IMR Press / FBL / Volume 28 / Issue 3 / DOI: 10.31083/j.fbl2803049
Open Access Original Research
Physiological and Proteomic Analysis of Seed Germination under Salt Stress in Mulberry
Show Less
1 College of Horticulture and Landscape Architecture, Zhongkai University of Agriculture and Engineering, 510225 Guangzhou, Guangdong, China
*Correspondence: wangyi@zhku.edu.cn (Yi Wang)
These authors contributed equally.
Front. Biosci. (Landmark Ed) 2023, 28(3), 49; https://doi.org/10.31083/j.fbl2803049
Submitted: 15 September 2022 | Revised: 11 November 2022 | Accepted: 9 December 2022 | Published: 13 March 2023
Copyright: © 2023 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Background: Salinity is the main abiotic stress that affects seed germination, plant growth and crop production. Plant growth begins with seed germination, which is closely linked to crop development and final yields. Morus alba L. is a well-known saline-alkaline tree with economic value in China, and the most prominent method of expanding mulberry tree populations is seed propagation. Understanding the molecular mechanism of Morus alba L. salt tolerance is crucial for identifying salt-tolerant proteins in seed germination. Here, we explored the response mechanism of mulberry seed germination to salt stress at physiological and protein omics levels. Methods: Tandem mass tag (TMT)-based proteomic profiling of Morus alba L. seeds germinated under 50 mM and 100 mM NaCl treatment for 14 days was performed, and the proteomic findings were validated through parallel reaction monitoring (PRM). Results: Physiological data showed that salt stress inhibited the germination rate and radicle length of mulberry seeds, decreased the malondialdehyde (MDA) content and significantly increased superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT) activities. Then, a TMT marker technique was used to analyze the protein groups in mulberry seeds with two salt treatment stages, and 76,544 unique peptides were detected. After removing duplicate proteins, 7717 proteins were identified according to TMT data, and 143 (50 mM NaCl) and 540 (100 mM NaCl) differentially abundant proteins (DAPs) were screened out. Compared with the control, in the 50 mM NaCl solution, 61 and 82 DAPs were upregulated and downregulated, respectively, and in the 100 mM NaCl solution, 222 and 318 DAPs were upregulated and downregulated, respectively. Furthermore, 113 DAPs were copresent in the 50 mM and 100 mM NaCl treatments, of which 43 were upregulated and 70 were downregulated. Gene Ontology (GO) annotation and Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment analysis revealed that the DAPs induced by salt stress during mulberry seed germination were mainly involved in photosynthesis, carotenoid biosynthesis and phytohormone signaling. Finally, PRM verified five differentially expressed proteins, which demonstrated the reliability of TMT in analyzing protein groups. Conclusions: Our research provides valuable insights to further study the overall mechanism of salt stress responses and salt tolerance of mulberry and other plants.

Keywords
Morus alba L.
proteomic
seed germination
salt stress
tandem mass tag
parallel reaction monitoring
1. Introduction

Soil salinization is among the most serious abiotic factors responsible for decreasing crop yields worldwide; the damage is mainly due to high sodium salt concentrations (NaCl or Na2SO4), which damage plants through osmotic stress, oxidative stress, ionic stress, nutritional deficiency, and other adverse conditions [1]. According to statistics, more than 60 million hm2 of arable land worldwide is threatened by varying degrees of salinization, accounting for approximately 20% of the world’s total area of irrigated land [2]. Due to improper irrigation practices, poor fertilizer application and industrial pollution, the salinized land area will continue to increase, and it is predicted that more than half of the arable land will be salinized by 2050 [3]. Soil salt stress not only causes direct ionic toxicity to plants, but also indirectly induces osmotic stress and oxidative stress, inhibition of cell division and proliferation, and affects plants at the physiological, biochemical, and molecular levels [4, 5]. These adverse physio-biochemical and molecular conditions simultaneously leads to inhibits plant growth, and its development. Many crops and vegetables, such as Vigna aconitifolia [6], Beta vulgaris [7] and Avena sativa [8], are very sensitive to salt stress. In addition, salt stress adversely affects the growth and development of many woody plants, such as Populus [9], Eucalyptus grandis [10] and Camellia sinensis [11], reducing the yield and quality of plants and causing very large economic losses. Previous studies have shown that plants are most sensitive to salinity during the seed germination and seedling stages; salt stress produces osmotic stress and excessive ion toxicity, causing retarded plant development and inhibiting tissue differentiation and growth [12]. The germination rate, shoot length and root length of seeds at the germination stage are extremely sensitive to salt stress in plants [13]. Salt stress leads to increased Na+ and Cl- accumulation in the plant seed hypocotyl and endosperm, which hinders seed germination, seedling growth, and DNA, RNA and protein synthesis [14, 15]. It has been reported that salt stress delays seed germination in tomato varieties, inhibits cottonseed germination and endogenous melatonin accumulation, reduces the content of α-amylase and β-galactosidase in cotton seeds and decreases the germination rate of Suaeda salsa seeds [16, 17, 18]. Over the years, researchers have explored various approaches, such as operating seed biopriming techniques [19] and nanoparticles [20], to identify salt-tolerant genes, cultivate salt-tolerant varieties and alleviate the impact of salt stress on plants. Currently, due to the high degree of salinization and poor agricultural production conditions in Songnen Plain of China, the popularization of mulberry is expected to transform saline-alkali land into fertile land [21]. Therefore, understanding the molecular and physiological mechanisms of mulberry salt tolerance is necessary for maintaining mulberry productivity.

Morus alba L. is an important perennial woody plant, and its leaves are the main food source for silkworms. Mulberry has high nutritional and economic value and has various uses in ecology, food and medicine [22]. Its high content of active ingredients and high resistance to adversity stresses, such as drought, salinity and flooding, make mulberry an excellent material for researching stress in woody plants [23] and one of the ecological tree species with great potential for afforestation, soil and water conservation, bioremediation of pollutants, land desertification and saline land management [24, 25]. In depth research on genes related to mulberry stress signaling has been conducted. For example, under ultraviolet-B and dark stress, mulberry leaves activate the cell redox and ER quality control systems, enhancing protein synthesis and weakening N-glycan biosynthesis in the ER to resist damage [26]. The activities of superoxide dismutase (SOD), catalase (CAT) and peroxidase (POD) in mulberry can eliminate Cd stress-induced oxidative stress [27]. After abscisic acid (ABA) biosynthesis is activated in mulberry, some drought-responsive genes (such as CBP6A, MGST3, ERF1B, etc.) are also activated [28]. In recent years, mulberry planting has been performed for the ecological restoration of land, which is of great significance for restoring and reconstructing vegetation in saline-alkali land and to improve the ecological environment. Seed sowing propagation is an important agricultural production method for mulberry propagation. Mulberry seeds in the germination stage are extremely sensitive to salt stress, and China contains a large area of saline-alkali land, which urgently needs to be utilized. It is necessary to study the response mechanism of mulberry seeds to salt stress. Proteins are directly involved in the stress response of plants. Studying the response of mulberry to salt stress conditions at the protein level can provide a powerful tool for revealing the physiological mechanism of mulberry stress resistance.

Furthermore, physiological responses in cells are produced by proteins, and proteomics, a global analysis strategy, is the key to studying gene functions [29]. Proteomic analysis can provide information on multiple processes in complex events, such as seed germination and seed filling. Seed germination is the very beginning of plant life and is among the most critical and decisive stages in the plant growth cycle, which determines the formation of adult plants and the final crop yield [30, 31]. Analyzing seed germination through proteomics is mainly carried out by sequencing plants, such as Oryza sativa [32], Brassica napus [33], Populus [34], Jatropha curcas [35], Araucaria [36] and Platanus orientalis [37]. Relatively few studies have focused on the germination of seeds from Morus alba. In recent years, the molecular basis of salt responses and tolerance during seed germination has been investigated using proteomic approaches in some crops and woody plant species, including Triticum aestivum [38], Zea mays [39], Sorghum bicolor [40] and Ulmus pumila [41]. The results showed that the abundance of H2O2 scavenging enzymes decreased under salt stress, and seed priming with ascorbic acid could improve the salt tolerance of durum wheat; MFT (MOTHER OF FT ANDTFL1) was upregulated by the repressor of the GA signaling pathway under salt treatment through inhibition of ABA-positive transcription factor ABI5, which promoted maize seed germination; under salt stress conditions, the germination of salt-sensitive sweet sorghum germplasm was significantly inhibited. With the publication of the mulberry genome sequence, genome-wide transcriptome and proteome analyses can be performed with mulberry [22, 42]. However, previous studies on mulberry have mainly focused on the tolerance of transgenic mulberries to drought stress [28], transcriptomic analysis of mulberry roots under cadmium stress [43] and molecular mechanisms of copper stress in mulberry leaves and identification of quality resources in resistant species [44]. The molecular mechanism for the proteomic response of mulberry to salt stress during seed germination remains unclear. Therefore, this study aimed to identify salt stress-responsive proteins during mulberry seed germination at the physiological and proteomic levels and to provide some new insights to further understand the abiotic stress response of mulberry and mulberry breeding and cultivation.

2. Materials and Methods
2.1 Mulberry Seed Germination under Salt Stress Treatment

The experimental material of this study is “Yuesang No. 11” provided by Botanical Garden of Zhongkai University of Agriculture and Engineering. First, the mature mulberries were harvested; the seeds were washed; mature, full and uniform seeds were selected; and the seeds were stored at 4 °C for subsequent experiments. The seeds were soaked at room temperature for 24 h and the soaking solution was discarded. The seeds were then surface disinfected with 75% ethanol and sodium hypochlorite solution (1% effective chlorine content). 75% ethanol was soaked for 5 min, rinsed with sterile water 3 times [17], sodium hypochlorite solution was soaked for 30 min, rinsed with sterile water 3 times, and the moisture on the surface of the seeds was absorbed with filter paper. The samples were placed in a sterile 90 mm diameter Petri dish covered with three layers of filter paper. The control group (CK) was treated with a proper amount of sterilized water (approximately 12 mL/dish) and 50 and 100 mM NaCl solutions. Each dish was placed with 100 seeds for germination, and biological reproduction was carried out three times. After 14 days, the seedlings were sampled for physiological analysis and protein extraction.

2.2 Physiological Analysis of Mulberry Seed Germination

Mulberry seeds germination began after 2 days of treatment (seeds were considered germinated when radicles appeared), and germinating seeds were observed and counted every 48 hours to calculate the germination percentage. After 14 days, mulberry seedlings from different treatments were collected and rinsed with deionized water. The germination rate and radicle length of mulberry seedlings were determined, and the final germination rate was the number of seeds germinated divided by the total number of seeds used for the test × 100.

In this experiment, SOD, POD, CAT activities as well as H2O2 content and MDA content were detected using biological assay kits A001-1-1, A084-3, A007-1, A064-1 and A003-1-1 respectively (Nanjing Jiancheng Bioengineering Institute, Nanjing,China) according to the manufacturer’s instructions. SOD activity was determined by measuring the absorbance at 550 nm using the xanthine oxidase method. In addition, the POD activity was determined by gauging the absorbance at 420 nm using guaiacol and H2O2 as substrates, and CAT activity was determined by measuring the absorbance at 405 nm using H2O2 as a substrate. The malondialdehyde content was determined through a thiobarbituric acid reactive substance (TBARS) assay. Three biological replicates were detected.

2.3 Protein Extraction, Peptide Enzymatic Hydrolysis and Peptide Desalination

Mulberry seedling samples that were stored at –80 °C were ground into a fine powder, and four volumes of phenol extraction buffer (1% protease inhibitor and 2 mmol/L EDTA, 10 mmol/L dithiothreitol) were added for ultrasonic cracking. Add the same amount of Tris balanced phenol, centrifuge at 5500 × g at 4 ℃ for 10 min, add 5 times of 0.1 mol/L ammonium acetate/methanol for precipitation for 12 h, and wash with methanol and acetone respectively. The extracted protein was quantified by bicinchoninic acid disodium (BCA) method.

The extracted proteins were subjected to trypsin (317107 Promega) hydrolysis (protein/trypsin, 50/1) using the method of filter-aided proteome preparation (FASP) described by Wisniewski et al. [45]. Add the total protein solution into a protein concentrator (10K MWCO, 0.5 mL Pierce™ Protein Concentrator, Thermo Scientific, Waltham, MA, USA), dilute the sample with 8 mol/L urea, and after reduction and alkylation, digest it with trypsin at 37 °C overnight. After digestion, centrifuge at 10,000 × g for 15 min, collect the filtrate, and then filter and wash it twice with 50 μL NH4HCO3 solution. Then, 10% trifluoroacetic acid (TFA) was added, and when the final concentration was 0.4%, the reaction was terminated. The proteolytic product was steamed in a rotary evaporator until the volume was less than 50 μL, and then buffer A (64% acetonitrile, 0.2% TFA) was added to about 200 μL for later use. The peptide desalting was carried out by PierceTM C-18 desalting column (PierceTM C-18 Spin columns, Thermo Scientific, USA). After the desalting column was activated by methanol, the enzymolysis product was added to the desalting column, centrifuged at 1000 × g for 30 s, and the filtered sample was returned to the desalting column and centrifuged again. The filtrates were combined, concentrated and dried.

2.4 TMT Labeling and Peptide Fractionation

100 μg of peptide was extracted from each sample for labeling according to TMT labeling kit instructions (Thermo Fisher Scientific, USA). The mulberry seedlings from the three groups under different salt stresses are labeled. Each group included three biological replicates. The mulberry seedling peptides from the control treatment were labeled with mass 126, 127N and 127C isobaric TMT tags, while the seedling peptides from the 50 mmol salt treatment were labeled with mass 128N, 128C and 129N isobaric TMT tags, and the 100 mmol salt treatment was labeled with mass 129C, 130N and 130C isobaric TMT tags.

Each set of labeled peptide fragments was uniformly mixed and fractionated using the Reverse-Phase High pH Peptide Fractionation Kit (Thermo Scientific, USA). The TMT-labeled digested sample was fractionated into 10 fractions using a gradient of acetonitrile elution using the Pierce™ High pH Reverse-Phase Peptide Fractionation Kit (Thermo scientific, USA) according to the manufacturer’s instructions.

2.5 LC-MS/MS Analysis

Each fraction was injected and analyzed by Nano LC-MS/MS. The peptide mixture was loaded onto a reversed-phase trap column (Thermo Scientific Acclaim PepMap100, 100 μm × 2 cm, nanoViper C18) that was connected to the C18-reversed-phase analytical column (Thermo Scientific Easy Column, 10 cm long, 75 μm inner diameter, 3 μm resin) in buffer A (0.1% Formic acid) and separated with a linear gradient of buffer B (84% acetonitrile and 0.1% Formic acid) at a flow rate of 300 nL/min. After column separation, all samples were analyzed by mass spectrometry using a Q-Exactive mass spectrometer (Proxeon Biosystems, Thermo Fisher Scientific, USA).

2.6 Protein Identification and Quantification

Raw data for the Q-Exactive spectrometry analysis were obtained as RAW files. Identification and quantitative analyses were performed by searching the database using Mascot version 2.2 and Proteome Discoverer version 1.4. The UniProt_Lycopersicon_esculentum Mulberry_35921_20211214.fasta (35921 sequences, downloaded on 14 December 2021) database was searched. Trypsin was selected as a random cleavage enzyme, and the parameter of max missed cleavages was set to 2. Carbamidomethyl (C), TMT 9 plex (N-term), and TMT 9 plex (K) was selected as fixed modifications. The variable modifications were set to oxidation (M), TMT 9 plex (Y). The peptide mass tolerance was set to ±20 ppm and the fragment mass tolerance was 0.1 Da. The database schema was used as a decoy for calculating the false discovery rate (FDR). The screening criteria for plausible peptides were FDR 0.01. DAPs were screened using a power analysis Student’s t test p < 0.05 (n = 3) and the criterion of a fold change greater than 1.2 (thresholds of fold changes 1.2 for up-regulation proteins and the thresholds of fold changes 0.83 for down-regulation proteins).

2.7 Bioinformatics Analysis

All annotation and DAP classification of Gene Ontology (GO) features were performed using the Blast2GO (https://www.blast2go.com/) program. The Automatic Annotation Server was used to perform Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway annotations [46]. The significance of protein enrichment of GO functional annotations and KEGG pathway annotations was evaluated by Fisher’s exact test.

2.8 PRM Analysis

The peptide information suitable for PRM analysis was imported into Xcalibur software (Thermo scientific, USA) to establish the PRM method. An equivalent amount of digested peptide of approximately 1 μg from each sample was removed and mixed with an equivalent amount of standard peptide of 20 fmol (PRTC:ELGQSGVDTYLQTK) for detection. PRM analyses were performed on a Q-Exactive Plus mass spectrometer equipped with an Easy nLC-1200 system (Thermo Fisher Scientific, Bremen, Germany). The liquid chromatographic column was equilibrated with 95% solution A (0.1% formic acid). 2 μL of each sample was injected into the column at a flow rate of 300 nL/min for gradient separation. The gradient of the liquid phase separation was as follows: linear gradient of 5–10% solution B (acetonitrile solution of 0.1% formic acid) for 0–2 min, followed by 10–30% solution B for 2–45 min, followed by 30–100% solution B for 45–55 min; finally, solution B was kept at 100%.

PRM analysis of samples separated by HPLC was performed for 60 min using a Q-Exactive HF mass spectrometer (Thermo Scientific, USA). The analytical column was a homemade tip column (75 μm × 200 mm, 3 μm-C18). Finally, Skyline software 3.5.0 (Thermo Scientific, USA) was used to statistically analyze the PRM data.

2.9 Statistical Analysis

Analysis and examination of the obtained experimental data were performed using Excel (Microsoft office professional plus 2010; Microsoft corporation, Redmond, WA, USA) and SPSS software 17.0 (SPSS Inc., Chicago, IL, USA). All data are the mean ± standard deviation (SD) of three repetitions, and the differences among different treatments were compared by one-way ANOVA and LSD. The level of significance was set at p < 0.05, and a very significant difference was set at p < 0.01 for all tests.

3. Results
3.1 Physiological Responses of Mulberry Seeds to Salt Stress During Germination

The germination percentage reflects the resistance of the seed germination process to abiotic stress. As shown in Fig. 1A, the germination rate of mulberry seeds increased significantly with increasing incubation time. Different concentrations of the same treatment days had different effects on the germination of mulberry seeds. With 0 mmol/L salt concentration, the germination rate continued to increase and reached the highest rate of 94.33% at day 14. With the 50 mmol/L salt treatment, the germination rate of seeds increased slowly at 4 days, increased rapidly from 4 to 8 days, and then increased slowly again from 8 to 14 days, and the germination rate of seeds at day 14 was 50%, which decreased by 44.33% compared with that of the control. With the 100 mmol/L salt treatment, the seed germination rate increased slowly from 0 to 8 days, but both were slower than the control and 50 mM NaCl stress treatments, and the seed germination rate on day 14 was 36.67%, which was decreased by 57.66% compared to that of the control. The germination rate of mulberry seeds was significantly lower than the germination rate of the control on the same day when the salt concentration was 50 mmol/L and 100 mmol/L. The growth of embryonic roots is crucial for plants to properly develop into seedlings out of the soil. As shown in Fig. 1B, the embryo root length increased with increasing embryo root growth days at the same concentration. Under 0 mmol/L salt, the embryo root length of mulberry seeds increased significantly, and the longest radicle length was 32.20 mm on the 14th day. With the 50 mmol/L salt stress treatment, the radicle length increased slowly from 0 to 6 days and rapidly from 6 to 14 days and basically stopped increasing after day 14. After 14 days, the length basically stopped increasing. At a salt concentration of 100 mmol/L, mulberry embryo roots basically did not grow. The length of the mulberry radicles varied at the same time for different concentrations. Compared to the control, both the 50 mmol/L and 100 mmol/L salt treatments significantly reduced the radicle length on the same day.

Fig. 1.

Detection of the physiological parameter of mulberry seeds at different salt treatment stages. (A) 14-day germination rate of mulberry seeds. (B) Radicle length of mulberry seeds in 14 days. (C) Superoxide dismutase (SOD) activity. (D) Peroxidase (POD) activity. (E) Catalase (CAT) activity. (F) H2O2 content. (G) Malondialdehyde (MDA) content. Note: Three replicates were performed. The values are presented as the means ± SDs. Significant differences are expressed by different lowercase letters (p < 0.05), and very significant differences are expressed by different capital letters (p < 0.01).

In response to salt stress, antioxidant enzymes (GST, POD, SOD, DHAR, APX and CAT) were differentially accumulated during seed germination of mulberry. Therefore, we studied the physiological indicators of salt stress, including the determination of SOD, POD and CAT activities and the content of H2O2 and MDA. The results are shown in Fig. 1. Under different salt treatment conditions, compared with the control, SOD, POD and CAT in mulberry seeds under salt stress exhibited significantly increased activities (Fig. 1C–E), SOD activity increased by 1.26-fold and 1.13-fold, POD activity increased by 4.17-fold and 2.72-fold, and CAT activity increased by 1.55- and 1.64-fold, respectively. The effect of different levels of salt stress on H2O2 content was not significant and did not differ significantly among treatments (Fig. 1F). However, the MDA content in mulberry seeds decreased significantly with increasing salt concentration (Fig. 1G), thus alleviating the damage caused by salt to the cell membrane.

3.2 Identification of the Proteins in Mulberry Seeds under Salt Stress during Germination

TMT-based quantitative proteomics was used to analyze the dynamic changes in proteins in mulberry seeds. A total of 76,544 unique peptides were detected, and 7717 proteins were identified based on TMT data after removing duplicate proteins. We counted the molecular weights of the identified proteins (Fig. 2A), and the results showed that there were 811 (10.51%) of 10.0–20.0 kD, 1123 (14.55%) of 20.0–30.0 kD, 1193 (15.46%) of 30.0–40.0 kD, and 1067 (13.83%) of 40.0–50.0 kD, 959 (12.43%) were 50.0–60.0 kD, and 90 (1.17%) were larger than 200.0 kD. The numbers of proteins with 1 peptide, 2–5 peptides, 6–10 peptides and more than 11 unique peptides were 2224, 3207, 1351 and 935, respectively. In addition, the number of peptides corresponding to almost 70% of the identified proteins was 2, which indicated that the MS data could be further analyzed (Fig. 2B). Proteins with peptide sequence coverage changes of 40%–100%, 30%–40%, 20%–30%, 10%–20%, and 10% accounted for 9.33%, 7.62%, 11.77%, 22.68% and 48.61%, respectively (Fig. 2C). The peptide sequences of most proteins have good coverage, and the results show that the protein data are accurate and reliable.

Fig. 2.

The overall distribution of mulberry seed proteins identified by TMT. (A) Protein molecular weight distribution. (B) Peptide count distribution. (C) Distribution of peptide sequence coverage.

3.3 Identification and Analysis of DAPs

In this study, with the mulberry seedling protein sample (CK group) as the control, the quantitative intensity of mass spectrometry of the mulberry seedling protein sample in the 50 mmol/L NaCl and 100 mmol/L NaCl salt-treated groups was compared with the protein sample in the CK group. The mean value of three biological replicates was used as the fold difference of protein expression. A fold-change cut off of 1.2 and p < 0.05 for upregulated proteins and 0.83 and p < 0.05 for downregulated proteins was applied to identify the DAPs between salt-treated and control plants (Fig. 3). A total of 143 DAPs were identified in salt-treated (50 mmol/L NaCl) and control (CK) mulberry seeds, of which 61 DAPs were upregulated and 82 were downregulated (Fig. 3A). Then, 540 DAPs were identified in salt-treated (100 mmol/L NaCl) and control (CK) mulberry seeds, of which 222 DAPs were upregulated and 318 DAPs were downregulated (Fig. 3B). It is obvious that the DAP of the 100 mmol/L NaCl treatment is more than three times higher than that of the 50 mmol/L NaCl treatment, indicating that a high salt concentration has a greater impact on seed germination.

Fig. 3.

Analysis of DAPs in mulberry seeds. (A) The volcano plots of DAPs in T50 vs. CK. (B) The volcano plots of DAPs in T100 vs. CK.

A total of 113 DAPs were screened in both salt treatments, of which 43 DAPs were upregulated and 70 DAPs were downregulated. By analyzing the hierarchical clustering analysis of DAP in different treatments, every three replicates of CK and salt-treated samples were clustered together separately with high reproducibility. Furthermore, the protein expression abundance and expression levels in mulberry seeds differed significantly between the CK and salt-treated samples (Fig. 4). At the same time, the functional analysis of the differentially abundant proteins screened from the two salt-stressed mulberry seeds showed that the differentially expressed proteins mainly belonged to reactive oxygen species (ROS) scavenging, stress defense, energy production, carbohydrate metabolism, transcription and translation, growth and development, signal transduction, substance transport, protein synthesis, folding and degradation categories (Table 1). Statistics of differential proteins include molecular weight, fold difference and p value.

Fig. 4.

Hierarchical clustering analysis of the DAPs in different treatments.

Table 1.Partial DAPs of mulberry seeds induced by salt stress.
Function classification Accession Description MW [kDa] Fold Change p-value
T50/CK T100/CK
Reactive oxygen scavenging and stress defense W9QU60 Peroxidase 34.19 1.379 1.678 0.024
W9RZP5 Superoxide dismutase 26.41 1.276 1.405 0.036
A0A1D8KVU3 Dihydroflavonol reductase 1 38.30 0.793 0.775 0.008
W9QH65 Glutathione peroxidase 26.50 1.268 1.361 0.024
W9T1J8 SHSP domain-containing protein 24.41 1.206 1.323 0.003
W9RN30 Jacalin-type lectin domain-containing protein 15.71 1.321 1.740 0.017
W9RJC9 Putative disease resistance protein RGA4 128.22 0.833 0.800 0.030
W9SE40 Sulfiredoxin 13.76 0.827 0.773 0.002
W9RK87 Spermine synthase 43.61 1.258 1.284 0.024
W9SCX0 Lycopene beta cyclase 56.26 0.768 0.806 0.029
W9QK12 Lycopene epsilon cyclase 59.98 0.827 0.725 0.026
W9RBR2 Allene oxide synthase 57.48 0.649 0.538 0.042
W9RAN9 Anthocyanin 5-aromatic acyltransferase 52.24 0.795 0.685 0.028
Energy production W9S8G4 Lipase 45.45 1.458 1.866 0.030
W9QX55 Lipoxygenase 86.97 0.821 0.758 0.009
Q09X21 Photosystem II CP43 reaction center protein 51.79 0.827 0.783 0.048
Q09X17 Photosystem I P700 chlorophyll a apoprotein A1 83.06 0.807 0.724 0.029
W9QPP1 Oxygen-evolving enhancer protein 3-2 24.70 0.628 0.545 0.019
W9R2T4 Glucose-1-phosphate adenylyltransferase 59.46 1.249 1.299 0.020
W9RFS1 Short-chain type 64.75 1.397 1.413 0.011
W9SKC6 Flavin-containing monooxygenase 47.43 0.736 0.543 0.034
W9R035 Frataxin 22.18 1.292 1.279 0.035
W9R2T4 Glucose-1-phosphate adenylyltransferase 59.46 1.249 1.299 0.020
W9RCL2 Lipase_3 domain-containing protein 58.76 1.288 1.437 0.036
W9QJ14 Sphingoid base hydroxylase 2 30.37 1.290 1.458 0.015
W9RFT7 Alpha-carbonic anhydrase domain-containing protein 30.62 0.718 0.644 0.023
Carbohydrate metabolism W9QF53 Cellulose synthase 120.43 0.829 0.830 0.024
W9RH68 Endoglucanase 55.19 0.802 0.741 0.019
W9SJJ0 Alpha-L-fucosidase 1 58.36 1.218 1.277 0.003
W9S5L8 Fructose-1,6-bisphosphatase 44.87 0.810 0.744 0.029
W9R8E3 Putative polygalacturonase 63.38 1.442 1.426 0.033
Transcription and translation W9R0R0 Histone 53.57 0.721 0.748 0.013
W9SN28 Histone H1 23.17 0.687 0.628 0.023
W9RWW0 Transcriptional corepressor SEUSS 108.92 0.785 0.583 0.023
W9SG17 BHLH domain-containing protein 39.45 0.719 0.658 0.021
W9QUN0 Transcription initiation factor IIF subunit alpha 52.71 0.817 0.801 0.025
W9RKQ5 HMG1/2-like protein 15.88 0.736 0.762 0.002
W9RJ53 High mobility group B protein 2 16.04 0.755 0.735 0.001
W9QWT1 Peptidylprolyl isomerase 24.58 0.794 0.731 0.025
W9QWT8 Glyceraldehyde-3-phosphate dehydrogenase B 147.81 0.795 0.713 0.034
W9STM4 Btz domain-containing protein 73.82 0.771 0.695 0.016
W9SN10 AT-hook motif nuclear-localized protein 31.51 0.631 0.621 0.045
W9QMK8 TCP domain-containing protein 38.21 0.752 0.679 0.015
W9RVK1 SWI/SNF chromatin-remodeling complex subunit snf22 246.07 0.803 0.750 0.008
W9SBM7 U3 small nucleolar RNA-associated protein 15-like protein 58.89 0.788 0.725 0.028
Growth and development W9R8E1 GDSL esterase/lipase 38.53 1.695 1.822 0.010
W9RZ71 Aspartic proteinase nepenthesin-2 63.51 1.457 1.571 0.030
W9RMA6 Beta-galactosidase 88.18 1.644 1.785 0.022
W9RY25 Subtilisin-like protease 87.92 1.310 1.359 0.011
W9QYC8 GDSL esterase/lipase 41.61 1.216 1.203 0.023
W9SHN8 GDSL esterase/lipase 40.75 1.409 1.446 0.011
W9RD56 GDSL esterase/lipase 6 39.63 1.432 1.543 0.002
W9RG64 Purple acid phosphatase 50.47 1.344 1.368 0.039
W9QMY5 Purple acid phosphatase 50.11 1.345 1.382 0.025
W9S1A8 Pectinesterase 62.59 0.813 0.784 0.013
W9SKF4 Caffeic acid 3-O-methyltransferase 29.05 0.657 0.647 0.016
Signal transduction W9RBX0 Putative receptor-like protein kinase 91.76 1.333 1.388 0.007
W9RZ35 Putative L-type lectin-domain containing receptor kinase S.5 76.53 1.204 1.248 0.037
W9QVJ6 Protein kinase domain-containing protein 31.02 0.796 0.702 0.002
W9RA31 Systemin receptor 127.65 0.806 0.811 0.046
W9R8I4 G-type lectin S-receptor-like serine/threonine-protein kinase 134.05 0.794 0.800 0.048
Transport proteins W9R5Q9 Putative cyclic nucleotide-gated ion channel 8 85.60 0.790 0.783 0.037
W9SAJ0 ABC transporter G family member 28 112.05 1.247 1.334 0.049
W9RS36 ABC transporter G family member 8 66.42 0.790 0.746 0.048
W9RGK8 Nitrate transporter 1.1 29.38 0.755 0.680 0.023
W9SDQ8 Nitrate transporter 1.3 66.41 0.768 0.639 0.020
W9QFD5 Protein DETOXIFICATIO 59.23 1.213 1.301 0.031
W9RWS8 Transmembrane ascorbate ferrireductase 1 25.28 1.232 1.226 0.047
W9QQQ4 Putative sulfate transporter 3.3 74.52 0.806 0.789 0.046
W9QN95 Protein translocase subunit SecA 118.21 0.829 0.750 0.016
W9S950 CMP-sialic acid transporter 2 44.79 1.355 1.335 0.006
W9SK42 t-SNARE coiled-coil homology domain-containing protein 56.65 1.394 1.414 0.017
Protein synthesis, folding and degradation W9SIU2 30S ribosomal protein S21 20.85 0.824 0.735 0.012
W9QWL4 50S ribosomal protein L34 16.60 0.680 0.683 0.030
W9SXH8 Carboxypeptidase 58.05 1.335 1.451 0.012
W9SPH4 Ribonuclease 3 23.25 2.357 2.638 0.015
3.4 Functional Annotation of DAPs

In the GO enrichment analysis of DAP in T50 vs. CK, DAP was significantly enriched into 66 functional GO terms, of which 40 belonged to biological processes (BPs), 25 to molecular functions (MFs), and 1 to cellular components (CCs). Among the top 20 enriched GO terms (Fig. 5A), carotenoid biosynthesis, tetraterpenoid biosynthesis, tetraterpene metabolism, carotenoid metabolism, lipid metabolism, and transcription elongation from the RNA polymerase II promoter were the most significantly enriched GO terms in the BP category. In the MF category, hydrolase activity (acting on ester bonds), polygalacturonase activity, secondary activity transmembrane transporter activity, nuclease activity and hydrolase activity were the most significantly enriched GO terms. Moreover, the most important term for GOs in the CC category is the extracellular region. GO enrichment analysis of the DAPs identified in T100 vs. CK resulted in a significant enrichment of 122 functional GO terms, of which 50 belonged to BPs, 46 to MFs, and 26 to CCs. In the top 20 enriched GO terms (Fig. 5B), cellular glucan metabolism, glucan metabolism, cellular polysaccharide metabolism, carotenoid biosynthesis, and abscisic acid-activated signaling pathway were the most significantly enriched GO terms in the BP category. Protein dimerization activity, protein heterodimerization activity, DNA binding, isoprenoid binding, alcohol binding, and abscisic acid binding were the most significantly enriched GO terms in the MF category. The most significantly enriched GO terms in the CC category were protein‒DNA DNA packaging complex, DNA packaging DNA packaging complex, nucleosome, chromatin and chloroplast part.

Fig. 5.

Gene ontology annotation of all differentially abundant proteins identified in mulberry seeds. The top 20 significantly enriched GO terms in (A) T50 vs. CK and (B) T100 vs. CK.

3.5 KEGG Pathway Enrichment Analysis of DAPs

Furthermore, DAPs were also annotated according to the KEGG pathway. KEGG enrichment analysis of T50 vs. CK DAPs showed that 55 pathways were enriched, including photosynthesis, chlorocyclohexane and chlorobenzene degradation, fluorobenzoate degradation, toluene degradation, and prodigiosin biosynthesis. Notably, only one pathway, carotenoid biosynthesis, was significantly enriched among these enriched pathways (Fig. 6A). In the KEGG pathway enrichment analysis of DAPs of T100 vs. CK, a total of 147 pathways were enriched, of which 7 were significantly enriched; these pathways mainly related to photosynthesis, systemic lupus erythematosus, carotenoid biosynthesis, alcoholism, mineral absorption, plant hormone signal transduction and necroptosis (Fig. 6B).

Fig. 6.

KEGG pathway enrichment analysis of DAP in mulberry seeds. The significantly enriched KEGG pathways in DAPs from (A) T50 vs. CK, (B) T100 vs. CK. The abscissa is the number of proteins involved in metabolic pathway; the ordinate is metabolic pathway. The color gradient represents the p-value below 0.05. The darker the color (closer to red), the smaller the p-value, and the more significant is the enriched KEGG pathways. The number on the bar graph indicates the rich factor, referring to the percentage of differential proteins that are involved in this pathway.

3.6 PRM-Based Validation of TMT Data

To validate the TMT proteomic data, we selected five differentially abundant proteins for PRM quantification (Table 2). PRM identified five candidate differentially abundant proteins, including cellulose synthase, glyceraldehyde-3-phosphate dehydrogenase B, superoxide dismutase, glutathione peroxidase, and photosystem II CP43 reaction center protein. The PRM results exhibited consistency in the differential expression patterns of the selected proteins with the TMT data. In 50 mmol/L NaCl solution, the expression levels of glutathione peroxidase and superoxide dismutase increased significantly in PRM and TMT analysis, while the expression levels of cellulose synthase, glyceraldehyde 3-phosphate dehydrogenase B and photosystem II CP43 reaction center decreased significantly in PRM and TMT analysis. In the 100 mmol/L NaCl solution, the expression of cellulose synthase, glyceraldehyde-3-phosphate dehydrogenase B and photosystem II CP43 reaction center protein was downregulated in the PRM data and did not change according to the TMT results. Overall, the consistency of PRM and TMT results validates the credibility of TMT in proteomic analysis.

Table 2.Comparison of the quantitative results between TMT and PRM.
Accession Gene name Description TMT PRM
T50/CK T100/CK T50/CK T100/CK
W9QF53 L484_012515 Cellulose synthase 0.829 0.830 0.835 0.455
W9QWT8 L484_024833 Glyceraldehyde-3-phosphate dehydrogenase B 0.795 0.713 0.682 0.175
W9RZP5 L484_013984 Superoxide dismutase 1.276 1.405 1.363 1.868
W9QH65 L484_024293 Glutathione peroxidase 1.268 1.361 1.016 1.508
Q09X21 psbC Photosystem II CP43 reaction center protein 0.827 0.783 0.648 0.196
4. Discussion

Seed germination and seedling growth are among the most sensitive stages in plant life, representing the first contact with water and soil, whereas salt stress affects seed germination and seedling growth [47]. In the current study, we identified salt stress-responsive proteins during mulberry seed germination at the proteomic level by the TMT method. A total of 143 DAPs were identified when treated with 50 mmol/L NaCl solution, of which 61 were upregulated and 82 were downregulated, and 540 DAPs were identified when treated with 100 mmol/L NaCl solution, of which 222 proteins were upregulated and 318 were downregulated. A total of 113 DAPs were screened in the two salt treatments, of which 43 proteins were upregulated and 70 proteins were downregulated. Functional analysis of DAPs screened from mulberry seeds treated with salt stress showed that the differentially expressed proteins mainly belonged to ROS scavenging, stress defense, energy production, carbohydrate metabolism, transcription and translation, growth and development, signal transduction, substance transport, protein synthesis, folding and degradation. In the following section, we will discuss the differential proteins that fall into these functional categories.

4.1 Salt Stress Induces Changes in ROS Scavenging and Stress Defense

During seed germination, respiration produces large amounts of ROS, and salt stress accelerates the accumulation of ROS, resulting in an imbalance between ROS production and detoxification. In response to this oxidative stress, an effective antioxidant system can scavenge excess ROS [48]. ROS can also contribute to the endosperm weakening during germination as a result of the cell wall loosening, but uncontrolled ROS production can lead to cell damage, resulting in seed deterioration, retarded germination and early seedling development, and the production of some antioxidants, such as SOD, POD, CAT, glutathione-related enzymes and dehydrogenases, to reduce ROS-induced cell damage in plants [49]. In this study, physiological data showed that the activities of SOD, POD, and CAT in mulberry seedling samples continued to increase under NaCl-induced salt stress. It has been reported that the activities of antioxidant enzymes, such as SOD, POD, CAT, and glutathione-related enzymes, generally increase in plants under oxidative stress conditions, resulting in enhanced tolerance to stress [50]. Proteomics research in this study showed that antioxidant-related proteins, such as peroxidase (W9QU60), superoxide dismutase (W9RZP5) and glutathione peroxidase (W9QH65), were significantly upregulated during seed germination under salt stress. Moreover, the changes in peroxidase, superoxide dismutase and glutathione peroxidase were higher under 100 mM NaCl treatment than under salt stress of 50 mM NaCl. These results may indicate that the plants suffer from relatively severe oxidative stress and need to initiate antioxidant enzyme systems to scavenge excessive ROS. Under the induction of 50 mM NaCl and 100 mM NaCl, the MDA content decreased from 1.39-fold to 3.81-fold compared with the control. MDA is mainly produced by polyunsaturated fatty acid peroxides and is an important indicator of cell membrane lipid peroxidation and plasma membrane damage caused by biotic and abiotic stresses [51]. In general, the lower the MDA content is, the less damage the plasma membrane incurs. The results of this study showed that the damage to the plasma membrane in mulberry seeds was relatively mild, indicating that mulberry seedlings exhibit significant salt tolerance under salt stress.

Salt stress can induce the production of ROS and damage plant cells, and to regulate ROS levels, many defense proteins are induced during plant responses to biotic or abiotic stresses [52]. In this study, in response to changes in ROS scavenging and stress defense induced by salt stress, 15 DAPs (7 upregulated and 8 downregulated) were screened under salt stress (50 mM NaCl), while under 100 mM NaCl salt stress, 40 DAPs were screened (24 upregulated and 16 downregulated); a total of 13 DAPs were screened in both salt treatments (5 upregulated and 8 downregulated). Heat shock proteins (HSPs), also known as heat stress proteins, play an important role in regulating the disease resistance and stress response of plants [53]. sHSP is the plant heat shock protein with the smallest molecular weight, and this protein mainly protects other proteins from various stresses in plants through molecular chaperones. A large number of studies have shown that sHSP members can respond to various external stresses [54]. Overexpression of the rice sHSP member Os HSP16.9 can improve the salt resistance and drought resistance of rice plants [55]. Overexpression of the sHSP member Ms HSP17.7 in alfalfa can improve the defense ability of alfalfa plants against heat, drought, and antioxidant stresses [56]. In this study, the expression of sHSP domain-containing protein (W9T1J8) was upregulated, and the fold change ratio increased with increasing salt concentration, which indicated that the upregulated expression of sHSP domain-containing protein was a defense mechanism of mulberry seed germination in response to salt stress. Furthermore, KEGG enrichment analysis of DAPs showed that seed germination under salt stress was mainly enriched in carotenoid biosynthesis. Carotenoids are natural pigments that exhibit antioxidant activity, play an important role in the production of plant hormones and can effectively protect plant cells [57]. Lycopene cyclization is an important part of the carotenoid biosynthetic pathway in plants and is catalyzed by both lycopene β-cyclase (LCY-β) and lycopene ε-cyclase (LCY-ε) [58]. In this study, proteins related to carotenoid biosynthesis, such as zeta-carotene desaturase (W9QKC9), lycopene ε cyclase (W9QK12) and lycopene β cyclase (W9SCX0), were significantly downregulated. Studies have shown that downregulated IbLCY-ε increases the synthesis of carotenoids in the nonembryonic callus of transgenic sweet potato and enhances the tolerance of sweet potato to environmental stress and the ability to grow on marginal land [59]. The above results further indicate that salt stress induces changes in DAP that were related to ROS scavenging and stress defense; thus, the tolerance and antioxidant capacity of mulberry seeds to salt stress were maintained during germination.

4.2 Salt Stress Induces Changes in Energy Production and Carbohydrate Metabolism

In plants, glycolysis and the tricarboxylic acid (TCA) cycle are the main features of carbohydrate and energy metabolism, which not only satisfy energy demands but also produce many essential cofactors and other substrates of metabolism [60]. In this study, to examine changes in energy production and carbohydrate metabolism induced by salt stress, 26 DAPs (9 upregulated and 17 downregulated) were screened under a salt stress of 50 mM NaCl, 77 DAPs (31 upregulated and 46 downregulated) were screened under a 100 mM NaCl salt stress, and a total of 17 DAPs (8 upregulated and 9 downregulated) were screened under both salt treatments. Glyceraldehyde 3-phosphate dehydrogenase catalyzes the conversion of 3-phosphoglyceraldehyde to 1,3-diphosphoglyceric acid and plays an important role in sugar metabolic pathways in plants [61]. Phosphoglycerate translocase plays an important role in glycolysis by catalyzing the conversion of phosphate from 3-phosphoglycerate to a second carbon to form 2-phosphoglycerate. In this study, the proteomic analysis results showed that related proteins involved in glycolysis and the tricarboxylic acid cycle, including short-chain type (W9RFS1), 2,3-diphosphoglycerate nondependent phosphoglycerate translocase (W9SCE4), 3-phosphoglyceraldehyde dehydrogenase (W9SF99), and phosphoenolpyruvate carboxylase 2 (W9RNP2), were significantly upregulated, while fructose diphosphate aldolase (W9RMK1) was significantly downregulated, indicating that when most identified glycolytic enzymes exhibited increased expression, the production of more NADH was promoted during mulberry seed germination under salt stress; then, the oxidative phosphorylation process by mitochondria began for ATP synthesis, providing the necessary energy for normal cell metabolism under salt stress and playing an important role in the process of improving salt tolerance of mulberry. Moreover, seeds require energy for germination and growth, and lipids are the most efficient form of energy storage in plant seeds [62]. Plant lipases are responsible for the hydrolysis of acylglycerides during lipid processing and can hydrolyze the oil stored in the seeds to provide energy and protection [63]. The expression of lipase (W9S8G4) was significantly upregulated in this study, indicating that this lipase may have enhanced the salt tolerance of mulberry seeds by providing energy and thus promoting germination. It has been found that the overexpression of Arabidopsis lithium-tolerant lipase increased salt tolerance in transgenic Arabidopsis plants, allowed seed to germinate in the presence of NaCl, and stimulated nutritional growth, flowering, and fruit set in the presence of NaCl [64]. Photosynthesis is a fundamental process that promotes the nutritional growth and development of plants, providing the necessary energy for seed germination and early seedling establishment [65]. It has been shown that during seed development, photosynthesis plays an important role in seed maturation; furthermore, inhibition of photosynthesis in seeds results in delayed germination and reduced seed storability, as well as reduced seedling vigor [66]. In this study, proteins related to photosynthesis cytochrome f (Q09X04), chloroplast heme oxygenase (A0A076K090) and oxygen precipitation enhancing protein 3-2 (W9QPP1) were significantly downregulated, indicating that photosynthesis-related proteins may be inhibited by salt stress during seed germination.

In addition, carbohydrate metabolism regulates the synthesis and conversion of sugars, and environmental stresses can disrupt carbohydrate metabolism in plant species [67]. It has been reported that glycoside hydrolase is among the key enzymes in carbohydrate metabolism, which can affect cellulose biosynthesis, such as β-D-xylosidase and xyloglucan transglycosidase/hydrolase, which is involved in sugar biosynthesis and remodeling, and glycoside hydorlase has been identified in the seed development process of many species [68]. Altering the expression levels of glycoside hydrolase genes can regulate cellulose levels and thus biomass production [69]. Glycoside hydrolase genes are upregulated in rice in response to biotic and abiotic stress treatments, indicating their role in stress adaptation [70]. Various proteins related to carbohydrate metabolism in this study, including the putative xyloglucan endoglucosylase/hydrolase protein 32 (W9SJJ4) and β-D-xylosidase 1 (W9R6G8), were significantly upregulated, indicating that these proteins have an important role in adaptation to salt stress during mulberry seed germination.

4.3 Salt Stress Induces Changes in Transcription, Translation and Growth and Development

In this study, the changes in transcription, translation and growth and development caused by salt stress were examined, and 32 DAPs were screened under 50 mM NaCl stress treatment (11 upregulated and 21 downregulated), 116 DAPs (50 upregulated and 66 downregulated) were screened under 100 mM NaCl salt stress treatment, and a total of 22 DAPs were screened in both salt treatments (10 upregulated and 12 downregulated). Seed germination is a complex trait influenced by many genetic, endogenous and environmental factors and is driven by many cellular processes, including transcription and translation [71]. Proteins involved in the regulation of transcriptional and translational activities are important because oxidative stress induced by high NaCl concentrations may affect protein integrity, indicating that the production of new proteins is essential to maintain plant growth [72]. In this study, we found that various transcription and translation factor-related proteins were significantly upregulated, especially under 100 mM NaCl stress, including transcription factor VIP1 (W9RA57), nuclear transcription factor Y subunit C-9 (W9QV28), translation initiation factor IF-2 (W9RQB9), and eukaryotic translation initiation factor 5A (W9S2P4). This indicates that salt stress may affect the process of initiating protein translation during seed germination in Morus alba. It has been shown that under stress conditions, epigenetic changes (i.e., DNA methylation and histone modifications can regulate the expression of important genes associated with stress tolerance and can activate the transcription of genes that respond to abiotic stresses [73]. Histones can be involved in promoting the unwinding, replication and transcription of DNA as well as stabilizing chromatin structure under salinity [60]. It has been reported that SRT-like histone deacetylases in Arabidopsis inhibit transcription, thereby reducing excessive H2O2 accumulation and DNA damage under salt stress and enhancing salt tolerance during seed germination in Arabidopsis [74]. High mobility group proteins (HMG) are widely involved in gene expression and regulation, including DNA replication, transcription, recombination, and DNA repair processes [75]. In the present study, histone (W9R0R0), histone H1 (W9SN28) and high mobility group B protein 2 (W9RJ53) were significantly downregulated, indicating that these proteins may inhibit transcription for germination of mulberry seeds, resulting in better adaptation to salt stress.

Studies have shown that GDSL esterase/lipase is widely distributed in plants, is mainly involved in the regulation of growth and development, and plays an important role in seed germination [76, 77]. It has been reported that GDSL esterase/lipase is overexpressed in transgenic plants [78]. The overexpression of Arabidopsis LTL1, a salt-inducible gene encoding a GDSL motif lipase, also improved salt tolerance in yeast and transgenic plants [79]. The expression levels of GDSL esterase/lipase (W9R8E1) and GDSL esterase/lipase 6 (W9RD56) were significantly upregulated in this study, indicating that these proteins may promote the tolerance of mulberry seed germination to salt stress. Aspartic proteinases (APs) are involved in the degradation of storage proteins, hydrolyze seed storage proteins to provide amino acids for seed germination, and play an important regulatory role in plant growth and development, disease resistance, and adversity stress processes [80]. Studies have shown that the Fagopyrum esculentummoench AP member Fe AP9 is upregulated in response to darkness, drought and salicylic acid stress [81]. Overexpressing the grape AP gene VlAP17 can improve salt tolerance and drought tolerance by protecting the integrity of the plasma membrane and increasing the root quality of seedlings [82]. In this study, the expression of aspartic proteinase nepenthesin-2 (W9RZ71) and bark storage protein A (W9RDI0) were significantly upregulated, which indicated that the salt tolerance of mulberry seeds might be enhanced by hydrolyzing and storing protein during germination.

4.4 Salt Stress Induces Changes in Signal Transduction and Substance Transport

In this study, we studied the changes in signal transduction and substance transport caused by salt stress. We screened 14 DAPs (7 upregulated and 7 downregulated) under salt stress (50 mM NaCl) and 51 DAPs (16 upregulated and 35 downregulated) under salt stress (100 mM NaCl), while 17 DAPs (7 upregulated and 10 downregulated) were screened in two salt treatments. Seed germination determines the beginning of the seed plant life cycle, which is a complex process that involves various signal transduction pathways [83]. Receptor-like kinase is a common key signaling molecule in many stress reactions that can mediate the response of plants to development and environmental stimuli [84]. Receptor-like protein kinase may play a role in transducing extracellular information into cells. The kinase consists of an extracellular domain, transmembrane domain and cytoplasmic-located serine/threonine protein kinase domain [85]. Studies have shown that the plant hormone abscisic acid (ABA) regulates seed germination, plant growth and development, while protein kinase may positively regulate ABA signal transduction in seed germination and seedling growth [86]. In this study, many protein kinases related to signal transduction were found to be significantly upregulated, including putative receptor-like protein kinase (W9RBX0), receptor-like protein kinase HERK 1 (W9RFT8), putative serine/threonine protein kinase (W9RNL9), serine/threonine protein kinase PRP4-like protein (W9RJN3) and putative LRR receptor-like serine/threonine protein kinase (W9SN33), suggesting that these proteins may play a critical role in the germination of mulberry seeds. Under stress, plants induce transporters to transport signal molecules to express some stress resistance genes and cope with adverse environments [87]. In wheat roots under NaCl stress, several K+ transporters, such as voltage-gated potassium channels, cyclic nucleotide-gated potassium channels and ABC transporters, increased; thus, increasing potassium absorption is an important strategy to regulate root ion homeostasis to cope with salt and alkali stress [88, 89]. As a potential nonselective cation transporter, cyclic nucleotide gated channels (CNGCs) can participate in the absorption and/or transport of monovalent and possible divalent cations and play a role in plant defense and seed germination [90]. It has been reported that CNGC10 negatively regulates the salt tolerance of Arabidopsis thaliana, indicating that it may inhibit the salt tolerance of Arabidopsis thaliana seed germination by participating in mediating Na+ transport [91]. In this study, the estimated cyclic nucleotide gated ion channel 8 (W9R5Q9) was significantly downregulated, while the ABC transporter G family member 28 (W9SAJ0) was significantly upregulated. However, studies have shown that the ABC transporter is directly powered by ATP, which can transport complex organic substances against the concentration gradient. ATP is essential for plant growth and development and plays an important role in seed development, seed germination, abiotic stress response and the interaction between plants and the environment [92, 93]. Overexpressing Arabidopsis thaliana BCG 36 reduces the sodium content in buds and roots, thus endowing the plant with salt tolerance [94]. Early studies have shown that salt stress mainly reduces NO3- absorption by regulating different nitrate transporters [95]. The expression of the nitrate transporter gene family (OsNRT) in rice is downregulated, which indicates that salt stress affects its expression level, which may inhibit the accumulation of NO3- during rice seed germination [96]. In this study, nitrate transporter 1.1 (W9RGK8) was significantly downregulated, which may indicate that the accumulation of NO-3 under salt stress was also inhibited. These results indicate that these substance transport signaling molecules may respond to salt stress through different expression patterns.

4.5 Salt Stress Induces Changes in Protein Synthesis, Folding and Degradation

In this study, to examine the changes in protein synthesis, folding and degradation caused by salt stress, we screened 7 DAPs (2 upregulated and 5 downregulated) under salt stress (50 mM NaCl) and 39 DAPs (9 upregulated and 30 downregulated) under salt stress (100 mM NaCl), while 5 DAPs (3 upregulated and 2 downregulated) were screened under two salt treatments. Proteins enable life activities, and when seeds germinate, the change in soluble protein content can reflect various information on protein synthesis, denaturation and degradation in cells. In many plants, such as Arabidopsis [97], Ricinus communis [98] and Phoenix dactylifera [99], protein folding and stability involve protein synthesis, including ribosomal proteins. Ribosomes consist of two parts, namely, ribosomal protein (RP) and ribosomal RNA. RP not only maintains the configuration of RNA but also participates in the synthesis, transportation and localization of proteins [100]. Studies have shown that under Na2CO3 stress, almost all the genes involved in protein synthesis were downregulated in Puccinellia tenuiflora seedlings, which indicated that the protein synthesis mechanism of Puccinellia tenuiflora was inhibited by Na2CO3 stress [101]. Additionally, the abundance of NaHCO3 and NaCl was reduced in ribosomal proteins in tomato [102] and Arabidopsis [103]. Moreover, 30S ribosomal protein S1 in annual ryegrass and 50S ribosomal protein in perennial ryegrass decreased under salt stress, which affected the folding and synthesis of protein in annual and perennial ryegrass [104]. In this study, many of the proteins involved in protein synthesis were significantly downregulated, including the ribosomal protein family 30S ribosomal protein S21 (W9SIU2), 50S ribosomal protein L28 (W9QPT4), 50S ribosomal protein L34 (W9QWL4) and 60S ribosomal protein L28-1 (W9QUV3), indicating that the protein synthesis mechanism of mulberry is inhibited by salt stress during seed germination, thus affecting protein folding and synthesis.

Studies have shown that ubiquitin-binding enzymes are highly involved in proteolysis, protein modification and protein degradation, including abiotic stress, and play an important role in all aspects of plant growth, development and cell processes [105]. It has been reported that overexpressing the soybean ubiquitin-binding enzyme gene GmUBC2 can enhance salt tolerance by regulating the expression of abiotic stress-responsive genes in Arabidopsis thaliana [106]. DnaJ plays an important role in the NaCl tolerance of Arabidopsis thaliana and is a molecular chaperone involved in many cellular processes, including the regulation of protein folding, protein degradation, protein activity and protein aggregation prevention [107]. In this study, proteins related to protein folding and degradation were significantly upregulated, including ubiquitin-binding enzyme E2 27 (W9RHB5), DnaJ-like protein (W9QJ41) and peptidyl prolyl cis-trans isomerase (W9RAD0). This result indicated that the folding and stability of these proteins might be enhanced to cope with the salt tolerance of mulberry seed germination.

5. Conclusions

This study explored the response mechanism of mulberry seed germination to salt stress at the physiological and proteomic levels. Physiological data showed that salt stress inhibited the germination rate and radicle length of mulberry seeds, decreased the content of MDA, and significantly increased the activities of SOD, POD and CAT. Then, the TMT marker technique was used to analyze the protein group of mulberry seeds in two salt treatment stages, and 76,544 unique peptides were detected. After removing duplicate proteins, 7717 proteins were identified according to TMT data. Through proteomic analysis, we obtained 143 (50 mM NaCl) and 540 (100 mM NaCl) DAPs associated with mulberry seed germination. Furthermore, a total of 113 DAPs were treated with 50 mM NaCl and 100 mM NaCl solutions, of which 43 were upregulated and 70 were downregulated. KEGG analysis showed that the proteins involved in photosynthesis, carotenoid biosynthesis and plant hormone signal transduction were usually enriched in mulberry seeds. Finally, five differential proteins were verified by PRM, which validated the credibility of TMT in proteomic analyses. The data presented herein provide new insights into plant salt-responsive proteins, with potential implications for enhancing the salt tolerance of mulberry plants. Future research requires further analysis at the RNA and RT-PCR levels for deeper molecular level studies to provide a more in-depth functional study of the genes and proteins involved in salinity tolerance.

Availability of Data and Materials

The datasets used and analyzed during the current study are available from the corresponding author on reasonable request.

Author Contributions

YW conceived and designed this study; YW, WJ, JC, WG and YL performed the experiments, interpreted the results and wrote the article; YW, WJ, WG, YL and CL revised the article and analyzed the data. All the authors read and approved the final version.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

Not applicable.

Funding

This work was supported by Guangdong Provincial Department of Education Youth Innovative Talents Project (No. 2019KQNCX050), Guangzhou Basic and Applied Basic Research Foundation (No. 202201011164), Forestry Scientific technology Innovation Project of Guangdong Province (No. 2021KJCX015) and Scarce and Quality Economic Forest Engineering Technology Research Center (No. 2022GCZX002). The funders had no role in the study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
Long R, Gao Y, Sun H, Zhang T, Li X, Li M, et al. Quantitative proteomic analysis using iTRAQ to identify salt-responsive proteins during the germination stage of two Medicago species. Scientific Reports. 2018; 8: 9553.
[2]
Nadeem M, Li J, Yahya M, Wang M, Ali A, Cheng A, et al. Grain Legumes and Fear of Salt Stress: Focus on Mechanisms and Management Strategies. International Journal of Molecular Sciences. 2019; 20: 799.
[3]
Ivushkin K, Bartholomeus H, Bregt AK, Pulatov A, Kempen B, Sousa LD. Global mapping of soil salinity change. Remote Sensing of Environment.2019; 231: 111260.
[4]
Shaheen S, Naseer S, Ashraf M, Akram NA. Salt stress affects water relations, photosynthesis, and oxidative defense mechanisms in Solanum melongena L. Journal of Plant Interactions. 2013; 8: 85–96.
[5]
Singh A, Rajput V, Rawat S, Singh AK, Bind A, Singh A, et al. Monitoring soil salinity and recent advances in mechanism of salinity tolerance in plants. Biogeosystem Technique. 2021; 2: 66–87.
[6]
Mathur N, Singh J, Bohra S, Bohra A, Vyas A. Biomass Production, Productivity and Physiological Changes in Moth Bean Genotypes at Different Salinity Levels. American Journal of Plant Physiology. 2006; 1: 210–213.
[7]
Jamil M, Rehman S, Rha ES. Salinity effect on plant growth, PSII photochemistry and chlorophyll content in sugar beet (Beta vulgaris L.) and cabbage (Brassica oleracea capitata L.). Pakistan Journal of Botany. 2007; 39: 753–760.
[8]
Zhao GQ, Ma BL, Ren CZ. Growth, Gas Exchange, Chlorophyll Fluorescence, and Ion Content of Naked Oat in Response to Salinity. Crop Science. 2007; 47: 123–131.
[9]
Lan Y, Zhang K, He T, Wang H, Jiang C, Yan H, et al. Systematic analysis of the Serine/Arginine-Rich Protein Splicing Factors (SRs) and focus on salt tolerance of PtSC27 in Populus trichocarpa. Plant Physiology and Biochemistry. 2022; 173: 97–109.
[10]
Dai J, Hu A, Zhang J, Liao W, Ma H, Wu J, et al. NF-YB-Mediated Active Responses of Plant Growth under Salt and Temperature Stress in Eucalyptus grandis. Plants. 2021; 10: 1107.
[11]
Chen Z, Lin S, Li J, Chen T, Gu Q, Yang T, et al. Theanine Improves Salt Stress Tolerance via Modulating Redox Homeostasis in Tea Plants (Camellia sinensis L.). Frontiers in Plant Science. 2021; 12: 770398.
[12]
Mahajan S, Tuteja N. Cold, salinity and drought stresses: an overview. Archives of Biochemistry and Biophysics. 2005; 444: 139–158.
[13]
Zhang G, Zhou J, Peng Y, Tan Z, Li L, Yu L, et al. Genome-Wide Association Studies of Salt Tolerance at Seed Germination and Seedling Stages in Brassica napus. Frontiers in Plant Science. 2022; 12: 772708.
[14]
Alencar NLM, Gadelha CG, Gallão MI, Dolder MAH, Prisco JT, Gomes-Filho E. Ultrastructural and biochemical changes induced by salt stress in Jatropha curcas seeds during germination and seedling development. Functional Plant Biology. 2015; 42: 865–874.
[15]
Shrivastava P, Kumar R. Soil salinity: A serious environmental issue and plant growth promoting bacteria as one of the tools for its alleviation. Saudi Journal of Biological Sciences. 2015; 22: 123–131.
[16]
Sholi N. Effect of Salt Stress on Seed Germination, Plant Growth, Photosynthesis and Ion Accumulation of four Tomato Cultivars. American Journal of Plant Physiology. 2012; 7: 269–275.
[17]
Chen L, Lu B, Liu L, Duan W, Jiang D, Li J, et al. Melatonin promotes seed germination under salt stress by regulating ABA and GA_3 in cotton (Gossypium hirsutum L.). Plant Physiology and Biochemistry. 2021; 162: 506–516.
[18]
Wang H, Tian C, Wang L. Germination of dimorphic seeds of Suaeda aralocaspica in response to light and salinity conditions during and after cold stratification. PeerJ. 2017; 5: e3671.
[19]
El Moukhtari A, Cabassa-Hourton C, Farissi M, Savouré A. How Does Proline Treatment Promote Salt Stress Tolerance During Crop Plant Development? Frontiers in Plant Science. 2020; 11: 1127.
[20]
Zulfiqar F, Ashraf M. Nanoparticles potentially mediate salt stress tolerance in plants. Plant Physiology and Biochemistry. 2021; 160: 257–268.
[21]
Zhang HH, Shi GL, Shao JY, Li X, Li MB, Meng L, et al. Photochemistry and proteomics of mulberry (Morus alba L.) seedlings under NaCl and NaHCO_3 stress. Ecotoxicology and Environmental Safety. 2019; 184: 109624.
[22]
He N, Zhang C, Qi X, Zhao S, Tao Y, Yang G, et al. Draft genome sequence of the mulberry tree Morus notabilis. Nature Communications. 2013; 4: 2445.
[23]
Liu Y, Ji D, Turgeon R, Chen J, Lin T, Huang J, et al. Physiological and Proteomic Responses of Mulberry Trees (Morus alba. L.) to Combined Salt and Drought Stress. International Journal of Molecular Sciences. 2019; 20: 2486.
[24]
Rohela GK, Muttanna PS, Kumar R, Chowdhury SR. Mulberry (Morus spp.): An Ideal Plant for Sustainable Development. Trees Forests and People. 2020; 2: 100011.
[25]
Liu C, Fan W, Zhu P, Xia Z, Hu J, Zhao A. Mulberry RGS negatively regulates salt stress response and tolerance. Plant Signaling & Behavior. 2019; 14: 1672512.
[26]
Liu S, Ou Y, Li Y, Sulaiman K, Tao M, Shawky E, et al. Tandem mass tag-based proteomic analysis of endoplasmic reticulum proteins in mulberry leaves under ultraviolet-B and dark stress. Physiologia Plantarum. 2022; 174: e13667.
[27]
Guo Z, Zeng P, Xiao X, Peng C. Physiological, anatomical, and transcriptional responses of mulberry (Morus alba L.) to Cd stress in contaminated soil. Environmental Pollution. 2021; 284: 117387.
[28]
Li R, Su X, Zhou R, Zhang Y, Wang T. Molecular mechanism of mulberry response to drought stress revealed by complementary transcriptomic and iTRAQ analyses. BMC Plant Biology. 2022; 22: 36.
[29]
Ma Q, Kang J, Long R, Zhang T, Xiong J, Zhang K, et al. Comparative proteomic analysis of alfalfa revealed new salt and drought stress-related factors involved in seed germination. Molecular Biology Reports. 2017; 44: 261–272.
[30]
Liu D, Han C, Deng X, Liu Y, Liu N, Yan Y. Integrated physiological and proteomic analysis of embryo and endosperm reveals central salt stress response proteins during seed germination of winter wheat cultivar Zhengmai 366. BMC Plant Biology. 2019; 19: 29.
[31]
Xu X, Fan R, Zheng R, Li C, Yu D. Proteomic analysis of seed germination under salt stress in soybeans. Journal of Zhejiang University-Science B. 2011; 12: 507–517.
[32]
Liu S, Xu H, Wang W, Li N, Wang W, Møller IM, et al. A proteomic analysis of rice seed germination as affected by high temperature and ABA treatment. Physiologia Plantarum. 2015; 154: 142–161.
[33]
Gu J, Chao H, Gan L, Guo L, Zhang K, Li Y, et al. Proteomic Dissection of Seed Germination and Seedling Establishment in Brassica napus. Frontiers in Plant Science. 2016; 7: 1482.
[34]
Zhang H, Wang W, Liu S, Møller IM, Song S. Proteome Analysis of Poplar Seed Vigor. PLoS ONE. 2015; 10: e0132509.
[35]
Yang M, Liu Y, Liu Y, Chen H, Chen F, Shen S. Proteomic analysis of oil mobilization in seed germination and postgermination development of Jatropha curcas. Journal of Proteome Research. 2009; 8: 1441–1451.
[36]
Balbuena TS, Jo L, Pieruzzi FP, Dias LLC, Silveira V, Santa-Catarina C, et al. Differential proteome analysis of mature and germinated embryos of Araucaria angustifolia. Phytochemistry. 2011; 72: 302–311.
[37]
Pawłowski TA, Staszak AM. Analysis of the embryo proteome of sycamore (Acer pseudoplatanus L.) seeds reveals a distinct class of proteins regulating dormancy release. Journal of Plant Physiology. 2016; 195: 9–22.
[38]
Fercha A, Capriotti AL, Caruso G, Cavaliere C, Samperi R, Stampachiacchiere S, et al. Comparative analysis of metabolic proteome variation in ascorbate-primed and unprimed wheat seeds during germination under salt stress. Journal of Proteomics. 2014; 108: 238–257.
[39]
Fu Z, Jin X, Ding D, Li Y, Fu Z, Tang J. Proteomic analysis of heterosis during maize seed germination. Proteomics. 2011; 11: 1462–1472.
[40]
Yang Z, Li J, Liu L, Xie Q, Sui N. Photosynthetic Regulation Under Salt Stress and Salt-Tolerance Mechanism of Sweet Sorghum. Frontiers in Plant Science. 2020; 10: 1722.
[41]
Batkhuu N, Enkhchimeg T, Jamsran T. Salt tolerance of Ulmus pumila L. seedlings from geographycally different seed sources [C]. XXIV IUFRO World Congress Sustaining Forests, Sustaining People: The Role of Research. 2014.
[42]
Sarkar T, Mogili T, Sivaprasad V. Improvement of abiotic stress adaptive traits in mulberry (Morus spp.): an update on biotechnological interventions. 3 Biotech. 2017; 7: 214.
[43]
Dai F, Luo G, Li Z, Wei X, Wang Z, Lin S, et al. Physiological and transcriptomic analyses of mulberry (Morus atropurpurea) response to cadmium stress. Ecotoxicology and Environmental Safety. 2020; 205: 111298.
[44]
Du Q, Guo P, Shi Y, Zhang J, Zheng D, Li Y, et al. Genome-Wide Identification of Copper Stress-Regulated and Novel MicroRNAs in Mulberry Leaf. Biochemical Genetics. 2021; 59: 589–603.
[45]
Wiśniewski JR, Zougman A, Nagaraj N, Mann M. Universal sample preparation method for proteome analysis. Nature Methods. 2009; 6: 359–362.
[46]
Kanehisa M, Goto S, Sato Y, Furumichi M, Tanabe M. KEGG for integration and interpretation of large-scale molecular data sets. Nucleic Acids Research. 2012; 40: D109–D114.
[47]
Tlahig S, Bellani L, Karmous I, Barbieri F, Loumerem M, Muccifora S. Response to Salinity in Legume Species: An Insight on the Effects of Salt Stress during Seed Germination and Seedling Growth. Chemistry & Biodiversity. 2021; 18: e2000917.
[48]
Gao Y, Cui Y, Long R, Sun Y, Zhang T, Yang Q, et al. Salt-stress induced proteomic changes of two contrasting alfalfa cultivars during germination stage. Journal of the Science of Food and Agriculture. 2019; 99: 1384–1396.
[49]
Han X, Yang F, Zhao Y, Chen H, Wan Z, Li L, et al. iTRAQ based protein profile analysis revealed key proteins involved in regulation of drought-tolerance during seed germination in Adzuki bean. Scientific Reports. 2021; 11: 23725.
[50]
Bela K, Horváth E, Gallé Á, Szabados L, Tari I, Csiszár J. Plant glutathione peroxidases: emerging role of the antioxidant enzymes in plant development and stress responses. Journal of Plant Physiology. 2015; 176: 192–201.
[51]
Davey MW, Stals E, Panis B, Keulemans J, Swennen RL. High-throughput determination of malondialdehyde in plant tissues. Analytical Biochemistry. 2005; 347: 201–207.
[52]
Luo Q, Peng M, Zhang X, Lei P, Ji X, Chow W, et al. Comparative mitochondrial proteomic, physiological, biochemical and ultrastructural profiling reveal factors underpinning salt tolerance in tetraploid black locust (Robinia pseudoacacia L.). BMC Genomics. 2017; 18: 648.
[53]
Chen X, Shi L, Chen Y, Zhu L, Zhang D, Xiao S, et al. Arabidopsis HSP70-16 is required for flower opening under normal or mild heat stress temperatures. Plant, Cell & Environment. 2019; 42: 1190–1204.
[54]
Waters ER, Vierling E. Plant small heat shock proteins - evolutionary and functional diversity. The New Phytologist. 2020; 227: 24–37.
[55]
Jung YJ, Nou IS, Kang KK. Overexpression of Oshsp16.9 gene encoding small heat shock protein enhances tolerance to abiotic stresses in rice. Plant Breeding and Biotechnology. 2014; 2: 370–379.
[56]
Li Z, Long R, Zhang T, Yang Q, Kang J. Molecular cloning and characterization of the MsHSP17.7 gene from Medicago sativa L. Molecular Biology Reports. 2016; 43: 815–826.
[57]
Osorio CE. The Role of Orange Gene in Carotenoid Accumulation: Manipulating Chromoplasts Toward a Colored Future. Frontiers in Plant Science. 2019; 10: 1235.
[58]
Zhu C, Yamamura S, Nishihara M, Koiwa H, Sandmann G. cDNAs for the synthesis of cyclic carotenoids in petals of Gentiana lutea and their regulation during flower development. Biochimica Et Biophysica Acta. 2003; 1625: 305–308.
[59]
Ke Q, Kang L, Kim HS, Xie T, Liu C, Ji CY, et al. Down-regulation of lycopene ε-cyclase expression in transgenic sweetpotato plants increases the carotenoid content and tolerance to abiotic stress. Plant Science. 2019; 281: 52–60.
[60]
Liu Z, Zou L, Chen C, Zhao H, Yan Y, Wang C, et al. iTRAQ-based quantitative proteomic analysis of salt stress in Spica Prunellae. Scientific Reports. 2019; 9: 9590.
[61]
Wang Y, Peng X, Salvato F, Wang Y, Yan X, Zhou Z, et al. Salt-adaptive strategies in oil seed crop Ricinus communis early seedlings (cotyledon vs. true leaf) revealed from proteomics analysis. Ecotoxicology and Environmental Safety. 2019; 171: 12–25.
[62]
Zhang H, Irving LJ, McGill C, Matthew C, Zhou D, Kemp P. The effects of salinity and osmotic stress on barley germination rate: sodium as an osmotic regulator. Annals of Botany. 2010; 106: 1027–1035.
[63]
Jung H, Moon S. Purification, distribution, and characterization activity of lipase from oat seeds (Avena sativa L.). Journal of the Korean Society for Applied Biological Chemistry. 2013; 56: 639–645.
[64]
Naranjo MA, Forment J, Roldán M, Serrano R, Vicente O. Overexpression of Arabidopsis thaliana LTL1, a salt-induced gene encoding a GDSL-motif lipase, increases salt tolerance in yeast and transgenic plants. Plant, Cell & Environment. 2006; 29: 1890–1900.
[65]
Sela A, Piskurewicz U, Megies C, Mène-Saffrané L, Finazzi G, Lopez-Molina L. Embryonic Photosynthesis Affects Post-Germination Plant Growth. Plant Physiology. 2020; 182: 2166–2181.
[66]
Allorent G, Osorio S, Vu JL, Falconet D, Jouhet J, Kuntz M, et al. Adjustments of embryonic photosynthetic activity modulate seed fitness in Arabidopsis thaliana. The New Phytologist. 2015; 205: 707–719.
[67]
Shaki F, Ebrahimzadeh Maboud H, Niknam V. Differential proteomics: Effect of growth regulators on salt stress responses in safflower seedlings. Pesticide Biochemistry and Physiology. 2020; 164: 149–155.
[68]
Minic Z. Physiological roles of plant glycoside hydrolases. Planta. 2008; 227: 723–740.
[69]
Lopez-Casado G, Urbanowicz BR, Damasceno CMB, Rose JKC. Plant glycosyl hydrolases and biofuels: a natural marriage. Current Opinion in Plant Biology. 2008; 11: 329–337.
[70]
Sharma R, Cao P, Jung K, Sharma MK, Ronald PC. Construction of a rice glycoside hydrolase phylogenomic database and identification of targets for biofuel research. Frontiers in Plant Science. 2013; 4: 330.
[71]
Xu H, Liu S, Song S, Wang R, Wang W, Song S. Proteomics analysis reveals distinct involvement of embryo and endosperm proteins during seed germination in dormant and non-dormant rice seeds. Plant Physiology and Biochemistry. 2016; 103: 219–242.
[72]
Vítámvás P, Prášil IT, Kosová K, Planchon S, Renaut J. Analysis of proteome and frost tolerance in chromosome 5A and 5B reciprocal substitution lines between two winter wheats during long-term cold acclimation. Proteomics. 2012; 12: 68–85.
[73]
Passamani LZ, Barbosa RR, Reis RS, Heringer AS, Rangel PL, Santa-Catarina C, et al. Salt stress induces changes in the proteomic profile of micropropagated sugarcane shoots. PLoS ONE. 2017; 12: e0176076.
[74]
Tang W, Zhong L, Ding Q, Dou Y, Li W, Xu Z, et al. Histone deacetylase AtSRT2 regulates salt tolerance during seed germination via repression of vesicle-associated membrane protein 714 (VAMP714) in Arabidopsis. The New Phytologist. 2022; 234: 1278–1293.
[75]
Grasser KD, Launholt D, Grasser M. High mobility group proteins of the plant HMGB family: dynamic chromatin modulators. Biochimica Et Biophysica Acta. 2007; 1769: 346–357.
[76]
Volokita M, Rosilio-Brami T, Rivkin N, Zik M. Combining comparative sequence and genomic data to ascertain phylogenetic relationships and explore the evolution of the large GDSL-lipase family in land plants. Molecular Biology and Evolution. 2011; 28: 551–565.
[77]
Ling H. Sequence analysis of GDSL lipase gene family in Arabidopsis thaliana. Pakistan Journal of Biological Sciences. 2008; 11: 763–767.
[78]
Shen G, Sun W, Chen Z, Shi L, Hong J, Shi J. Plant GDSL Esterases/Lipases: Evolutionary, Physiological and Molecular Functions in Plant Development. Plants. 2022; 11: 468.
[79]
Lai C, Huang L, Chen LO, Chan M, Shaw J. Genome-wide analysis of GDSL-type esterases/lipases in Arabidopsis. Plant Molecular Biology. 2017; 95: 181–197.
[80]
Lu X, Zhang X, Mei M, Liu G, Ma B. Proteomic analysis of Magnolia sieboldii K. Koch seed germination. Journal of Proteomics. 2016; 133: 76–85.
[81]
Timotijević GS, Milisavljević MD, Radović SR, Konstantinović MM, Maksimović VR. Ubiquitous aspartic proteinase as an actor in the stress response in buckwheat. Journal of Plant Physiology. 2010; 167: 61–68.
[82]
Guo R, Zhao J, Wang X, Guo C, Li Z, Wang Y, et al. Constitutive expression of a grape aspartic protease gene in transgenic Arabidopsis confers osmotic stress tolerance. Plant Cell, Tissue and Organ Culture. 2015; 121: 275–287.
[83]
Li H, Li X, Wang G, Zhang J, Wang G. Analysis of gene expression in early seed germination of rice: landscape and genetic regulation. BMC Plant Biology. 2022; 22: 70.
[84]
Zhang X, Yang G, Shi R, Han X, Qi L, Wang R, et al. Arabidopsis cysteine-rich receptor-like kinase 45 functions in the responses to abscisic acid and abiotic stresses. Plant Physiology and Biochemistry. 2013; 67: 189–198.
[85]
Braun DM, Stone JM, Walker JC. Interaction of the maize and Arabidopsis kinase interaction domains with a subset of receptor-like protein kinases: implications for transmembrane signaling in plants. The Plant Journal. 1997; 12: 83–95.
[86]
Fujii H, Verslues PE, Zhu J. Identification of two protein kinases required for abscisic acid regulation of seed germination, root growth, and gene expression in Arabidopsis. The Plant Cell. 2007; 19: 485–494.
[87]
Li P, Luo T, Pu X, Zhou Y, Yu J, Liu L. Plant transporters: roles in stress responses and effects on growth and development. Plant Growth Regulation. 2021; 93: 253–266.
[88]
Peng Z, Wang M, Li F, Lv H, Li C, Xia G. A proteomic study of the response to salinity and drought stress in an introgression strain of bread wheat. Molecular & Cellular Proteomics. 2009; 8: 2676–2686.
[89]
Wang M, Peng Z, Li C, Li F, Liu C, Xia G. Proteomic analysis on a high salt tolerance introgression strain of Triticum aestivum/Thinopyrum ponticum. Proteomics. 2008; 8: 1470–1489.
[90]
Gobert A, Park G, Amtmann A, Sanders D, Maathuis FJM. Arabidopsis thaliana cyclic nucleotide gated channel 3 forms a non-selective ion transporter involved in germination and cation transport. Journal of Experimental Botany. 2006; 57: 791–800.
[91]
Jin Y, Jing W, Zhang Q, Zhang W. Cyclic nucleotide gated channel 10 negatively regulates salt tolerance by mediating Na+ transport in Arabidopsis. Journal of Plant Research. 2015; 128: 211–220.
[92]
Do THT, Martinoia E, Lee Y. Functions of ABC transporters in plant growth and development. Current Opinion in Plant Biology. 2018; 41: 32–38.
[93]
Yao Y, Sun Y, Feng Q, Zhang X, Gao Y, Ou Y, et al. Acclimation to nitrogen × salt stress in Populus bolleana mediated by potassium/sodium balance. Industrial Crops and Products. 2021; 170: 113789.
[94]
Kim D, Jin J, Alejandro S, Martinoia E, Lee Y. Overexpression of AtABCG36 improves drought and salt stress resistance in Arabidopsis. Physiologia Plantarum. 2010; 139: 170–180.
[95]
Kang J, Park J, Choi H, Burla B, Kretzschmar T, Lee Y, et al. Plant ABC Transporters. The Arabidopsis Book. 2011; 9: e0153.
[96]
Shi Y, Gao L, Wu Z, Zhang X, Wang M, Zhang C, et al. Genome-wide association study of salt tolerance at the seed germination stage in rice. BMC Plant Biology. 2017; 17: 92.
[97]
Job C, Rajjou L, Lovigny Y, Belghazi M, Job D. Patterns of protein oxidation in Arabidopsis seeds and during germination. Plant Physiology. 2005; 138: 790–802.
[98]
Nogueira FCS, Palmisano G, Schwämmle V, Soares EL, Soares AA, Roepstorff P, et al. Isotope labeling-based quantitative proteomics of developing seeds of castor oil seed (Ricinus communis L.). Journal of Proteome Research. 2013; 12: 5012–5024.
[99]
Sghaier-Hammami B, Valledor L, Drira N, Jorrin-Novo JV. Proteomic analysis of the development and germination of date palm (Phoenix dactylifera L.) zygotic embryos. Proteomics. 2009; 9: 2543–2554.
[100]
Yan S, Chong P, Zhao M, Liu H. Physiological response and proteomics analysis of Reaumuria soongorica under salt stress. Scientific Reports. 2022; 12: 2539.
[101]
Zhao Q, Suo J, Chen S, Jin Y, Ma X, Yin Z, et al. Na2CO3-responsive mechanisms in halophyte Puccinellia tenuiflora roots revealed by physiological and proteomic analyses. Scientific Reports. 2016; 6: 32717.
[102]
Gong B, Zhang C, Li X, Wen D, Wang S, Shi Q, et al. Identification of NaCl and NaHCO3 stress responsive proteins in tomato roots using iTRAQ-based analysis. Biochemical and Biophysical Research Communications. 2014; 446: 417–422.
[103]
Jiang Y, Yang B, Harris NS, Deyholos MK. Comparative proteomic analysis of NaCl stress-responsive proteins in Arabidopsis roots. Journal of Experimental Botany. 2007; 58: 3591–3607.
[104]
Peng X, Yu D, Yan J, Zhang N, Lin J, Wang J. Physiological and Proteomic Analyses Reveal Adaptive Mechanisms of Ryegrass (Annual vs. Perennial) Seedlings to Salt Stress. Agronomy. 2019; 9: 843.
[105]
Sharma B, Bhatt TK. Genome-wide identification and expression analysis of E2 ubiquitin-conjugating enzymes in tomato. Scientific Reports. 2017; 7: 8613.
[106]
Zhou G, Chang R, Qiu L. Overexpression of soybean ubiquitin-conjugating enzyme gene GmUBC2 confers enhanced drought and salt tolerance through modulating abiotic stress-responsive gene expression in Arabidopsis. Plant Molecular Biology. 2010; 72: 357–367.
[107]
Zhao Z, Zhang W, Yan J, Zhang J, Liu Z, Li X, et al. Over-expression of Arabidopsis DnaJ (Hsp40) contributes to NaCl-stress tolerance. African Journal of Biotechnology. 2010; 9: 972–978.

Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share
Back to top