, Spot 1757 was up-regulated whereas Spot 1810 was down-regulated. Similarly, the stress-induced protein SAM22 and copper amino oxidase proteins also had two different isoelectric points and different mass weights. These two proteins showed opposite down/up change trends after salt treatment in Wenfeng07. In Union85140, the peroxisomal voltage-dependent anion-selective channel protein, fumarylacetoacetase-like, isocitrate dehydrogenase (NADP) (EC 1.1.1.42) and glyceraldehyde-3-phosphate dehydrogenase protein showed three or more spots in the 2-DE gels. Altogether, 14 nonredundant proteins were identified from two or three DEP spots with different isoelectric points and/or molecular weights in the two soybean varieties (Table VI). This implied that the isoforms of the abovementioned proteins might play significant roles in the two varieties.

Table VI

Differential expressed proteins identified with two or more spots on 2-DE gels. *W: Wenfeng07; U: Union 85140
Uniprot accession No.Protein descriptionIDs on 2D gelTheoretial MW/pI
Q9ZNZ6peroxidase precursor1218/138835644.1/7.12
I1LU76peroxidase 39-like1006/108835644.1/7.12
I1MRA7copper amino oxidase; diamine oxidase1303/131475776/6.21
I1MJC7cytosolic phosphoglycerate kinase2155/216842408.6/5.96
C6T8Y4Ferrodoxin NADP oxidoreductase1528/156142241.3/8.38
I1M561fructokinase 21263/2324/238835375.4/5.29
I1KZY9fumarylacetoacetase-like969/97040512.1/6.49
C6TL98glucan endo-1,3-beta-glucosidase precursor1379/178438088.3/8.72
C6T857isocitrate dehydrogenase (NADP) (EC1.1.1.42)3199/320846050.5/5.87
Q9SDZ0isoflavone reductase homolog 22219/229333918.7/5.6
Q39843l-ascorbate peroxidase 21757/181027108.8/5.65
C6THQ0peroxidase 73-like1380/139335475/9.03
I1L602peroxisomal voltage-dependent anion-selective channel protein1527/1535/1544/1539/1877/1941/281429737.6/8.57
Q43453stress-induced protein SAM222908/350316746.6/4.93

Phosphopeptide Identification and Quantitative Analysis

The intensity of each phosphopeptide was normalized to the mean of intensities of all phosphopeptides within each biological replicate. Subsequently, the log2 intensity value changes (salt stress time point Tx/T0) in each condition were calculated for each phosphopeptide (supplemental Table S4). The Student's t test (p values) was performed using the standard deviation of the pooled sample (standard) between different biological replicates for assessing the global variability of all tested samples (supplemental Table S4).

In total, 5509 phosphorylated sites corresponding to 2692 phosphoproteins were identified (supplemental Tables S4 and S5), and 2344 phosphoproteins containing 3744 phosphorylation sites were quantitatively analyzed. Of these, 34.04% of phosphopeptides were detected in all three biological replicates, and 24.29% in two biological replicates (Fig. S2A). In addition, 31.41% of phosphoproteins were detected in all three biological replicates, and 24.97% in two biological replicates (supplemental Fig. S2B). Besides, there were 673 protein, which were found by LC-MSMS approaches (supplemental Tables S1 and S2), been also identified as phosphoproteins (supplemental Tables S4 and S5).

Identification of Differentially Expressed Proteins with Phosphorylation Sites

Among the 179 differentially expressed nonredundant proteins (89 nonredundant proteins from LC-MS/MS and 90 nonredundant proteins from 2-DE MS/MS), 16 proteins were also identified as phosphoproteins (Table VII, Fig. 5 and supplemental Fig. S1), such as PIP2,2 (Uniprot accession No. C6TBC3), stress-induced protein SAM22 (Uniprot accession no. Q43453), histone H2A OS (Uniprot accession no. C6SV65), eukaryotic translation initiation factor 3 subunit C (Uniprot accession no. I1JQD9) and glyceraldehyde-3-phosphate dehydrogenase cytosolic-like (Uniprot accession no. I1KC70). These phosphoproteins were involved in signal transduction, chromosome remodeling, gene translation, and energy metabolism (1014, 29).

Table VII

Differential expressed proteins identified with reliable phosphorylated sites
Protein No.Uniprot accession No.Protein descriptionPhosphorylated peptidePhosphorylated site (probabilities)
Glyma09g40690.1I1L6W02,3-bisphosphoglycerate-independent phosphoglycerate mutaseAHGTAVGLPTEDDMGNSEVGHNALGAGR/AHGTAVGLPTEDDMGNSEVGHNALGAGRT(4): 0.0; T(10): 0.0; S(17): 100.0/T(4): 0.0; T(10): 95.9; S(17): 4.1
Glyma01g45020.1Q5NUF32-hydroxyisoflavanone dehydrataseLLSSENVAASPEDPQTGVSSKS(3): 0.0; S(4): 0.0; S(10): 100.0; T(16): 0.0; S(19): 0.0; S(20): 0.0
Glyma06g12780.3C6TD82alcohol dehydrogenase 1-likeIIGVDLVSSRS(8): 100.0; S(9): 0.0
Glyma11g33560.1C6TJN5cytosolic glutamine synthetase GSbeta1WNYDGSSTGQAPGEDSEVIIYPQAIFRY(3): 0.0; S(6): 33.3; S(7): 33.3; T(8): 33.3; S(16): 0.0; Y(21): 0.0
Glyma11g33560.1C6TJN5cytosolic glutamine synthetase GSbeta1WNYDGSSTGQAPGEDSEVIIYPQAIFRY(3): 0.0; S(6): 33.3; S(7): 33.3; T(8): 33.3; S(16): 0.0; Y(21): 0.0
Glyma04g37120.1C6SXP1elongation factor 1-delta-likeAAVAEDDDDDDVDLFGEETEEEKT(19): 100.0
Glyma03g36470.1I1JQD9eukaryotic translation initiation factor 3 subunit CYFVDNASDSDDSDGQK/SDSEASQYDNEKY(1): 0.0; S(7): 100.0; S(9): 100.0; S(12): 100.0/S(1): 100.0; S(3): 0.0; S(6): 0.0; Y(8): 0.0
Glyma14g36850.1C6TMG1fructose-bisphosphate aldolase, cytoplasmic isozyme-likeLASISVENVESNR/LADGASESLHVEDYK/GILAADESTGTIGKS(3): 98.5; S(5): 1.5; S(11): 0.0/S(6): 0.1; S(8): 99.9; Y(14): 0.0/S(8): 98.3; T(9): 1.7; T(11): 0.0
Glyma06g18110.7I1KC70glyceraldehyde-3-phosphate dehydrogenase, cytosolic-likeEASYDEIKS(3): 98.9; Y(4): 1.1
Glyma19g42760.1C6SV65histone H2A OSGEIGSASQEFS(5): 0.0; S(7): 100.0
Glyma19g29210.1A9XE62KS-type dehydrin SLTI629EHGHEHGHDSSSSSDSD/EHGHEHGHDSSSSSDSD/EHGHEHGHDSSSSSDSDS(10): 32.9; S(11): 32.9; S(12): 32.9; S(13): 0.6; S(14): 0.6; S(16): 0.0/S(10): 91.0; S(11): 8.3; S(12): 4.5; S(13): 0.5; S(14): 4.7; S(16): 91.0/S(10): 0.5; S(11): 0.5; S(12): 67.0; S(13): 67.0; S(14): 67.0; S(16): 98.1
Glyma10g08010.1C6ZRY3leucine-rich repeat family protein/protein kinase family proteinEEDFSYSGIFPSTR/SSELNPFANWEQNTNSGTAPQLKS(5): 0.0; Y(6): 0.0; S(7): 0.0; S(12): 98.3; T(13): 1.7/S(1): 0.0; S(2): 0.0; T(14): 33.3; S(16): 33.3; T(18): 33.3
Glyma01g01180.3I1J4J8malic enzyme OSIWLVDSKS(6): 100.0
Glyma14g39290.1C6ZRR4NAK-type protein kinaseVQSPNALVIHPRS(3): 100.0
gi 62546339C6TBC3PIP2,2DVEQVTEQGEYSAKT(6): 0.0; Y(11): 1.1; S(12): 98.9
Glyma17g03350.1Q43453stress-induced protein SAM22SVENLEGNGGPGTIKS(1): 100.0; T(13): 0.0

Phosphorylation Motif Analysis for Quantitative Phosphopeptides

To extract overrepresented patterns from the 1164 quantitative phosphorylated peptides with differential changes between the two cultivars, the software MEME Suite and motif-X were used to analyze the motifs generated at different time points after salinity treatment from the two soybean cultivars. The intensities of phosphopeptides from Wenfeng07 (IpW) were compared with those from Union85140 (IpU) and the ratio values (IpW:IpU) with significant (p value < 0.05) differences were divided into two groups. When the intensity value IpW > IpU, its corresponding phosphopeptide was categorized into the Up group, whereas the phosphopeptide with IpW < IpU was categorized into the Down group. The Up group represented the peptides with higher phosphorylation level in the salt-tolerant cultivar and lower phosphorylation level in the salt-sensitive cultivar. There were ten phosphorylation motifs enriched from the Up group (Fig. 6A) and 14 motifs enriched from the Down group (Fig. 6B). In addition, Ser and Thr were observed as the central phosphorylated amino acid residue in both groups, with much higher frequency for Ser. In both the Up and Down groups, the amino acid closely neighboring the phosphorylated Ser/Thr was mainly Pro or Asp (Fig. 6). There were six phosphorylation motifs ([sP], [xDsDx], [xsxxD], [xsxSx], [xsxDx], and [Sxxsx]) enriched from both Up and Down groups. Four motifs ([xsxPx], [xsDxE], [xsxEx], and [Pt]) were only found in the Up group, and eight ([xPxsPx], [xDsx], [xsxDD], [xsSPx], [Dxxsx], [Axxsx], [xtPx], and [xtDx]) only in the Down group. These differentially regulated motifs were then searched for their target kinases in relevant databases, for example, [sP] is a potential substrate of plant MAPK and [sDxE] is recognized by casein kinase-II (29, 48, 49).

An external file that holds a picture, illustration, etc.
Object name is zjw0111551750006.jpg

Phosphorylation motifs enriched from sequence of peptide with different modification levels in two cultivars. A, phosphorylation motifs extracted from the phosphopeptides in the Up group by motif-X. B, phosphorylation motifs extracted from the phosphopeptides in the Down group by motif-X.

The Phosphorylated TFs and Their Specific Binding Motif in Enzymes Involved in Chalcone Metabolism

Several transcription factors, including MYB, bZIP, WRKY, ERF, BTF and GTE families were identified with fluctuating phosphorylation modifications at different time points of salt treatment (supplemental Tables S4 and S5). For example, ten GmMYB family proteins were quantitatively analyzed on one or more phosphorylated peptides. Interestingly, the phosphorylated peptide TVPSAsG in GmMYB I1KQI5 was detected in both cultivars, and another phosphorylated peptide FsPNLNQNPNPNLGK could only be detected in Union85140 (supplemental Tables S4 and S5), indicating that phosphorylation of the same protein could be modified at different sites in the two cultivars and might generate various activations. In addition, the phosphorylated peptide QKIDDsDESPNPK in GmMYB K7MQI8 in both cultivars was only detected at late time points (T12-T48) (supplemental Table S4). Similar results were observed in GmMYBs (K7LAB8 and I1JE71), GmbZIPs (Q00M78, I1JDF7, K7MV95, and C6T6L1), GmWRKY (I1MT25)and GmERF (I1KN17). This suggested a temporary regulation of this modification in response to salt stress.

To reveal the potential interaction network between TFs and differentially expressed proteins, the TF-specific binding motifs of the promoters from enzymes involved in chalcone metabolism are summarized in Fig. 7. Motif structures of these promoters were retrieved from the JASPAR database (50). All the promoters of genes encoding chalcone synthase (GmCHS), chalcone isomerase (GmCHI), and cytochrome P450 monooxygenase (GmCMP) were predicted to contain the conserved motifs recognized by MYBs, indicating that the promoters of these 13 enzymes should be regulated by this TF family. In addition, promoters of the two GmCMP and one GmCHI also included motifs recognized by bZIP. Additionally, GmERF had potential binding motifs in promoters of some GmCHS, GmCMP and GmCHI genes. Because their activities might be regulated by phosphorylation modifications, these TFs should play significant roles in the bridge between stress signal and the transcription of salt responsive genes.

An external file that holds a picture, illustration, etc.
Object name is zjw0111551750007.jpg

TFs specific binding motifs in promoters of GmCHS, GmCHI and GmCMP genes in soybean. All the promoters (2000 bp) of tested genes were scanned for discovering conserved motifs recognized by MYB, bZip and ERF TFs (88% threshold) at JASPAR (http://jaspar.genereg.net/cgi-bin/jaspar_db.pl?rm=browse&db=core&tax_group=plants). The abbreviations of gene names were listed in supplemental Table S3.

Rapid Function Tests of the Genes Involved in Chalcone Metabolism

In this research, the enzymes involved in chalcone metabolism were proposed to have potential correlations with soybean salt tolerance at both proteomic and transcriptional levels. To further validate that the chalcone synthase (CHS), chalcone isomerase (CHI) and cytochrome P450 monooxygenase (CPM) were determinants of plant salt-tolerance, gain-of-function and loss-of-function analyses were tested in soybean composites (Fig. 8) and A. thaliana mutants (Fig. 9), respectively, at seedling stage.

An external file that holds a picture, illustration, etc.
Object name is zjw0111551750008.jpg

Effects of salinity stress on seedlings of soybean composites. The seedlings of negative control (Union85140/pCAMBIA1301), gmchs-ox (Union85140/GmCHS), gmchi-ox (Union85140/GmCHI) and gmcpm-ox ((Union85140/GmCPM)) composites were treated in 1/2 MS medium with or without 100 mm NaCl for 10 days. Bar: 1 cm.

An external file that holds a picture, illustration, etc.
Object name is zjw0111551750009.jpg

Effects of salinity stress on seedlings of Arabidopsis thaliana mutants. The germination of Col-0 (WT), chs, chs/chi and chs/cpm plants grown in 1/2 Murashige and Skoog (MS) medium for 5 days and then transferred to 1/2 MS medium with or without 150 mm NaCl for 10 days. A and B, comparison of salt tolerance between WT and deletion mutant chs. C and D, comparison of salt tolerance between WT and double deletion mutant chs/chi. E and F: comparison of salt tolerance between WT and double deletion mutant chs/cpm. A, C, E: 0 mm NaCl; B, D, F: 150 mm NaCl; Bar: 2 cm.

When subjected to NaCl treatments, the Union85140/gmchs-ox composites showed higher tolerance than Union85140/pCAMBIA1301 (negative control) (Fig. 8), indicating that chalcone was a positive regulating factor in salt tolerance. Both of the gain-of-function Union85140/gmchi-ox and Union85140/gmcpm-ox composites (Fig. 8) showed a slight lower tolerance than the negative control. Similar results were observed in Arabidopsis, the single deletion mutant (chs) showed significantly lower tolerance than wild type (Fig. 9A and and99B), indicating that chalcone was a positive regulating factor in salt tolerance. Both of the loss-of-function double mutants chs/cpm (Fig. 9C and and99D) and chs/chi (Fig. 9E and and99F) also showed lower tolerance than wild type. However, these double mutants (chs/cpm and chs/chi) showed higher tolerance than the single deletion mutant (chs), suggesting that chalcone isomerase and cytochrome P450 monooxygenase were two negative regulating factors in salt tolerance. To summarize, chalcone synthase dominated the response to salt stress in chalcone metabolism.

DISCUSSION

Compared with the salt-sensitive Union85140, the salt-tolerant Wenfeng07 showed no significant advantage in exportation or compartmentalization of salts, but much higher capacity for ROS elimination within 48 h of NaCl treatment. Plants have evolved very complex mechanisms for ROS elimination at the transcription, translation and post-translational modification levels (12, 13, 15, 29, 51). The present study involved a comparative analysis of salt stress responses between a salt-tolerant and a salt-sensitive soybean variety using proteomic and phosphoproteomic approaches. Among them, 89 representative differentially expressed proteins were checked with their changes at transcriptional level using quantitative RT-PCR. Our results confirmed the view that expression differences at proteomic level are involved in functional proteins, whereas differences at phosphoproteomic level are mainly related to regulatory proteins (29). Interestingly, a series of proteins related to ROS scavenging and protein folding/degradation—such as GST, APX, SOD, heat shock protein 90–2, and Hsp70-Hsp90 organizing protein 1—were involved in salt responses of both salt-tolerant and salt-sensitive varieties, which were almost in accordance with previous studies (17, 52, 53). However, tolerance discriminations were possibly dominated by: (1) synthesis of flavonoid/isoflavonoid involved in the salicylic acid defense pathway by chalcone metabolism (54, 55) in Wenfeng07, compared with initiation of lateral roots by auxin response factor, auxin-induced protein AUX22 and PIN6a (10) in Union85140; (2) up-regulation of ERF and MYB TFs for activating MAPK and SOS pathways to eliminate ROS and excessive salts (12, 13) in Wenfeng07; and (3) regulating innate immunity via cytochrome P450 monooxygenase, chalcone isomerase, and sterol 24-C methyltransferase (56, 57) specifically in Wenfeng07.

However, phosphoproteomic comparisons revealed the details of dissimilarities in stress signal perception and transduction, transcription/translation of response genes and protein transporting. The protein samples were analyzed based on 2-DE MS/MS and LC MS/MS proteomics. A total of 89 differentially expressed nonredundant proteins were identified in LC MS/MS analysis and 90 in 2-DE MS/MS analysis. Of the 179 nonredundant differentially expressed proteins from LC-MS/MS and 2-DE MS/MS, 16 were also identified as phosphoproteins, including the stress-induced protein SAM22, histone H2A OS, eukaryotic translation initiation factor 3 subunit C, elongation factor 1-delta-like, fructose-bisphosphate aldolase, cytosolic glutamine synthetase GSbeta1 for signal transduction, chromosome remodeling, gene translation, energy and small molecular metabolism, etc.

Perception of Salinity and Signal Transduction

The SOS system (e.g. SOS1: H9CDQ2, supplemental Table S4) acts as a central hub in preventing Na+ toxicity in the plant, especially for Wenfeng07. The most common role of the SOS system is to sequestrate Na+ ions from the plant cytosol (58). In general, the high salt stress suddenly triggers a cytosolic Ca2+ signature (59), which can be perceived by the calcineurin B-like protein, SOS3 and Ser/Thr protein kinase, SOS2 (60). After perceiving the Ca2+ signature, SOS3 is phosphorylated by the protein kinase SOS2. The SOS2/SOS3 complex activates the plasma membrane Na+/H+ antiporter, SOS1. Downstream of the SOS cascade, SOS1 mediates Na+ efflux at the root epidermis (61). In our study, there were many SOS2 and SOS3 homologs found with multiphosphorylated sites and with different regulation levels. For example, GmSOS2 (K7KTI3) was observed with four phosphorylation sites, in which phosphor-Ser in peptide LPEsPREGSEEDNFLENLTGMPIR only occurred at early time points T0.5-T4, but the phosphor-Ser in peptide EGsEEDNFLENLTGMPIR only occurred at late time points T12-T48 (supplemental Table S4). Interestingly, GmSOS3 (C6T458), both in Wenfeng07 and Union85140, was detected at T0 and all treatment times except T4 (supplemental Table S4). In addition, another GmSOS3 homolog (K7KLX6), from both cultivars, was detected at time points T12-T48.

The Ca2+ signature could also be perceived by calcium-dependent protein kinases (CDPKs or CPKs) (62). The latter two transmit the signal into phosphorylation cascades capable of modulating gene expression and target protein activity (63). CDPKs, through their interaction with ion channels and transporters, seem to represent part of membrane-delimited plant stress responses (64). In the present study, the GmCDPK (D3G9M7) in Wenfeng07 showed much higher phosphorylation levels than Union85140 for time points T0.5-T48 (supplemental Table S4). This suggested that this GmCDPK might significantly contribute to the salt tolerance of Wenfeng07.

Reactive oxygen species (ROS) and hormones are key elements in intricate switches used by plants to trigger highly dynamic responses to changing environment. Although ROS may have deleterious effects in cells, they also act as signal transduction molecules involved in mediating responses to environmental stresses (65). Plant plasticity in response to the environment is linked to a complex signaling module in which ROS and antioxidants operate together with hormones, including auxin (66). The auxin resistant double-mutant tir1 afb2 showed increased tolerance to salinity as measured by chlorophyll content, germination rate and root elongation. In addition, mutant plants displayed reduced hydrogen peroxide (H2O2) and superoxide anion (O2−−.) levels, as well as enhanced antioxidant metabolism (67). Microarray analyses indicated that auxin responsive genes are repressed by oxidative and salt treatments in rice (68). More recently, the transcriptomic data of Blomster et al. (69) showed that various aspects of auxin homeostasis and signaling are modified by apoplastic ROS. Together, these findings suggest that the suppression of auxin signaling might be a strategy of plants to enhance their tolerance to abiotic stress, including salinity. In this study, the auxin response factor K7M7H1 was found with phosphorylated serine (in peptide sPPQPR). However, this modification was only detected at late time points T12-T48 (supplemental Table S4). Recent research found that a salt-responsive ethylene response factor1 (ERF1) regulates ROS-dependent signaling during the initial response to salt stress (13). However, the GmERF (I1KN17) was only observed with phosphorylation modification in the sensitive cultivar Union85140 (supplemental Table S4).

Other reported pathways of salt signaling include mitogen-activated protein kinase (MAPK or MPK) cascades (70). A MAPK cascade consists of a MAPK kinase (MAPKkk)–MAPK kinase (MAPKK/MKK)–MAPK module that links salt-signal receptors to downstream targets (71). For a rapid signal transduction, the GmMAPKK2 (Uniprot accession no. Q5JCL0) showed a much higher level phosphorylation modification after NaCl treatment in both Wenfeng07 and Union85140 (supplemental Table S4).

Metabolism of Small Molecules Related to Detoxification and Defense Pathways

Under salinity stress, the plant employs detoxification and defense pathways to increase their tolerance (58). Several abiotic stresses, such as salt, drought and cold can induce ROS accumulation including O2−−., H2O2 and hydroxyl radicals (10). Suitable concentrations of ROS are acquired as substrates in lipid, sugar and protein metabolisms. Peak values of ROS concentration usually act as signals for inducing ROS scavengers, which are mainly substrates involved in these metabolisms. In this study, copper amino oxidase and quinone oxidoreductase, which produces ROS (72, 73), were up-regulated after salt treatment. Meanwhile, universal scavengers, such as APX, SOD, GST and POD, also showed up-regulation in roots of both salt-sensitive and -tolerant soybean. Among these scavengers, APX has been shown to reduce H2O2 to H2O, with the concomitant generation of monodehydroascorbate. Many reports demonstrating that APX overexpression can enhance the salt tolerance of different plants have confirmed that APX plays an important role in scavenging ROS produced by salinity stress (7477). Moreover, the two homologs of APX might have different efficiencies in ROS elimination, because APX2 was significantly up-regulated in the tolerant cultivar (Wenfeng07), whereas APX1 was significantly up-regulated in the sensitive cultivar (Union85140) after salt treatment. This result is consistent with findings in two rice APXs (78).

Chalcone Metabolism Pathway is Involved in Soybean Tolerance to Salt Stress

Up to now, the chalcone metabolism pathway has mainly been considered as a feasible strategy for enhancing plant immunity to microbes (7981). In plants, chalcone biosynthesis begins with the hydroxylation of cinnamic acid by cytochrome P450 monooxygenase (82). The intermediate product p-coumaric acid is then activated by 4-coumaroyl:CoA ligase, yielding p-coumaroyl-coenzyme A (CoA) (83, 84). Subsequently, malonyl-CoA is added to p-coumaroyl-CoA and yields tetrahydroxychalcone by the enzyme chalcone synthase. Finally, chalcone isomerase converts the C15 compound tetrahydroxychalcone into (2S)-flavanones (8587). These flavonoids, including a diverse family of polyphenols, have been proven with health-promoting effects especially in preventing various human pathological risks (88, 89). Hence, significant amounts of research have been stimulated to elucidate the biosynthetic networks of flavonoids (90, 91). However, there are very few reports on the contribution of chalcone metabolism to plant salt tolerance (92, 93). Recently, a cytochrome P450 monooxygenase mutant was shown to be involved in a series of abiotic stresses including ABA and salt in Arabidopsis (94). Our proteomic and phosphoproteomic analyses showed that key enzymes, such as cytochrome P450 monooxygenase, chalcone synthase and chalcone isomerase, were correlated with salt stress especially in tolerant cultivar Wenfeng07. Their salt-responsive dynamics were also confirmed at the transcriptional level. The functions of these enzymes were preliminarily tested in soybean composites and Arabidopsis mutants. Both the gain of function and loss-of-function tests demonstrated that cytochrome P450 monooxygenase and chalcone isomerase were negatively related with salt tolerance in plant seedlings, whereas chalcone synthase was positively related.

Interestingly, 10 MYB (MYB like) transcription factors (TFs) were identified with significantly changed phosphorylation sites (supplemental Table S4 and S5). Commonly, MYB TFs play crucial roles in flavonol accumulation by regulating the expression of series genes coding for key enzymes involved in chalcone metabolism in plants (9597). In addition, three chalcone metabolism enzymes have been found in response to salt stress. These results indicate that the network between phosphorylated MYB TFs and chalcone metabolism enzymes might play potential crucial roles in soybean's tolerance to salinity.

CONCLUSION

Plants have evolved a set of physiological and biochemical responses for adaptation to salinity stress. Generally, glutathione and proline as well as several secondary metabolites, such as flavonoids, play a pivotal role in tolerance/detoxification of plants (98100).

In the present research, quantitative proteomic and phosphoproteomic analyses were conducted with both salt-tolerant (Wenfeng07) and -sensitive (Union85140) soybean varieties under salt stress. LC-MSMS and 2-D gel based proteomic analysis of these two variants from a series of time points after salt treatment identified 179 differentially expressed nonredundant proteins in total. Of these, 16 proteins also showed changes at phosphorylation level. These differential protein expression characteristics were mostly involved in functional pathways which possibly dominated the capacity of the two varieties concerning salt tolerance.

The quantitative phosphoproteomic analysis identified 3744 phosphorylated sites and 1163 differentially changed sites between the two cultivars, which revealed an activated signaling cascade involved in salt response. The comparison at phosphorylation level indicated that the hub signals fitted with salt tolerance in the tolerant variety.

In summary, the proteomic and phosphoproteomic comparisons between tolerant and sensitive variants could aid understanding of the response and defense mechanisms of soybean in response to salinity stress. The transcriptional and functional analyses confirmed the correlation of significantly changed proteins with salt tolerance. Moreover, the identified significantly changed proteins and phosphorylated sites provide an array of potential salt-response markers for future work. More importantly, the chalcone metabolism pathway was shown as a likely novel candidate for further research on plant salt tolerance. Based on these findings, we hypothesized a novel soybean salinity-tolerance pathway involved in chalcone metabolism (Fig. 10). After the perception of salinity signal, the GmMYBs are phosphorylated and further activated the genes GmCHS, GmCHI and GmCPM. Then, these activated key enzymes GMCHS, GmCHI and GmCPM mediated the accumulation patterns of flavonoids. Finally, these flavonoids appropriately reduced the ROS or play roles in other functions for enhancing the soybean's tolerance to salinity.

An external file that holds a picture, illustration, etc.
Object name is zjw0111551750010.jpg

A hypothetical model for transcription factor GmMYB in regulating genes GmCHS, GmCHI and GmCPM during soybean's response to salinity. After the perception of salinity signal, the GmMYBs are phosphorylated and further activated the genes GmCHS, GmCHI and GmCPM. Then, these key enzymes GMCHS, GmCHI and GmCPM regulated the accumulation patterns of flavonoids. Finally, these flavonoids appropriately reduced the ROS or play roles in other functions for soybean's tolerance to salinity.

Supplementary Material

Supplemental Data:

Acknowledgments

We thank Prof. Rencun Jin, Prof. Weiqin Zhu, Miss Yanqin Gu and Mr Zhengzhe Zhang, from Hangzhou Normal University, for their kind assistance during the ion content analysis. Special appreciations should also be paid to Dr. Rui Wang and Dr. Xiaojing Gao, from Shanghai Applied Protein Technology Co. Ltd., for their kind help for bioinformatics analysis. The mass spectrometry proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE partner repository with the dataset identifier PXD002856.

Footnotes

Contributed by

Author contributions: E.P. and S.N. designed research; E.P., L. Qu, J.H., Y.H., C.L., T.P., Y.Z., and L.D. performed research; E.P., H.W., and S.N. contributed new reagents or analytic tools; E.P., L. Qu, J.H., Y.H., H.L., B.J., and L.D. analyzed data; E.P., S.T., and S.N. wrote the paper; L. Qiu provide the soybean seed.

* This work was partially supported by grants 31301053 and U1130304 from the National Science Foundation of China, the Hong Kong RGC Collaborative Research Fund (CUHK3/CRF/11G), the Hong Kong RGC General Research Fund (468610), the Lo Kwee-Seong Biomedical Research Fund and Lee Hysan Foundation, and grant PF14002004014 from Hangzhou Normal University.

An external file that holds a picture, illustration, etc.
Object name is sbox.jpg This article contains supplemental material.

1 The abbreviations used are:

ROS
reactive oxygen species
SOS
salt overly sensitive
CDPK
calcium dependent protein kinase
MAPK
mitogen activated protein kinase.

REFERENCES

1. Berman K. H., Harrigan G. G., Riordan S. G., Nemeth M. A., Hanson C., Smith M., Sorbet R., Zhu E., and Ridley W. P. (2009) Compositions of seed, forage, and processed fractions from insect-protected soybean MON 87701 are equivalent to those of conventional soybean. J. Agr. Food Chem. 57, 11360–11369 [PubMed] [Google Scholar]
2. Prakash D., Niranjan A., Tewari S. K., and Pushpangadan P. (2001) Underutilised legumes: potential sources for low-cost protein. Int. J. Food Sci. Nutr. 52, 337–341 [PubMed] [Google Scholar]
3. Lam H. M., Xu X., Liu X., Chen W., Yang G., Wong F. L., Li M. W., He W., Qin N., Wang B., Li J., Jian M., Wang J., Shao G., Wang J., Sun S. S., and Zhang G. (2010) Resequencing of 31 wild and cultivated soybean genomes identifies patterns of genetic diversity and selection. Nat. Genet. 42, 1053–1059 [PubMed] [Google Scholar]
4. Li W. Y., Shao G., and Lam H. M. (2008) Ectopic expression of GmPAP3 alleviates oxidative damage caused by salinity and osmotic stresses. New Phyt. 178, 80–91 [PubMed] [Google Scholar]
5. Yamaguchi T., and Blumwald E. (2005) Developing salt-tolerant crop plants: challenges and opportunities. Trends Plant Sci. 10, 615–620 [PubMed] [Google Scholar]
6. Munns R., and Tester M. (2008) Mechanisms of salinity tolerance. Annu. Rev. Plant Biol. 59, 651–681 [PubMed] [Google Scholar]
7. Wang W., Vinocur B., and Altman A. (2003) Plant responses to drought, salinity and extreme temperatures: towards genetic engineering for stress tolerance. Planta 218, 1–14 [PubMed] [Google Scholar]
8. Kim B. G., Waadt R., Cheong Y. H., Pandey G. K., Dominguez-Solis J. R., Schultke S., Lee S. C., Kudla J., and Luan S. (2007) The calcium sensor CBL10 mediates salt tolerance by regulating ion homeostasis in Arabidopsis. Plant J. 52, 473–484 [PubMed] [Google Scholar]
9. Ward J. M., Hirschi K. D., and Sze H. (2003) Plants pass the salt. Trends Plant Sci. 8, 200–201 [PubMed] [Google Scholar]
10. Galvan-Ampudia C. S., and Testerink C. (2011) Salt stress signals shape the plant root. Curr. Opin. Plant Biol. 14, 296–302 [PubMed] [Google Scholar]
11. Osakabe Y., Yamaguchi-Shinozaki K., Shinozaki K., and Tran L. S. P. (2013) Sensing the environment: key roles of membrane-localized kinases in plant perception and response to abiotic stress. J. Exp. Bot. 64, 445–458 [PubMed] [Google Scholar]
12. Yu L. J., Nie J. N., Cao C. Y., Jin Y. K., Yan M., Wang F. Z., Liu J., Xiao Y., Liang Y. H., and Zhang W. H. (2010) Phosphatidic acid mediates salt stress response by regulation of MPK6 in Arabidopsis thaliana. New Phytol. 188, 762–773 [PubMed] [Google Scholar]
13. Schmidt R., Mieulet D., Hubberten H. M., Obata T., Hoefgen R., Fernie A. R., Fisahn J., Segundo B. S., Guiderdoni E., Schippers J. H. M., and Mueller-Roeber B. (2013) SALT-RESPONSIVE ERF1 regulates reactive oxygen species-dependent signaling during the initial response to salt stress in rice. Plant Cell 25, 2115–2131 [PMC free article] [PubMed] [Google Scholar]
14. Latz A., Mehlmer N., Zapf S., Mueller T. D., Wurzinger B., Pfister B., Csaszar E., Hedrich R., Teige M., and Becker D. (2013) Salt stress triggers phosphorylation of the Arabidopsis vacuolar K channel TPK1 by calcium-dependent protein kinases (CDPKs). Mol. Plant 6, 1274–1289 [PMC free article] [PubMed] [Google Scholar]
15. Lin H. X., Yang Y. Q., Quan R. D., Mendoza I., Wu Y. S., Du W. M., Zhao S. S., Schumaker K. S., Pardo J. M., and Guo Y. (2009) Phosphorylation of SOS3-LIKE CALCIUM BINDING PROTEIN8 by SOS2 protein kinase stabilizes their protein complex and regulates salt tolerance in Arabidopsis. Plant cell 21, 1607–1619 [PMC free article] [PubMed] [Google Scholar]
16. Yan J. H., Wang B. A., Jiang Y. N., Cheng L. J., and Wu T. L. (2014) GmFNSII-Controlled soybean flavone metabolism responds to abiotic stresses and regulates plant salt tolerance. Plant Cell Physiol. 55, 74–86 [PubMed] [Google Scholar]
17. Aghaei K., Ehsanpour A., Shah A., and Komatsu S. (2009) Proteome analysis of soybean hypocotyl and root under salt stress. Amino Acids 36, 91–98 [PubMed] [Google Scholar]
18. Toorchi M., Yukawa K., Nouri M.-Z., and Komatsu S. (2009) Proteomics approach for identifying osmotic-stress-related proteins in soybean roots. Peptides 30, 2108–2117 [PubMed] [Google Scholar]
19. Ge Y., Li Y., Zhu Y.-M., Bai X., Lv D.-K., Guo D., Ji W., and Cai H. (2010) Global transcriptome profiling of wild soybean (Glycine soja) roots under NaHCO3 treatment. BMC Plant Biol. 10, 153. [PMC free article] [PubMed] [Google Scholar]
20. Lu Y., Lam H., Pi E., Zhan Q., Tsai S., Wang C., Kwan Y., and Ngai S. (2013) Comparative metabolomics in Glycine max and Glycine soja under salt stress to reveal the phenotypes of their offspring. J. Agr. Food Chem. 61, 8711–8721 [PubMed] [Google Scholar]
21. Qin J., Gu F., Liu D., Yin C., Zhao S., Chen H., Zhang J., Yang C., Zhan X., and Zhang M. (2013) Proteomic analysis of elite soybean Jidou17 and its parents using iTRAQ-based quantitative approaches. Proteome Sci. 11, 12. [PMC free article] [PubMed] [Google Scholar]
22. Ficarro S. B., McCleland M. L., Stukenberg P. T., Burke D. J., Ross M. M., Shabanowitz J., Hunt D. F., and White F. M. (2002) Phosphoproteome analysis by mass spectrometry and its application to Saccharomyces cerevisiae. Nat Biotechnol 20, 301–305 [PubMed] [Google Scholar]
23. Gruhler A., Olsen J. V., Mohammed S., Mortensen P., Færgeman N. J., Mann M., and Jensen O. N. (2005) Quantitative phosphoproteomics applied to the yeast pheromone signaling pathway. Mol. Cell. Proteomics 4, 310–327 [PubMed] [Google Scholar]
24. Beausoleil S. A., Jedrychowski M., Schwartz D., Elias J. E., Villén J., Li J., Cohn M. A., Cantley L. C., and Gygi S. P. (2004) Large-scale characterization of HeLa cell nuclear phosphoproteins. Proc. Natl. Acad. Sci. U.S.A. 101, 12130–12135 [PMC free article] [PubMed] [Google Scholar]
25. Larsen M. R., Thingholm T. E., Jensen O. N., Roepstorff P., and Jørgensen T. J. (2005) Highly selective enrichment of phosphorylated peptides from peptide mixtures using titanium dioxide microcolumns. Mol. Cell. Proteomics 4, 873–886 [PubMed] [Google Scholar]
26. Mazanek M., Mituloviae G., Herzog F., Stingl C., Hutchins J. R., Peters J. M., and Mechtler K. (2007) Titanium dioxide as a chemo-affinity solid phase in offline phosphopeptide chromatography prior to HPLC-MS/MS analysis. Nat. Protoc. 2, 1059–1069 [PubMed] [Google Scholar]
27. Guan R. X., Qu Y., Guo Y., Yu L. L., Liu Y., Jiang J. H., Chen J. G., Ren Y. L., Liu G. Y., Tian L., Jin L. G., Liu Z. X., Hong H. L., Chang R. Z., Gilliham M., and Qiu L. J. (2014) Salinity tolerance in soybean is modulated by natural variation in GmSALT3. Plant J. 80, 937–950 [PubMed] [Google Scholar]
28. Yuan C. P., Zhou G., Li Y. H., Wang K. J., Wang Z., Li X. H., Chang R. Z., Qiu L. J. (2008) Cloning and sequence diversity analysis of GmHs1pro-1 in Chinese domesticated and wild soybeans. Mol Breeding 22, 593–602 [Google Scholar]
29. Lv D. W., Subburaj S., Cao M., Yan X., Li X. H., Appels R., Sun D. F., Ma W. J., and Yan Y. M. (2014) Proteome and phosphoproteome characterization reveals new response and defense mechanisms of Brachypodium distachyon leaves under salt stress. Mol. Cell. Proteomics 13, 632–652 [PMC free article] [PubMed] [Google Scholar]
30. Kao S. H., Wong H. K., Chiang C. Y., and Chen H. M. (2008) Evaluating the compatibility of three colorimetric protein assays for two-dimensional electrophoresis experiments. Proteomics 8, 2178–2184 [PubMed] [Google Scholar]
31. Wisniewski J. R., Zougman A., and Mann M. (2009) Combination of FASP and StageTip-based fractionation allows in-depth analysis of the hippocampal membrane proteome. J. Proteome Res. 8, 5674–5678 [PubMed] [Google Scholar]
32. Wisniewski J. R., Zougman A., Nagaraj N., and Mann M. (2009) Universal sample preparation method for proteome analysis. Nat. Methods 6, 359-U360 [PubMed] [Google Scholar]
33. Ostasiewicz P., Zielinska D. F., Mann M., and Wisniewski J. R. (2010) Proteome, phosphoproteome, and N-glycoproteome are quantitatively preserved in formalin-fixed paraffin-embedded tissue and analyzable by high-resolution mass spectrometry. J. Proteome Res. 9, 3688–3700 [PubMed] [Google Scholar]
34. Dong M., Gu J., Zhang L., Chen P., Liu T., Deng J., Lu H., Han L., and Zhao B. (2014) Comparative proteomics analysis of superior and inferior spikelets in hybrid rice during grain filling and response of inferior spikelets to drought stress using isobaric tags for relative and absolute quantification. J. Proteomics 109, 382–399 [PubMed] [Google Scholar]
35. Cox J., and Mann M. (2008) MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol 26, 1367–1372 [PubMed] [Google Scholar]
36. Tran H. N. N., Brechenmacher L., Aldrich J. T., Clauss T. R., Gritsenko M. A., Hixson K. K., Libault M., Tanaka K., Yang F., Yao Q. M., Pasa-Tolic L., Xu D., Nguyen H. T., and Stacey G. (2012) Quantitative phosphoproteomic analysis of soybean root hairs inoculated with Bradyrhizobium japonicum. Mol. Cell. Proteomics 11, 1140–1155 [PMC free article] [PubMed] [Google Scholar]
37. Jones K. A., Kim P. D., Patel B. B., Kelsen S. G., Braverman A., Swinton D. J., Gafken P. R., Jones L. A., Lane W. S., Neveu J. M., Leung H. C., Shaffer S. A., Leszyk J. D., Stanley B. A., Fox T. E., Stanley A., Hall M. J., Hampel H., South C. D., de la Chapelle A., Burt R. W., Jones D. A., Kopelovich L., and Yeung A. T. (2013) Immunodepletion plasma proteomics by tripleTOF 5600 and Orbitrap elite/LTQ-Orbitrap Velos/Q exactive mass spectrometers. J. Proteome Res. 12, 4351–4365 [PMC free article] [PubMed] [Google Scholar]
38. Wang G., Zeng H., Hu X., Zhu Y., Chen Y., Shen C., Wang H., Poovaiah B. W., and Du L. (2015) Identification and expression analyses of calmodulin-binding transcription activator genes in soybean. Plant Soil 386, 205–221 [Google Scholar]
39. Ye J., Coulouris G., Zaretskaya I., Cutcutache I., Rozen S., and Madden T. L. (2012) Primer-BLAST: a tool to design target-specific primers for polymerase chain reaction. BMC Bioinformatics 13, 134. [PMC free article] [PubMed] [Google Scholar]
40. Schwartz D., and Gygi S. P. (2005) An iterative statistical approach to the identification of protein phosphorylation motifs from large-scale data sets. Nat. Biotechnol. 23, 1391–1398 [PubMed] [Google Scholar]
41. Mathelier A., Zhao X. B., Zhang A. W., Parcy F., Worsley-Hunt R., Arenillas D. J., Buchman S., Chen C. Y., Chou A., Ienasescu H., Lim J., Shyr C., Tan G., Zhou M., Lenhard B., Sandelin A., and Wasserman W. W. (2014) JASPAR 2014: an extensively expanded and updated open-access database of transcription factor binding profiles. Nucleic Acids Res. 42, D142-D147 [PMC free article] [PubMed] [Google Scholar]
42. Bryne J. C., Valen E., Tang M. H. E., Marstrand T., Winther O., da Piedade I., Krogh A., Lenhard B., and Sandelin A. (2008) JASPAR, the open access database of transcription factor-binding profiles: new content and tools in the 2008 update. Nucleic Acids Res. 36, D102-D106 [PMC free article] [PubMed] [Google Scholar]
43. Zhang Z., Pang X., Xuewu D., Ji Z., and Jiang Y. (2005) Role of peroxidase in anthocyanin degradation in litchi fruit pericarp. Food Chem. 90, 47–52 [Google Scholar]
44. Re R., Pellegrini N., Proteggente A., Pannala A., Yang M., and Rice-Evans C. (1999) Antioxidant activity applying an improved ABTS radical cation decolorization assay. Free Radical Biol. Med. 26, 1231–1237 [PubMed] [Google Scholar]
45. Qi X. P., Li M. W., Xie M., Liu X., Ni M., Shao G. H., Song C., Yim A. K. Y., Tao Y., Wong F. L., Isobe S., Wong C. F., Wong K. S., Xu C. Y., Li C. Q., Wang Y., Guan R., Sun F. M., Fan G. Y., Xiao Z. X., Zhou F., Phang T. H., Liu X., Tong S. W., Chan T. F., Yiu S. M., Tabata S., Wang J., Xu X., and Lam H. M. (2014) Identification of a novel salt tolerance gene in wild soybean by whole-genome sequencing. Nat. Commun. 5, 4340–4350 [PMC free article] [PubMed] [Google Scholar]
46. Shi H., Lee B. H., Wu S. J., and Zhu J. K. (2003) Overexpression of a plasma membrane Na+/H+ antiporter gene improves salt tolerance in Arabidopsis thaliana. Nat. Biotechnol. 21, 81–85 [PubMed] [Google Scholar]
47. Dudonne S., Vitrac X., Coutiere P., Woillez M., and Mérillon J.-M. (2009) Comparative study of antioxidant properties and total phenolic content of 30 plant extracts of industrial interest using DPPH, ABTS, FRAP, SOD, and ORAC assays. J. Agr. Food Chem. 57, 1768–1774 [PubMed] [Google Scholar]
48. Zulawski M., Braginets R., and Schulze W. X. (2013) PhosPhAt goes kinases-searchable protein kinase target information in the plant phosphorylation site database PhosPhAt. Nucleic Acids Res. 41, D1176-D1184 [PMC free article] [PubMed] [Google Scholar]
49. Huang H. D., Lee T. Y., Tzeng S. W., and Horng J. T. (2005) KinasePhos: a web tool for identifying protein kinase-specific phosphorylation sites. Nucleic Acids Res. 33, W226–229 [PMC free article] [PubMed] [Google Scholar]
50. Broin P. O., Smith T. J., and Golden A. (2015) Alignment-free clustering of transcription factor binding motifs using a genetic-k-medoids approach. BMC Bioinformatics 16, 22. [PMC free article] [PubMed] [Google Scholar]
51. Matsuura H., Ishibashi Y., Shinmyo A., Kanaya S., and Kato K. (2010) Genome-wide analyses of early translational responses to elevated temperature and high salinity in Arabidopsis thaliana. Plant Cell Physiol. 51, 448–462 [PubMed] [Google Scholar]
52. Wu T., Pi E.-X., Tsai S.-N., Lam H.-M., Sun S.-M., Kwan Y. W., and Ngai S.-M. (2011) GmPHD5 acts as an important regulator for crosstalk between histone H3K4 di-methylation and H3K14 acetylation in response to salinity stress in soybean. BMC Plant Biol. 11, 178. [PMC free article] [PubMed] [Google Scholar]
53. Sobhanian H., Razavizadeh R., Nanjo Y., Ehsanpour A. A., Jazii F. R., Motamed N., and Komatsu S. (2010) Proteome analysis of soybean leaves, hypocotyls and roots under salt stress. Proteome Sci. 8, 1–15 [PMC free article] [PubMed] [Google Scholar]
54. Chen F., Li B., Li G., Charron J. B., Dai M., Shi X., and Deng X. W. (2014) Arabidopsis phytochrome a directly targets numerous promoters for individualized modulation of genes in a wide range of pathways. Plant Cell 26, 1949–1966 [PMC free article] [PubMed] [Google Scholar]
55. Alimohammadi M., de Silva K., Ballu C., Ali N., and Khodakovskaya M. V. (2012) Reduction of inositol (1,4,5)-trisphosphate affects the overall phosphoinositol pathway and leads to modifications in light signalling and secondary metabolism in tomato plants. J. Exp. Bot. 63, 825–835 [PMC free article] [PubMed] [Google Scholar]
56. Wang K. R., Senthil-Kumar M., Ryu C. M., Kang L., and Mysore K. S. (2012) Phytosterols play a key role in plant innate immunity against bacterial pathogens by regulating nutrient efflux into the apoplast. Plant Physiol. 158, 1789–1802 [PMC free article] [PubMed] [Google Scholar]
57. O'Brien M., Chantha S. C., Rahier A., and Matton D. P. (2005) Lipid signaling in plants. Cloning and expression analysis of the obtusifoliol 14alpha-demethylase from Solanum chacoense Bitt., a pollination- and fertilization-induced gene with both obtusifoliol and lanosterol demethylase activity. Plant Physiol. 139, 734–749 [PMC free article] [PubMed] [Google Scholar]
58. Zhu J. K. (2003) Regulation of ion homeostasis under salt stress. Curr. Opin. Plant Biol. 6, 441–445 [PubMed] [Google Scholar]
59. Knight H., Trewavas A. J., and Knight M. R. (1997) Calcium signalling in Arabidopsis thaliana responding to drought and salinity. Plant J. 12, 1067–1078 [PubMed] [Google Scholar]
60. Qiu Q.-S., Guo Y., Dietrich M. A., Schumaker K. S., and Zhu J.-K. (2002) Regulation of SOS1, a plasma membrane Na+/H+ exchanger in Arabidopsis thaliana, by SOS2 and SOS3. Proc. Natl. Acad. Sci. U.S.A. 99, 8436–8441 [PMC free article] [PubMed] [Google Scholar]
61. Wu S.-J., Ding L., and Zhu J.-K. (1996) SOS1, a genetic locus essential for salt tolerance and potassium acquisition. Plant Cell 8, 617–627 [PMC free article] [PubMed] [Google Scholar]
62. Kudla J., Batistič O., and Hashimoto K. (2010) Calcium signals: the lead currency of plant information processing. Plant Cell 22, 541–563 [PMC free article] [PubMed] [Google Scholar]
63. Curran A., Chang F., Chang C.-L., Garg S., Miguel R. M., Barron Y. D., Li Y., Romanowsky S., Cushman J. C., and Gribskov M. (2011) Calcium-dependent protein kinases from Arabidopsis show substrate specificity differences in an analysis of 103 substrates. Frontiers Plant Sci. 2, 1–15 [PMC free article] [PubMed] [Google Scholar]
64. Hedrich R., and Kudla J. (2006) Calcium signaling networks channel plant K+ uptake. Cell 125, 1221–1223 [PubMed] [Google Scholar]
65. Miller G., Suzuki N., CIFTCI-YILMAZ S., and Mittler R. (2010) Reactive oxygen species homeostasis and signalling during drought and salinity stresses. Plant Cell Environ. 33, 453–467 [PubMed] [Google Scholar]
66. Tognetti V. B., Mühlenbock P., and Van Breusegem F. (2012) Stress homeostasis–the redox and auxin perspective. Plant Cell Environ. 35, 321–333 [PubMed] [Google Scholar]
67. Iglesias M. J., Terrile M. C., Bartoli C. G., D'Ippólito S., and Casalongué C. A. (2010) Auxin signaling participates in the adaptative response against oxidative stress and salinity by interacting with redox metabolism in Arabidopsis. Plant Mol. Biol. 74, 215–222 [PubMed] [Google Scholar]
68. Jain M., and Khurana J. P. (2009) Transcript profiling reveals diverse roles of auxin-responsive genes during reproductive development and abiotic stress in rice. FEBS J. 276, 3148–3162 [PubMed] [Google Scholar]
69. Blomster T., Salojärvi J., Sipari N., Brosché M., Ahlfors R., Keinänen M., Overmyer K., and Kangasjärvi J. (2011) Apoplastic reactive oxygen species transiently decrease auxin signaling and cause stress-induced morphogenic response in Arabidopsis. Plant Physiol. 157, 1866–1883 [PMC free article] [PubMed] [Google Scholar]
70. Nakagami H., Pitzschke A., and Hirt H. (2005) Emerging MAP kinase pathways in plant stress signalling. Trends Plant Sci. 10, 339–346 [PubMed] [Google Scholar]
71. Ichimura K., Mizoguchi T., Yoshida R., Yuasa T., and Shinozaki K. (2000) Various abiotic stresses rapidly activate Arabidopsis MAP kinases ATMPK4 and ATMPK6. Plant J. 24, 655–665 [PubMed] [Google Scholar]
72. Page C. C., Moser C. C., Chen X., and Dutton P. L. (1999) Natural engineering principles of electron tunnelling in biological oxidation-reduction. Nature 402, 47–52 [PubMed] [Google Scholar]
73. Brandt U. (2006) Energy converting NADH: quinone oxidoreductase (complex I). Annu. Rev. Biochem. 75, 69–92 [PubMed] [Google Scholar]
74. Wang J., Zhang H., and Allen R. D. (1999) Overexpression of an Arabidopsis peroxisomal ascorbate peroxidase gene in tobacco increases protection against oxidative stress. Plant Cell Physiol. 40, 725–732 [PubMed] [Google Scholar]
75. Lee S.-H., Ahsan N., Lee K.-W., Kim D.-H., Lee D.-G., Kwak S.-S., Kwon S.-Y., Kim T.-H., and Lee B.-H. (2007) Simultaneous overexpression of both CuZn superoxide dismutase and ascorbate peroxidase in transgenic tall fescue plants confers increased tolerance to a wide range of abiotic stresses. J. Plant Physiol. 164, 1626–1638 [PubMed] [Google Scholar]
76. Yoshimura K., Yabuta Y., Ishikawa T., and Shigeoka S. (2000) Expression of spinach ascorbate peroxidase isoenzymes in response to oxidative stresses. Plant Physiol. 123, 223–234 [PMC free article] [PubMed] [Google Scholar]
77. Allen R. D. (1995) Dissection of oxidative stress tolerance using transgenic plants. Plant Physiol. 107, 1049. [PMC free article] [PubMed] [Google Scholar]
78. Lu Z., Liu D., and Liu S. (2007) Two rice cytosolic ascorbate peroxidases differentially improve salt tolerance in transgenic Arabidopsis. Plant Cell Rep. 26, 1909–1917 [PubMed] [Google Scholar]
79. O'Keefe D. P., Tepperman J. M., Dean C., Leto K. J., Erbes D. L., and Odell J. T. (1994) Plant expression of a bacterial cytochrome P450 that catalyzes activation of a sulfonylurea pro-herbicide. Plant Physiol. 105, 473–482 [PMC free article] [PubMed] [Google Scholar]
80. Dixon R. A. (2001) Natural products and plant disease resistance. Nature 411, 843–847 [PubMed] [Google Scholar]
81. Weisshaar B., and Jenkins G. I. (1998) Phenylpropanoid biosynthesis and its regulation. Curr. Opin. Plant Biol. 1, 251–257 [PubMed] [Google Scholar]
82. Zhu Y., Nomura T., Xu Y., Zhang Y., Peng Y., Mao B., Hanada A., Zhou H., Wang R., Li P., Zhu X., Mander L. N., Kamiya Y., Yamaguchi S., and He Z. (2006) ELONGATED UPPERMOST INTERNODE encodes a cytochrome P450 monooxygenase that epoxidizes gibberellins in a novel deactivation reaction in rice. Plant Cell 18, 442–456 [PMC free article] [PubMed] [Google Scholar]
83. Mayer M. J., Narbad A., Parr A. J., Parker M. L., Walton N. J., Mellon F. A., and Michael A. J. (2001) Rerouting the plant phenylpropanoid pathway by expression of a novel bacterial enoyl-CoA hydratase/lyase enzyme function. Plant Cell 13, 1669–1682 [PMC free article] [PubMed] [Google Scholar]
84. Morita H., Shimokawa Y., Tanio M., Kato R., Noguchi H., Sugio S., Kohno T., and Abe I. (2010) A structure-based mechanism for benzalacetone synthase from Rheum palmatum. Proc. Natl. Acad. Sci. U.S.A. 107, 669–673 [PMC free article] [PubMed] [Google Scholar]
85. Heller W. (1986) Flavonoid biosynthesis, an overview. Prog. Clin. Biol. Res. 213, 25–42 [PubMed] [Google Scholar]
86. Pelletier M. K., and Shirley B. W. (1996) Analysis of flavanone 3-hydroxylase in Arabidopsis seedlings. Coordinate regulation with chalcone synthase and chalcone isomerase. Plant Physiol. 111, 339–345 [PMC free article] [PubMed] [Google Scholar]
87. Hur S., Newby Z. E., and Bruice T. C. (2004) Transition state stabilization by general acid catalysis, water expulsion, and enzyme reorganization in Medicago savita chalcone isomerase. Proc. Natl. Acad. Sci. U.S.A. 101, 2730–2735 [PMC free article] [PubMed] [Google Scholar]
88. Arts I. C., and Hollman P. C. (2005) Polyphenols and disease risk in epidemiologic studies. Am. J. Clin. Nutr. 81, 317S-325S [PubMed] [Google Scholar]
89. Hughes D. A. (2005) Plant polyphenols: modifiers of immune function and risk of cardiovascular disease. Nutrition 21, 422–423 [PubMed] [Google Scholar]
90. Xue Y., Zhang Y., Cheng D., Daddy S., and He Q. (2014) Genetically engineering Synechocystis sp. Pasteur Culture Collection 6803 for the sustainable production of the plant secondary metabolite p-coumaric acid. Proc. Natl. Acad. Sci. U.S.A. 111, 9449–9454 [PMC free article] [PubMed] [Google Scholar]
91. Yan Y., Kohli A., and Koffas M. A. (2005) Biosynthesis of natural flavanones in Saccharomyces cerevisiae. Appl. Environ Microb. 71, 5610–5613 [PMC free article] [PubMed] [Google Scholar]
92. Chen L. J., Guo H. M., Lin Y., and Cheng H. M. (2015) Chalcone synthase EaCHS1 from Eupatorium adenophorum functions in salt stress tolerance in tobacco. Plant Cell Rep. 34, 885–894 [PubMed] [Google Scholar]
93. Wang H., Hu T. J., Huang J. Z., Lu X., Huang B. Q., and Zheng Y. Z. (2013) The expression of Millettia pinnata chalcone isomerase in Saccharomyces cerevisiae salt-sensitive mutants enhances salt-tolerance. Int. J. Mol. Sci. 14, 8775–8786 [PMC free article] [PubMed] [Google Scholar]
94. Mao G., Seebeck T., Schrenker D., and Yu O. (2013) CYP709B3, a cytochrome P450 monooxygenase gene involved in salt tolerance in Arabidopsis thaliana. BMC Plant Biol. 13, 169. [PMC free article] [PubMed] [Google Scholar]
95. Stracke R., Ishihara H., Huep G., Barsch A., Mehrtens F., Niehaus K., and Weisshaar B. (2007) Differential regulation of closely related R2R3-MYB transcription factors controls flavonol accumulation in different parts of the Arabidopsis thaliana seedling. Plant J. 50, 660–677 [PMC free article] [PubMed] [Google Scholar]
96. Liao Y., Zou H. F., Wang H. W., Zhang W. K., Ma B., Zhang J. S., and Chen S. Y. (2008) Soybean GmMYB76, GmMYB92, and GmMYB177 genes confer stress tolerance in transgenic Arabidopsis plants. Cell Res. 18, 1047–1060 [PubMed] [Google Scholar]
97. Du H., Wang Y. B., Xie Y., Liang Z., Jiang S. J., Zhang S. S., Huang Y. B., and Tang Y. X. (2013) Genome-wide identification and evolutionary and expression analyses of MYB-related genes in land plants. DNA Res. 20, 437–448 [PMC free article] [PubMed] [Google Scholar]
98. Lafuente A., Perez-Palacios P., Doukkali B., Molina-Sanchez M. D., Jimenez-Zurdo J. I., Caviedes M. A., Rodriguez-Llorente I. D., and Pajuelo E. (2015) Unraveling the effect of arsenic on the model Medicago-Ensifer interaction: a transcriptomic meta-analysis. New Phytol. 205, 255–272 [PubMed] [Google Scholar]
99. Hellmann H., Funck D., Rentsch D., and Frommer W. B. (2000) Hypersensitivity of an Arabidopsis sugar signaling mutant toward exogenous proline application. Plant Physiol. 122, 357–368 [PMC free article] [PubMed] [Google Scholar]
100. Hellmann H., Funck D., Rentsch D., and Frommer W. B. (2000) Hypersensitivity of an Arabidopsis sugar signaling mutant toward exogenous proline application. Plant Physiol. 123, 779–789 [PMC free article] [PubMed] [Google Scholar]

Articles from Molecular & Cellular Proteomics : MCP are provided here courtesy of American Society for Biochemistry and Molecular Biology

-