Next Article in Journal
Liverwort and Hornwort Flora of Hoàng Liên National Park and the Adjacent Areas (North Vietnam, Indochina)
Previous Article in Journal
Optimizing Insecticide Application Timing for Broad Bean Weevil Control and Minimizing Crop Damage in Broad Bean (Vicia faba Linn.)
Previous Article in Special Issue
Variation in Seed Dormancy of Chaco Seasonally Dry Forest Species: Effects of Seed Traits and Population Environmental Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

iTRAQ-Based Proteomic and Physiological Analyses Reveal the Mechanisms of Dehydration and Cryopreservation Tolerance of Sophora tonkinensis Gagnep. Seeds

1
Hubei Key Laboratory of Biologic Resources Protection and Utilization, Hubei Minzu University, Enshi 445000, China
2
Agro-biological Gene Research Center, Guangdong Academy of Agricultural Sciences, Guangzhou 510640, China
3
Guangdong Key Laboratory for Crop Germplasm Resources Preservation and Utilization, Guangzhou 510640, China
4
Research Center for Germplasm Engineering of Characteristic Plant Resources in Enshi Prefecture, Hubei Minzu University, Enshi 445000, China
5
The Plant Germplasm Resources Laboratory, School of Forestry and Horticulture, Hubei Minzu University, Enshi 445000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2023, 12(9), 1842; https://doi.org/10.3390/plants12091842
Submission received: 9 March 2023 / Revised: 23 April 2023 / Accepted: 26 April 2023 / Published: 29 April 2023
(This article belongs to the Special Issue Advances in Seed Physiology)

Abstract

:
Sophora tonkinensi is a shrub of the genus Sophora in the family Fabaceae with anti-inflammatory and pain-relieving effects. While the cultivation, chemical makeup, and medicinal properties of S. tonkinensis have been reported, the physiological mechanisms governing its dehydration and cryopreservation tolerance of seeds remain unclear. In this study, we investigated the morphological, physiological, biochemical, and protein expression characteristics of S. tonkinensis seeds subjected to dehydration and cryopreservation techniques via the observation of cell microstructure, determination of antioxidant enzyme activity, and iTRAQ-based proteomic analysis, respectively. The results of the study demonstrated that the seeds possessed a certain level of tolerance to dehydration. The highest germination percentage of 83.0% was observed after 2 h of dehydration (10.1% water content), which was identified as the optimal time point for cryopreservation. However, the germination percentage was reduced to only 30.5% when the water content reached 5.4%, indicating that S. tonkinensis seeds exhibit intermediate storage behavior. Further investigation revealed that during seed dehydration and cryopreservation treatment, liposomes were gradually and highly fused, whereas the activities of ROS scavenging and stress defense were significantly enhanced. During dehydration, the seed tissues formed a protective mechanism of stress resistance based on protein processing in the endoplasmic reticulum and antioxidant system, which was related to the dehydration tolerance. Moreover, only three differentially expressed LEA proteins were identified, and it is speculated that the strengthening of intracellular metabolism and the absence of specific LEA and dehydrins could be crucial factors for the reduced germination percentage after excessive dehydration and cryopreservation.

1. Introduction

Dehydration tolerance (DT) is defined as the ability of an organism or tissue to survive without irreversible damage in the state of losing all or almost all of the cellular water [1]. DT of seeds is an adaptive mechanism for plants to ensure the survival and reproduction of species in the long-term evolution process and plays a key role in the preservation of plant seeds and germplasm resources [2]. It was found that the DT of orthodox seeds was gradually formed during the development process and reached the peak at physiological maturity stage. Orthodox seeds could maintain vitality after losing up to 95% of water, and then gradually lost DT during the stage of germination; recalcitrant seeds were intolerant to dehydration and highly sensitive to water loss throughout development; intermediate seeds had some tolerance to dehydration [3].
Seeds DT is a complex trait. A series of processes or mechanisms have been suggested to confer or contribute to DT, including accumulation of putatively protective molecules, e.g., late embryogenesis abundant (LEA) proteins and dehydrins [4], presence and efficient operation of antioxidant system [5], intracellular de-differentiation [6], metabolic ‘switching off’ [7], the presence and operation of repair systems during the rehydration of seeds [8,9,10], protective effects of heat shock proteins (HSP) and non-reducing raffinose families [11,12,13,14], and transcription factor networks and regulation of the AFL (ABA INSENSITIVE3, FUSCA3, and LEAFY COTYLEDON 2) subfamily [2,15,16]. The coordinated induction of protective mechanisms involves complex regulatory networks and signaling pathways. The degree of DT of seeds and embryos was negatively correlated with the respiratory activity of mitochondria, which was higher in recalcitrant seeds than in normal ones and the respiratory activity of recalcitrant seeds was higher than that of orthodox seeds [17]. During dehydration, the H2O2 content, the production rate of superoxide anion radical (O2·−), and the content of thiobarbituric acid reactive substance of DT embryos (axes) were significantly lower than that of dehydration-sensitive embryos (axes) [18], while the activity of reactive oxygen species (ROS) scavenging systems, including enzymatic and non-enzymatic, was significantly higher. During seed maturation, the accumulation of LEA proteins, small HSPs, and non-reducing oligosaccharides was closely related to the formation of seeds DT [2]. The AFL subfamily of B3 transcription factors increased the seeds (embryo) DT through positively regulating the storage substances and protective proteins accumulation [19,20]. In seeds, the parallel ABA and DOG1 (DELAY OF GERMINATION 1) signaling pathways activate the synthesis of raffinose family oligosaccharides, the expression of LEA genes and HSP genes, thereby regulating the initiation of DT and transition to dormancy [21,22,23]. In addition, seed dehydration and development were also regulated by epigenetic pathways. For example, seed desiccation in the late mature stage could also trigger nuclear DNA methylation [15,24], which was usually related to long-term repression of gene transcription [25]. Michalak et al. reported the effect of desiccation on orthodox seeds of Pyrus communis L., with an immediate increase in overall DNA methylation levels at seed maturity [24]. In addition, non-coding RNAs, PTMS of histones such as acetylation, methylation, and ubiquitination, played important roles in seed formation and dehydration [25,26,27]. In contrast to the remarkable progress made in the mechanisms of DT during orthodox seeds development, our understanding of intermediate and recalcitrant seeds’ desiccation-sensitivity remains limited.
Sophora tonkinensi was a shrub in the Sophora genus, belonging to the Fabaceae family, mainly in southwest China but also in northern Vietnam [28]. The matrine alkaloids in S. tonkinensis had the effect of anti-inflammatory and analgesic, which were the main ingredients of medicine [29,30,31]. A large number of studies has been conducted on identification, cultivation, chemical composition, and pharmacology of S. tonkinensis [32,33,34,35,36]. However, there is a lack of information regarding the preservation and reproduction of its seeds. The wild resources of S. tonkinensis were narrowly distributed, with a low natural reproduction rate coupled with indiscriminate mining, making the wild resources decrease year by year, and were close to endangered status, which made the study of seed preservation and seed characteristics critical [37]. Previous studies have shown that seed propagation was the main means of cultivating S. tonkinensis, using the method of sowing as it was harvested. The seeds of S. tonkinensis are considered to be intermediate seeds due to difficulty in long-term storage and reduced germination percentage after dehydration. In this study, we aimed to investigate the response mechanism of S. tonkinensis seeds to dehydration using cell microscopic observation, antioxidant analysis, and differentially expressed protein analysis and to explore the effect of cryopreservation treatment on seed DT and vigor. It provides a theoretical foundation for the safe preservation of resources of intermediate seeds.

2. Results

2.1. Dehydration Tolerance of S. tonkinensis Seeds

The initial water content of freshly matured S. tonkinensis seeds was as high as 31%, which were dehydrated rapidly in the active silica gel at first, dropping to 13.0% after 1 h-dehydration, and then the water content gradually slowed down until to 5.4% after 24 h-dehydration (Figure 1A). The slight degree of dehydration of S. tonkinensis seeds, i.e., the water content, was 10.1~13%, promoted germination, with a highest germination percentage of 84%, indicating that fresh mature S. tonkinensis seeds have a certain degree of DT. As the degree of dehydration increased, i.e., the water content was 5.4~9.7%, the germination percentage decreased, and the germination percentage was only 30.5% when the water content was 5.4% (Figure 1B).
The water content during the cryopreservation process affected the germination percentage of seeds. All freshly harvested seeds, with 31.0% water content, died after cryopreservation, while the germination percentage after cryopreservation was 67%, 75%, 77%, and 83% at water content between 8.5% and 13.0%, respectively, all of which were not significantly different from that before cryopreservation treatment. Among which, the highest germination percentage was 83.0% for the seeds treated with cryopreservation after dehydration to a water content of 10.1% (Figure 1B). The above results indicated that S. tonkinensis seeds were intermediate seeds. After appropriate rapid dehydration treatment, the seeds have a fairly high seed vigor and certain DT, which laid foundation for cryopreservation of S. tonkinensis seeds.

2.2. Effects of Dehydration and Cryopreservation after Dehydration Treatment on the Cell Structure of S. tonkinensis Seed Plumular Axis

When the S. tonkinensis seeds were not treated with dehydration, the plumular axis cell structure was clear; the nucleus, nucleolus, and nuclear membrane were clearly visible; the nucleolus was dense; and the nucleoplasm was well-defined and clearly visible, showing a good structural and functional state. The shape of lipid bodies was regular, the cytoplasm was thick, the degree of vacuolization was low, and there were few mitochondria (Figure 2A). After 2 h-dehydration, irregular plasmolysis was observed in the plumular axis cells, the nuclear membrane was wrinkled, and the size of the cell contents such as lipid bodies and protein bodies did not differ significantly from that of the non-dehydrated cells (Figure 2B). After 24 h-dehydration, air bubbles appeared in the cell wall; the nucleus was in a regular spherical shape; the lipid bodies were highly fused; the nuclear membrane shrank inward; the structure of the mitochondrial was disordered; and the lipid bodies were broken and deformed (Figure 2D). After 2 h-dehydration, the seeds were stored in liquid nitrogen for 24 h. It was observed that (Figure 2C), compared to pre-cryopreservation (Figure 2B), the space between the cells was bigger; the protein bodies were evenly distributed around the cell wall; the nuclear membrane and cell membrane shrunk more severely; the nucleolus was reduced and regular spherical shape; and the lipid bodies were highly fused. After 24 h-dehydration whereafter by cryopreservation, the cell wall collapsed severely; vacuoles appeared between the cells; the lipid bodies ruptured and deformed; the cell wall ruptured; contents overflowed; and the structures such as the endoplasmic reticulum could not be identified (Figure 2E).
Combined with Figure 2A,B,D, it can be seen that, after dehydration treatment, a large number of small lipid bodies fused into larger cellulites, and cryopreservation treatment made the fusion degree of lipid bodies higher, according to combinations Figure 2B,C and Figure 2D,E.

2.3. Response of Antioxidant Enzyme Activities in S. tonkinensis Seed Plumular Axis Treated with Dehydration and Dehydration Followed by Cryopreservation Treatment

The superoxide dismutase (SOD) enzyme activity in S. tonkinensis seed plumular axis showed an overall increasing trend with the dehydration time increased, reaching maximum at 2 or 8 h-dehydration. The SOD activity after cryopreservation reached maximum at 2 or 24 h-dehydration, and the difference was not significant before and after cryopreservation at 2 h-dehydration (Figure 3A). Catalase (CAT) enzyme was the most important antioxidant enzyme responsible for scavenging excess H2O2 in cells. After dehydration and dehydration followed by cryopreservation, the CAT activity showed an increasing trend with increasing dehydration time. It is worth noting that the CAT activity was greatly increased after 2h-dehydration followed by cryopreservation. After 24h-dehydration followed by cryopreservation, the CAT activity was significantly reduced, which was consistent with the trend of seed germination percentage (Figure 3B). The activity of the APX enzyme in the plumular axis of S. tonkinensis seed exhibited an initial increase and subsequently decreased with increasing dehydration time. However, the overall trend increased after the dehydration followed by cryopreservation treatment. The APX activity had a peak at 2 h-dehydration, and there was no significant difference before and after cryopreservation treatment. It was noted that, after 24 h-dehydration, APX enzyme treated with cryopreservation increased rapidly, contrary to the trend of CAT activity (Figure 3C).

2.4. Results of iTRAQ Analysis and Differential Protein Screening of S. tonkinensis Seeds Plumular Axis Treated with Dehydration and Dehydration Followed by Cryopreservation Treatment

To investigate the molecular response pathways of S. tonkinensis seeds plumular axis treated with dehydration and cryopreservation, all samples were analyzed by iTRAQ proteomics. The iTRAQ proteomics produced a total of 978,057 spectra from the 4 sets of samples. We identified 97,996 known spectra, 66,953 unique spectra, 12,530 peptides, 9568 unique peptides, and 3898 proteins, respectively (Table S1 and Figure S1). Based on previous studies [38], differentially expressed proteins (DEPs) were defined based on a 1.2–1.5 fold change threshold. Among the proteins, only one protein in two or three replicates with fold changes ≥1.2 was defined up-regulated or ≤0.83 as down-regulated. Based on above criteria, 41, 45, 98, 40 DEPs were identified in D2 vs. CK, CD2 vs. CK, D24 vs. CK, and CD24 vs. CK, respectively (Figure 4A; Table S2), using the control group as reference to compare the DEPs of samples. Among them, there were 7 DEPs in all comparison groups, and there were largest number of unique DEPs with 48 in D24 vs. CK (Figure 4B).

2.5. KEGG and Go Functional Annotation of DEPs in S. tonkinensis Seeds Axis Treated with Dehydration and Cryopreservation

GO (Gene Ontology) and KEGG (Kyoto Encyclopedia of Genes and Genomes) enrichment analyses were performed on the DEPs to investigate functions and biological pathways during dehydration and post-dehydration cryopreservation treatment. GO functional enrichment analysis showed that the DEPs before and after preservation cryopreservation significantly enriched in the structural molecule activity were structural molecule activity (GO:0005198), intracellular (GO:0005622), gene expression(GO:0010467) and intracellular ribonucleoprotein complex (GO:0030529)(Figure 5B,D). Notably, GO:0016209 antioxidant activity was significantly enriched in 2 and 24 h-dehydration but not in cryopreservation (Figure 5A–D, Tables S3–S6). To gain insight into the metabolic pathways of DEPs, scatter plots were used to show the top 10 KEGG enrichment pathways, which showed that DEPs were significantly enriched in ribosome (ko03010), carbon fixation in photosynthetic organisms (ko00710), and alpha-linolenic acid metabolism (ko00592) pathways after 2 h-dehydration (Figure 5E, D2 vs. CK). Compared with 2 h-dehydration, the DEPs of 24 h-dehydration were significantly enriched not only in ribosome (ko03010) but also in glutathione metabolism, protein processing in endoplasmic reticulum (ko04141), and spliceosome pathways (Figure 5G, D24 vs. CK). In 2 h-dehydration followed by cryopreservation seeds, DEPs were mainly enriched in ribosome (ko03010) and protein processing in endoplasmic reticulum (ko04141) (Figure 5F, CD2 vs. CK). DEPs of 24 h-dehydration followed by cryopreservation seeds were mainly enriched in ribosome (ko03010), proteasome, and flavonoid Biosynthesis (Figure 5H, CD24 vs. CK). The physiological processes of dehydration and post-dehydration cryopreservation treatment were very similar and were both involved in ribosome, protein processing in endoplasmic reticulum, antioxidant, and signal transduction activities.
To investigate the relationship between identified key proteins, 140 DEPs were predicted using Protein–Protein Interaction Networks. The STRING online software was used to analyze the protein interaction regulatory network; among them, 68 DEPs interacted with each other (Figure 6), and the other proteins formed independent branches. GO functional analysis showed that the molecular functions of DEPs were mainly enriched in structural molecule activity, antioxidant activity, carbon–sulfur lyase activity, endopeptidase activity, intramolecular transferase activity, transferring amino groups, oligosaccharyl transferase activity, and peroxidase activity pathways and involved in the biological process of macromolecule metabolic process and response to stress (Table S7). KEGG functional annotation showed that the DEPs were mainly involved in proteasome, protein processing in endoplasmic reticulum, RNA translation, oxidation-reductive process, and ribosome, etc. (Figure 6). However, the proteins involved in energy metabolism were not included in the network. PPI network analysis suggested that redox balance, nucleic acid metabolism, protein processing, and modification process played important roles during dehydration and cryopreservation. Interestingly, dehydration-related proteins such as LEA4, which was down-regulated in this network, and its homologous proteins SMPs (seed maturation proteins, PM41, LEA4), which were up-regulated at 2 h-dehydration and had no difference at 24 h-dehydration, were found to form an independent network (Figure 6).

3. Discussion

3.1. Molecular Mechanism of Dehydration Tolerance in S. tonkinensis Seeds

According to their storage behavior, seeds were classified into three types: orthodox seeds, intermediate seeds, and recalcitrant seeds [39,40]. DT and low temperature tolerance are the key characteristics of non-orthodox seeds, including recalcitrant seeds and intermediate seeds, to achieve long-term safe storage, which has important theoretical and practical significance. In this study, the fresh and mature S. tonkinensis seeds had a water content of 31% and demonstrated a certain degree of dehydration tolerance. Even after being dehydrated to 10.1% water content, the germination percentage remained at 83.0%. However, upon further dehydration to 5.4%, the germination percentage reduced to 30.5% (Figure 1). Although S. tonkinensis seeds could survive under relatively low water content, they could not tolerate more water loss as orthodox seeds. This storage behavior was similar to other intermediate seeds such as Carica papaya L. and Pistia stratiotes L., etc. Therefore, according to the existing definition of storage behavior [41], S. tonkinensis seeds should be classified as intermediate seeds.
The proteomics results showed that the expression of phosphoinositide phospholipase C was up-regulated both by dehydration and cryopreservation treatments. The electron microscopy sections revealed that there was a large number of lipid bodies that fused into large cellulites after dehydration treatment in plumular axis, and the degree of fusion increased with longer dehydration time. During cryopreservation, the medium outside the seed cells froze, whilst the cells have not yet frozen. With the increasing extracellular ice crystals, the cells lost water and gradually became dehydrated due to the vapor pressure difference. In this study, the degree of lipid bodies fusion was increased with cryopreservation treatment under the same dehydration time (Figure 2). It was reported that that oil seeds had higher DT compared to other seeds, and that high lipid body content was the most significant feature of oil seeds [42,43,44,45]. For example, rape seeds could maintain seed vigor when the water content was reduced to less than 1.5% [46]. Therefore, it is speculated that the presence of a large number of lipid bodies in the plumular axis is one of the important reasons for a certain degree of dehydration tolerance in S. tonkinensis seeds.
The activity of ROS scavenging system of seeds was closely related to seed recalcitrance production [47,48,49]. The activity of antioxidant enzymes of S. tonkinensis seeds was enhanced after 0~8 h-dehydration and cryopreservation (Figure 3), indicating that plant defense system generated self-help measures to enhance the activity of antioxidant enzymes to fight against ROS free radicals produced by dehydration and low temperature stress. SOD activity during dehydration and post-dehydration cryopreservation treatment of S. tonkinensis seeds gradually increased, consistent with the results of the study on dehydration tolerance of Citrus wilsnii Tanaka seeds [50]. Cryopreservation treatment exerted the greatest promoting effect on APX enzyme activity, and the longer the dehydration time was, the higher the APX activity was, similar to the pattern of Sanionia uncinata (Hedw.) Loeske reported by Pizarro et.al. [51]. The proteomic results showed that the expression of antioxidant enzymes such as peroxidase (POD), glutathione peroxidase (GPX), glutathione S-transferase (GST), CAT, and SOD were up-regulated during dehydration and post-dehydration cryopreservation (Table S2), similar to the change trend of enzyme activities. On the one hand, it suggested the reliability of proteomic results. On the other hand, it provided a speculation that the expression and enzyme activities of antioxidant enzymes were increased to fight and scavenge ROS free radicals during dehydration and post-dehydration cryopreservation treatment, which may be another important reason for DT of S. tonkinensis seeds.
In this study, the protein processing in the endoplasmic reticulum pathway showed different enriched degrees in multiple treated samples (Figure 5). In this pathway, the S-Phase Kinase Associated Protein 1 (SKP1) was down-regulated at all stages, while the rest of the proteins were up-regulated at different stages. The endoplasmic reticulum played an important role in the folding, assembly, transport, secretion, and glycosylation of protein synthesis [52]. In the normal course of growth, molecular chaperones and glycosylation modifications assisted proteins in folding correctly [53]. Once stimulated, a large number of misfolded and unfolded proteins were produced and accumulated, resulting in endoplasmic reticulum stress (ERS) due to the inability of the precise endoplasmic reticulum quality control system (ERQC) system to process in time, which in turn triggered an unfolded protein response (UPR), and if a series of regulated downstream genes still failed to restore the plant endoplasmic reticulum homeostasis, UPR would initiate programmed cell death [52]. HSP was an important anti-stress-protective protein of protein processing in the endoplasmic reticulum pathway, with molecular chaperone properties, and maintained the normal function through various biological processes such as protein synthesis and degradation, stabilization of protein structure, and cellular localization [54]. In this study, three HSPs were found to be up-regulated during 2 h-dehydration followed cryopreservation treatment, and one HSP was up-regulated in all treatments (Table S2). Notably, three PDIs (protein disulfide-isomerase) were up-regulated after 24 h-dehydration in this pathway, and studies have shown that up-regulation of PDI protein expression increased seed storage tolerance during development and abiotic stress [55,56,57], suggesting that PDI may increase storage tolerance by removing the wrong proteins in the endoplasmic reticulum, thereby regulating seed DT and low temperature tolerance [58]. In addition, the thioredoxin-like superfamily protein precursor was up-regulated in all treatments. The redox state of disulfide bonds played an important role in maintaining the correct folding and advanced structure of proteins, and excessive ROS would affect the equilibrium state of disulfide bonds between proteins, leading to structural changes and inactivation of proteins [59]. Thioredoxin (TRX) contained redox-active disulfide bonds, which could regulate redox in cells through reversible conversion of thiol-disulfide bonds and stabilize the structure and function of proteins [60]. It is hypothesized that S. tonkinensis seeds are subject to protein misfolding under stress conditions during dehydration, and the identified DEPs of endoplasmic reticulum-associated proteins are involved in the repair of misfolded protein macromolecules. This could potentially be another important factor for DT of S. tonkinensis seeds.
A reduction in metabolic activity was a characteristic of DT [61]. In this study, it was observed that DEPs from all samples were significantly enriched in the ribosomal pathway (Figure 6), and the expression of ribosomal proteins was up-regulated (Table S2). Additionally, a group of up-regulated DEPs with molecular functions related to carbon-sulfur lyase activity, endopeptidase activity, proteasome, RNA translation, intramolecular transferase activity, and other GO terms were identified (Figure 5A–D). It is speculated that high metabolic activity could be one of the important factors contributing to the intermediate storage behavior of S. tonkinensis seeds.
The accumulation of LEA proteins and non-reducing raffinose family oligosaccharides was closely related to DT formation [2]. During the dehydration stage, LEAs proteins form a protective aqueous membrane around the internal cell structure and macromolecules, thus making the seeds DT due to their hydrophilic quality [62]. The 126, 123, and 167 LEA proteins were identified in normal seeds of Arachis hypogaea L., Arachis hypogaea L., and Triticum aestivum L., respectively [63,64,65]. Li et al. identified 50 LEA proteins in Linum usitatissimum L., of which 36 LuLEA proteins were expressed in seed maturation [66]. Shen et al. identified 20 LEA proteins in lotus embryo, of which 7 protein families were identified in the dehydration maturation stage [67]. At the same time, Zhang et al. identified 12 LEA proteins by proteomics, 9 of which belong to the LEA protein family [68]. However, non-orthodox seeds were different. For example, 39 LEA genes were identified in the genome of tea tree, and only 6 CsLEA genes were involved in seed dehydration, and most of them were repressed [69]. Julien et al. compared the results of LEA polypeptide assays during dehydration in orthodox seeds of Medicago truncatula Gaertn. and recalcitrant seeds of Castanospermum australe A. Cunn. et Fraser cotyledons and found that transcripts of 16 homologues out of 17 LEA genes for which polypeptides were detected in orthodox seeds. For the 12 LEA genes, however, the polypeptide was either absent or strongly reduced in recalcitrant seeds [70]. In this study, only 17 LEA protein polypeptide segments were identified, belonging to the 5 subfamilies of LEA2, SMP, dehydrin, LEA4, and LEA5. However, only two homologous SMP proteins were up-regulated during 2 h-dehydration, and one Lea4 protein was down-regulated during seed dehydration and post-dehydration cryopreservation (Figure 6). The above results indicate that the number of LEA proteins and families differed greatly between orthodox and non-orthodox seeds [26]. During the dehydration process, a large amount of oligosaccharides and non-reducing sugars were accumulated to maintain the stability of membranes and proteins [71]. However, KEGG functional annotation showed that the proteins involved in the pathway of glucose metabolism were not included in the network (Figure 6), and there were no changes in the synthesis of proteins related to oligosaccharides. It is speculated that the lack of sufficient LEA, dehydrin, and oligosaccharides is one of the important factors for the intermediateness of S. tonkinensis seeds.

3.2. Cryopreservation Treatment May Be Used as a Candidate Method for the Preservation of S. tonkinensis Seeds

So far, recalcitrant seeds have not been effectively preserved on a large scale for a long time, and cryopreservation was considered to be the most promising research direction [72,73]. The DT and low temperature tolerance of intermediate seeds were between orthodox seeds and recalcitrant seeds, with limited DT, while dehydration treatment did not prolong storage life of intermediate seeds, and intermediate seeds were sensitive to low temperature, which could not be stored for a long time through traditional low-temperature seed banks [40]. Cryopreservation was an ideal method for long-term preservation of plant germplasm resources [74]. Cryopreservation was usually carried out in liquid nitrogen at −196 °C. Theoretically, the metabolism and growth activities of plant seeds were in a relatively stable biological state; cells, tissues, and organs did not undergo genetic mutations and trait changes during cryopreservation [75], and various physiological indicators have shown that seed activity was not affected, such as Phaius tankervilleae (L’Heritier) Blume seeds [76] and Tillandsia species seeds [77]. Cryopreservation safely preserved pollen and young shoots so that their activity was not affected. Pradeep et al. [78] found that the germination percentage of guava pollen was not significantly different from that of fresh state after cryopreservation, thus overcoming the problem of asynchronous flowering. Thomas Rasl reported that the regeneration percentage of winter-acclimated buds of Dracocephalum austriacum L. after cryopreservation reached up to 100% [79].
In the process of cryopreservation, the formation of ice crystals and the change of osmotic pressure caused damage and wrinkling of cell membranes, which in turn led to changes in cell and tissue morphology and function. A number of experiments have shown that damage to cell membranes and organelles was strongly related to the formation of intracellular ice crystals [80]. Therefore, some non-orthodox seed crops, such as mango and lychee, were difficult to preserve through cryopreservation. An important reason was that the water content in biological cells was relatively high; freezing occurs during the cooling process, causing damage and destruction to the cell structure, eventually leading to cell death [81]. Due to the sensitivity to dehydration, the water content of non-orthodox seeds was reduced to a safe threshold to survive freezing damage, which was a key factor before the safe cryopreservation [82,83,84].
The germination percentage of Senna tora (L.) Roxb. seeds decreased significantly with decreasing the water content with or without liquid nitrogen preservation, and its germination percentage was less than 20% at 9.3% water content [85]. In contrast, the seed germination percentage of Mimosa pudica L. seeds was more than 70% when the water content was between 8.09~15.67%, which was non-significantly different before and after liquid nitrogen preservation [85]. Similar to the phenotype of M. pudica seeds, the germination percentage of S. tonkinensis seeds in this study was not significantly different from the control group at 71.2~83% after cryopreservation when the water content was 8.5~10.1% and reached the highest germination rate of 83% at 2 h-dehydration. Since the germination percentage of seeds dehydrated to 5.4% water content was only 30.5%, without preservation value, no cryopreservation treatment was performed. The results of antioxidant enzyme activity identification showed that the SOD activity after cryopreservation reached maximum at 2 h-dehydration, and the difference was not significant before and after cryopreservation. The CAT activity was greatly increased after 2 h-dehydration followed by cryopreservation and reached maximum. The APX activity had a peak at 2 h-dehydration, and there was no significant difference before and after cryopreservation treatment. The physiological processes of dehydration and post-dehydration cryopreservation treatments were similar, both involving protein processing in the ribosome, endoplasmic reticulum, antioxidants, and signal transduction, indicating that cryopreservation treatment had no effect on seeds. In summary, it is concluded that S. tonkinensis seeds can be preserved by cryopreservation technology, and the most suitable dehydration time for cryopreservation is 2 h with a water content of 10.1%.

4. Materials and Methods

4.1. Experimental Materials

The seeds of S. tonkinensis used in the experiment were collected in October 2022 from Shishan District, Southwest Guilin, Guangxi (E: 110°16’, N: 25°15). Pick the fresh and mature S. tonkinensis seeds together with the pods, peel the seeds from the pods, place the peeled seeds in a cool and dry space, dry them to constant weight, and then put them in the incubator at 5 °C for future use.
The seeds were quickly dehydrated with allochroic silicagel [86]. Seeds with dehydration time of 1, 2, 4, 8, and 24 h were selected and stored in liquid nitrogen for 24 h. Part of seeds were thawed, and the embryos were peeled out; some embryos were fixed with glutaraldehyde fixative for subcellular structure microscopy, and the rest of embryos were ground into powder for protein content, antioxidant enzyme activity determination, and proteomic analysis. The other seeds were subjected to germination experiments.

4.2. Germination Experiment

Fresh mature seeds were germinated in incubators at 10 °C, 15 °C, 20 °C, 25 °C, 30 °C, and 35 °C, and the highest germination percentage was found at 25 °C (Figure S2). Fresh mature seeds, seeds dehydrated to various water contents, and seeds cryopreserved after dehydration were used as materials for the germination experiment. Perlite was used as the germination medium, and 50 seeds were sown in 4 replicates for each treatment. The seeds were incubated in an incubator at 25 °C for 30 d (12 h/day alternating light or total darkness, PPFD = 121 µmol m−2s−1, and the germination state was checked under a green light in total darkness). The survival rate after dehydration treatment and cryopreservation was expressed as the survival rate of seedlings. The data, which were shown as the means ± SDs, were subjected to ANOVA to determine significant differences.

4.3. Cell Structure and Antioxidant Enzyme Activity of Seedplumular Axis Treated with Cryopreservation after Dehydration and Dehydration Treatment

The seed plumular axis were washed with water, cut into small cubes of 1 × 1 mm, and placed in the prepared fixative solution (3% glutaraldehyde, 0.1 M PBS, pH 7.0). After rinsing, the phosphate buffer was removed and put into 1% H2[OsO4(OH)2] fixation for 8 h and was then dehydrated with ethanol with a volume fraction gradient of 30~100% for 30 min. After removing the ethanol, infiltration, and embedding, the samples were transferred to epoxy resin for polymerization at 60 °C for 24 h. After the processing, the sample was placed on an ultramicrotome for sectioning with a thickness of 50 nm. After fixing, it was observed and photographed by a transmission electron microscope.
The SOD was measured by the nitro blue tetrazolium (NBT) method [87]. The activity of CAT was measured according to the methods of Maehly and Chance [88], and the activity of APX was measured according to the methods of Hossain and Asada [89]. All assays described above were set up with four biological replicates. The data, which were shown as the means ± SDs, were subjected to t-test to determine significant differences.

4.4. Protein Extraction and Enzyme Hydrolysis

Fresh seeds (CK), seeds for 2 h-dehydration (D2), seeds for cryopreservation for 24 h in liquid nitrogen after a 2 h-dehydration (CD2), seeds for 24 h-dehydration (D24), and seeds for cryopreservation for 24 h in liquid nitrogen after 24 h-dehydration (CD24), with three repeat samples in each group, were respectively homogenized to powder in liquid nitrogen. The powder was resuspended in a lysis buffer (7 M urea, 2 M thiourea, 4% SDS, 40 mM Tris-HCl), with a final concentration of 1 mM PMSF and 2 mM EDTA. After mixing, the samples were added to dithiothreitol (DTT) to a final concentration 10 mM then sonicated at 200 W for 15 min. The supernatant was obtained following centrifugation at 13,000× g for 20 min at 4 °C and was added to 4 fold volume of ice-cold acetone and precipitated at −20 °C overnight. After centrifugation as described above, the pellets were dissolved in 400 mL, 100 mM TEAB, and 7 M urea. The supernatant was reduced with 10 mM DTT and a water bath for 30 min at 56 °C. Next, the reaction was blocked with 55 mM IAM at dark room temperature for an alkylation reaction lasting 30 min. The homogenates were centrifuged, and the resulting supernatant protein samples were stored at −80 °C. The protein concentration was quantified by Coomassie (Bradford) Protein Assay Kit (Thermo Scientific, Waltham, MA, USA. Bovine albumin (1000 μg/mL) was used as the standard. A total of 100 μg protein from each sample was used for trypsin digestion. Briefly, trypsin was added to protein solution, which was diluted 5 times with 100 mM TEAB at 1:50 and hydrolyzed overnight at 37 °C. The peptides hydrolyzed by the enzyme were desalted by C18 column and then desalinated by vacuum freeze-drying.

4.5. iTRAQ Labeling and Grouping

Samples were dissolved with 0.5 M triethylammonium bicarbonate. According to the specification of the iTRAQ-8 marker kit, the samples were marked and mixed. The proteins from each sample were labeled with iTRAQ reagents 113, 114, 115, and 116, respectively. The labeled samples were pooled and purified using a strong cation exchange chromatography column (Phenomenex, Torrance, CA, USA) and separated by liquid chromatography (LC) using the Ultimate 3000 HPLC system, USA). Additionally, the chromatographic column used was Durashell C18 (5 μm, 4.6 × 250 mm). The peptide was separated by increasing an can concentration under alkaline conditions, and the flow rate was 1 mL min−1. A total of 42 secondary components were collected and combined into 12 components. The combined components were desalted on the Strata-X column and dried under vacuum. A triple TOF 5600 liquid mass spectrometry system and liquid chromatography-mass spectrometry (LC-MS) were used for mass spectrometry data acquisition. Briefly, the polypeptide sample was dissolved in 2% acetonitrile 0.1% formic acid and analyzed by a Triple TOF 5600 plus mass spectrometer coupled with an Eksigent nano LC system. The polypeptide solution was added to the C18 capture column (5 μm, 100 μm × 20 mm) and eluted at a 90 min time gradient and at a flow rate of 300 nL min−1 on the C18 column (3 μm, 75 μm × 150 mm). For IDA (information-dependent collection), the primary mass spectrometry of 30 precursor ions was scanned with an ion accumulation time of 250 ms, and the secondary mass spectrometry of 30 precursor ions was collected with the ion accumulation time of 50 ms. The MS1 spectra were collected in the range of 350 ≤ 1500 m.z., and the MS2 spectra were collected in the range of 100 ≤ 1500 m.z. The dynamic exclusion time of precursor ions was set to 15 s.
In this experiment, peptides were identified using ProteinPilotTM 4.5 (https://en.freedownloadmanager.org/; accessed on 25 April 2022), a search engine matching with AB Sciex 5600 plus (SCIEX, Framingham, MA, USA), with the following paragon parameters. For the identification results of ProteinPilot, we conducted further filtration using unused scores. We chose the peptides with unused scores ≥1.3, which means that the reliability level is more than 95%. The filtrated proteins were used in the subsequent analysis.

4.6. Database Search and Proteomic Analysis

Protein identification was performed using ProteomeDiscovererTM 2.2 aligned to the NCBI NR database (http://www.ncbi.nlm.nih.gov/; accessed on 18 January 2022) and SwissProt/UniProt database (http://www.uniprot.org/; accessed on 18 January 2022) by DIAMOND. One protein had to contain at least one unique peptide. Peptide false discovery rate (FDR) analysis ≤0.01 was the filtering parameter. DEPs were identified using a FC (fold change) threshold, and p-values of < 0.05 were significant. FC > 1.2 was considered up-regulated, and FC < 0.83 was considered down-regulated.
The COG analysis of the proteins was performed for functional classification via searches in a database (http://www.ncbi.nlm.nih.gov/COG/; accessed on 18 March 2022). Blast2GO 2.5.0 was used for the GO annotations (http://www.geneontology.org/; accessed on 18 March 2022) of the DEPs to classify proteins based on molecular function, biological process, and cellular components. Blast2GO computes Fisher’s Exact Test by applying robust FDR correction for multiple testing and returns a list of significant GO terms ranked by their corrected or one-test p-values. KEGG (http://www.genome.jp/kegg/; accessed on 2 April 2022) annotation was performed using KOBAS 2.1.1 software to predict the metabolic pathways. Protein–protein interactions (PPI) were analyzed using the STRING v10 database (http://string-db.org; accessed on 4 February 2022) to determine the identified proteins.

5. Conclusions

In this study, we reported the dehydration characteristics and viability of cryopreservation for S. tonkinensis seeds. The results showed that these seeds possessed a certain degree of dehydration tolerance. The highest germination percentage reached 83.0% upon dehydration to a 10.1% water content. However, when the water content dipped below 9.7%, the germination percentage decreased significantly. At a mere 5.4% water content, the germination percentage was only 30.5%. Consequently, the S. tonkinensis seeds are best categorized as intermediate seeds.
Subsequent analysis indicated that the lipid bodies gradually fused during dehydration, and the degree of fusion was significantly improved by cryopreservation. Physiological experiments demonstrated that SOD, POD, and CAT in the antioxidant enzyme system increased during dehydration, aligning with the enrichment tendency of DEPs in proteomics results regarding ROS clearance pathway and stress defense. Proteomics outcomes also unveiled significant DEP enrichment in protein processing in endoplasmic reticulum and ribosomal pathway but not in intracellular energy metabolism and metabolic activities.
It is postulated that S. tonkinensis seeds form an anti-stress protection mechanism mainly by protein processing in the endoplasmic reticulum and the antioxidant system during dehydration, which is related to their dehydration tolerance. In addition, DEPs were significantly enriched in the ribosome pathway; only three differentially expressed LEA proteins were identified, and it is speculated that the lack of specific LEA and dehydrins is also one of the important factors for the reduced germination percentage after excessive dehydration. There was no significant difference in protein enrichment pathway between dehydrated seeds and 2 h-dehydrated followed by cryopreservation, suggesting that the cryopreservation of S. tonkinensis seeds is indeed feasible.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/plants12091842/s1, Figure S1: Total spectra, spectra, unique spectra, peptides, unique peptide, and proteins identified from iTRAQ proteomic analysis; Figure S2: Germination process of fresh mature Sophora tonkinensis seeds under constant temperature and alternating light/dark conditions; Table S1: Functional annotation of all proteins; Table S2: Annotation of all DEPs; Table S3: Go functional annotation of DEPs in D2 vs. CK; Table S4: Go functional annotation of DEPs in CD2 vs. CK; Table S5: Go functional annotation of DEPs in D24 vs. CK; Table S6: Go functional annotation of DEPs in D24 vs. CK; Table S7: Go functional annotation of PPI interaction network for DEPs.

Author Contributions

Y.L. participated in performing the experiments, data analysis, drafting and revising the manuscript; Y.Z. participated in performing the experiments, data analysis, drafting the manuscript; Y.J. analyzed data and revised the manuscript; Z.D. (Zhangyan Dai) participated in revising the manuscript; Q.L. and J.M. performed the experiment and analyzed the data, respectively; L.X., S.D., J.L. (Jitao Li) and R.W. participated in performing the experiments. Z.D. (Zhijun Deng) participated in conceiving the study, data analysis, and revising the manuscript. J.L. (Jun Liu) participated in conceiving the study and revising the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported the Science and Technology Program of Guangdong Province, China (2022B0202110003, 2020B121201008, 2019KJ106), the National Natural Science Foundation of China (31860073, 31871716), and the Foundation of Guangdong Academy of Agricultural Sciences (202132TD).

Data Availability Statement

The data is contained within the manuscript and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Smolikova, G.; Leonova, T.; Vashurina, N.; Frolov, A.; Medvedev, S. Desiccation Tolerance as the Basis of Long-Term Seed Viability. Int. J. Mol. Sci. 2021, 22, 101. [Google Scholar] [CrossRef]
  2. Song, S.-Q.; Liu, J.; Tang, C.-F.; Cheng, H.-Y.; Wang, W.-Q.; Zhang, Q.; Zhang, W.-H.; Gao, J.-D. Research Progress on the Physiology and Its Molecular Mechanism of Seed Desiccation Tolerance. Sci. Agric. Sin. 2022, 55, 1047–1063. [Google Scholar]
  3. Zhongqi, L.; Jiwai, H.; Haiqing, Z.; Aimin, L. Dehydration Tolerance of Plant Seeds: Current Research Situation and Prospects. Chin. Agric. Sci. Bull. 2020, 36, 36. [Google Scholar] [CrossRef]
  4. Bewley, J.D.; Bradford, K.J.; Hilhorst, H.W.M.; Nonogaki, H. Seeds: Physiology of Development, Germination and Dormancy, 3rd ed.; Springer: New York, NY, USA, 2013; ISBN 978-1-4614-4692-7. [Google Scholar]
  5. McKersie, B.D. Principles of Seed Science and Technology. Plant Sci. 2002, 162, 849. [Google Scholar] [CrossRef]
  6. Pammenter, N.W.; Berjak, P. Physiology of Desiccation-Sensitive (Recalcitrant) Seeds and the Implications for Cryopreservation. Int. J. Plant Sci. 2014, 175, 21–28. [Google Scholar] [CrossRef]
  7. Dekkers, B.J.W.; Costa, M.C.D.; Maia, J.; Bentsink, L.; Ligterink, W.; Hilhorst, H.W.M. Acquisition and Loss of Desiccation Tolerance in Seeds: From Experimental Model to Biological Relevance. Planta 2015, 241, 563–577. [Google Scholar] [CrossRef]
  8. Berjak, P.; Pammenter, N.W. From Avicennia to Zizania: Seed Recalcitrance in Perspective. Ann. Bot. 2007, 101, 213–228. [Google Scholar] [CrossRef]
  9. Berjak, P. Unifying Perspectives of Some Mechanisms Basic to Desiccation Tolerance across Life Forms. Seed Sci. Res. 2006, 16, 1–15. [Google Scholar] [CrossRef]
  10. Pammenter, N.W.; Berjak, P. A Review of Recalcitrant Seed Physiology in Relation to Desiccation-Tolerance Mechanisms. Seed Sci. Res. 1999, 9, 13–37. [Google Scholar] [CrossRef]
  11. Leprince, O.; Pellizzaro, A.; Berriri, S.; Buitink, J. Late Seed Maturation: Drying without Dying. EXBOTJ 2016, 68, 827–841. [Google Scholar] [CrossRef]
  12. Sano, N.; Rajjou, L.; North, H.M.; Debeaujon, I.; Marion-Poll, A.; Seo, M. Staying Alive: Molecular Aspects of Seed Longevity. Plant Cell Physiol. 2016, 57, 660–674. [Google Scholar] [CrossRef] [PubMed]
  13. Baud, S.; Dubreucq, B.; Miquel, M.; Rochat, C.; Lepiniec, L. Storage Reserve Accumulation in Arabidopsis: Metabolic and Developmental Control of Seed Filling. Arab. Book 2008, 6, e0113. [Google Scholar] [CrossRef] [PubMed]
  14. Marques, A.; Buijs, G.; Ligterink, W.; Hilhorst, H. Evolutionary Ecophysiology of Seed Desiccation Sensitivity. Funct. Plant Biol. 2018, 45, 1083–1095. [Google Scholar] [CrossRef] [PubMed]
  15. Lepiniec, L.; Devic, M.; Roscoe, T.J.; Bouyer, D.; Zhou, D.-X.; Boulard, C.; Baud, S.; Dubreucq, B. Molecular and Epigenetic Regulations and Functions of the LAFL Transcriptional Regulators That Control Seed Development. Plant Reprod. 2018, 31, 291–307. [Google Scholar] [CrossRef]
  16. Braybrook, S.A.; Stone, S.L.; Park, S.; Bui, A.Q.; Le, B.H.; Fischer, R.L.; Goldberg, R.B.; Harada, J.J. Genes Directly Regulated by LEAFY COTYLEDON2 Provide Insight into the Control of Embryo Maturation and Somatic Embryogenesis. Proc. Natl. Acad. Sci. USA 2006, 103, 3468–3473. [Google Scholar] [CrossRef] [PubMed]
  17. Song, S.-Q.; Tian, M.-H.; Kan, J.; Cheng, H.-Y. The Response Difference of Mitochondria in Recalcitrant Antiaris Toxicaria Axes and Orthodox Zea Mays Embryos to Dehydration Injury. J. Integr. Plant Biol. 2009, 51, 646–653. [Google Scholar] [CrossRef] [PubMed]
  18. Huang, H.; Møller, I.M.; Song, S.-Q. Proteomics of Desiccation Tolerance during Development and Germination of Maize Embryos. J. Proteom. 2012, 75, 1247–1262. [Google Scholar] [CrossRef]
  19. Carbonero, P.; Iglesias-Fernández, R.; Vicente-Carbajosa, J. The AFL Subfamily of B3 Transcription Factors: Evolution and Function in Angiosperm Seeds. J. Exp. Bot. 2017, 68, 871–880. [Google Scholar] [CrossRef]
  20. Fatihi, A.; Boulard, C.; Bouyer, D.; Baud, S.; Dubreucq, B.; Lepiniec, L. Deciphering and Modifying LAFL Transcriptional Regulatory Network in Seed for Improving Yield and Quality of Storage Compounds. Plant Sci. 2016, 250, 198–204. [Google Scholar] [CrossRef] [PubMed]
  21. Sall, K.; Dekkers, B.J.W.; Nonogaki, M.; Katsuragawa, Y.; Koyari, R.; Hendrix, D.; Willems, L.A.J.; Bentsink, L.; Nonogaki, H. DELAY OF GERMINATION 1-LIKE 4 Acts as an Inducer of Seed Reserve Accumulation. Plant J. 2019, 100, 7–19. [Google Scholar] [CrossRef] [PubMed]
  22. Carrillo-Barral, N.; del Rodríguez-Gacio, M.C.; Matilla, A.J. Delay of Germination-1 (DOG1): A Key to Understanding Seed Dormancy. Plants 2020, 9, 480. [Google Scholar] [CrossRef]
  23. Dekkers, B.J.W.; He, H.; Hanson, J.; Willems, L.A.J.; Jamar, D.C.L.; Cueff, G.; Rajjou, L.; Hilhorst, H.W.M.; Bentsink, L. The Arabidopsis DELAY OF GERMINATION 1 Gene Affects ABSCISIC ACID INSENSITIVE 5 (ABI5) Expression and Genetically Interacts with ABI3 during Arabidopsis Seed Development. Plant J. 2016, 85, 451–465. [Google Scholar] [CrossRef] [PubMed]
  24. Michalak, M.; Barciszewska, M.Z.; Barciszewski, J.; Plitta, B.P.; Chmielarz, P. Global Changes in DNA Methylation in Seeds and Seedlings of Pyrus Communis after Seed Desiccation and Storage. PLoS ONE 2013, 8, e70693. [Google Scholar] [CrossRef]
  25. Bartels, A.; Han, Q.; Nair, P.; Stacey, L.; Gaynier, H.; Mosley, M.; Huang, Q.Q.; Pearson, J.K.; Hsieh, T.-F.; An, Y.-Q.C.; et al. Dynamic DNA Methylation in Plant Growth and Development. Int. J. Mol. Sci. 2018, 19, 2144. [Google Scholar] [CrossRef]
  26. Radwan, A.; Hara, M.; Kleinwächter, M.; Selmar, D. Dehydrin Expression in Seeds and Maturation Drying: A Paradigm Change. Plant Biol. J. 2014, 16, 853–855. [Google Scholar] [CrossRef] [PubMed]
  27. Nonogaki, H. Seed Germination and Dormancy: The Classic Story, New Puzzles, and Evolution. J. Integr. Plant Biol. 2019, 61, 541–563. [Google Scholar] [CrossRef]
  28. Chen, D.; Chen, B.; Fang, Y.; Zheng, C.; Ding, C.; Li, J.; Ma, Q. Flora of China; Science Press: Beijing, China, 2018; Volume 40, p. 76. [Google Scholar]
  29. Zhang, S.; Li, X.; Yang, W.; Zhou, Y. Sophorae Tonkinensis Radix et Rhizome-Induced Pulmonary Toxicity: A Study on the Toxic Mechanism and Material Basis Based on Integrated Omics and Bioinformatics Analyses. J. Chromatogr. B 2021, 1179, 122868. [Google Scholar] [CrossRef]
  30. Zhang, S.-N.; Li, X.-Z.; Tan, L.-Y.; Zhu, K.-Y. A Review of Pharmacological and Toxicological Effects of Sophora Tonkinensis with Bioinformatics Prediction. Am. J. Chin. Med. 2021, 49, 359–389. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, S.; He, Y.; Li, X.; Yang, W.; Zhou, Y. Biolabel-Led Research Pattern Positions the Effects and Mechanisms of Sophorae Tonkinensis Radix et Rhizome on Lung Diseases: A Novel Strategy for Computer-Aided Herbal Medicine Research Based on Omics and Bioinformatics. Comput. Biol. Med. 2021, 136, 104769. [Google Scholar] [CrossRef]
  32. Song, L.; Jiang, N.; Tan, G.; Chen, Q.; Feng, S.; Zhang, Z. First Report of Round Leaf Spot on Sophora Tonkinensis Caused by Didymella Glomerata in China. Plant Dis. 2020, 105, 498. [Google Scholar] [CrossRef]
  33. Chen, D.-J.; Yuan, S.; Zhang, P.; An, Q.; Zou, J.-B.; Yuan, C.-M.; Zhao, L.-H.; Hu, Z.-X.; Hao, X.-J. Two New Isoflavones from the Roots of Sophora Tonkinensis. J. Asian Nat. Prod. Res. 2022, 25, 163–170. [Google Scholar] [CrossRef]
  34. Wei, K.H.; Xu, J.P.; Li, L.X.; Cai, J.Y.; Miao, J.H.; Li, M.H. In Vitro Induction and Generation of Tetraploid Plants of Sophora Tonkinensis Gapnep. Pharmacogn. Mag. 2018, 14, 149–154. [Google Scholar] [CrossRef] [PubMed]
  35. Sheng-Li, Y.; Hao-Ke, H.E.; Meng-Yin, T.; Xin, L.I.; Su-Juan, W. Phenylpropanoid glycosides from Sophorae Tonkinensis Radix et Rhizoma. Zhongguo Zhong Yao Za Zhi 2020, 45, 5525–5529. [Google Scholar] [CrossRef]
  36. Yang, J.-H.; Kang, S.-Q.; Su, X.-H.; Lin, N.; Kong, X.-Y. Visual analysis of research progress of Sophorae Tonkinensis Radix et Rhizoma based on CiteSpace. Chin. Tradit. Herbal. Drugs 2021, 46, 1763–1768. [Google Scholar] [CrossRef]
  37. Qin, W.-L.; Lin, Z.-Z.; Wu, Q.-H.; Lan, Z.-Z.; Jiang, N. Study of the Cultivation of Wid Sophora tonkinsis Gapnep. Lishizhen Med. Mater. Med. Res. 2006, 17, 1668–1669. [Google Scholar]
  38. Chen, J.; Du, X.; Xu, X.; Zhang, S.; Yao, L.; He, X.; Wang, Y. Comparative Proteomic Analysis Provides New Insights into the Molecular Basis of Thermal-Induced Parthenogenesis in Silkworm (Bombyx Mori). Insects 2023, 14, 134. [Google Scholar] [CrossRef] [PubMed]
  39. Roberts, E.H. Predicting the Storage Life of Seeds. Seed Sci. Technol. 1973, 1, 499–514. [Google Scholar]
  40. Ellis, R.H.; Hong, T.D.; Roberts, E.H. An Intermediate Category of Seed Storage Behaviour? I. COFFEE. J. Exp. Bot. 1990, 41, 1167–1174. [Google Scholar] [CrossRef]
  41. Hong, T.D.; Linington, S.; Ellis, R.H.; Engels, J. Seed Storage Behaviour: A Compendium; International Plant Genetic Resources Institute (IPGRI): Rome, Italy, 1996; ISBN 978-92-9043-280-7. [Google Scholar]
  42. Liu, D.; Liu, J.-S. Research progress in dehydration tolerance of seed. J. Agric. Sci. Yanbian Univ. 2014, 36, 362–367. [Google Scholar] [CrossRef]
  43. Zong, M.; Cai, Y.-P.; Jin, X. Effective Factor s of Br owning and Relationship between Br owning and Desiccation Toler ance during the Development of Chestnut Seed. J. Anhui Agric. Sci. 2005, 8, 1405–1407. [Google Scholar] [CrossRef]
  44. Tao, Y.-L.; Zhu, C. Relationship among Desiccation-tolerance Proteinsand Soluble Sugars before and after Maturation of Castanea mollissima Seeds. Sci. Silv. Sin. 2004, 2, 45–50. [Google Scholar]
  45. Yang, Q.-H.; Ye, W.-H.; Song, S.-Q.; Yin, W.-S. Seed desiccation -tolerance and its relationship to seed types and developmental stages. ActaBot Boreali Occident. Sin. 2002, 06, 238–245. [Google Scholar]
  46. Xin, P.-P.; Hu, X.-R.; Zhang, Y.-L.; Hu, C.-L.; Guo, X.-R. Study on energy saving preservation of scallion and rape seed drying and ultra drying. Seed World 2001, 3, 21–22. [Google Scholar]
  47. Varghese, B.; Sershen; Berjak, P.; Varghese, D.; Pammenter, N.W. Differential Drying Rates of Recalcitrant Trichilia Dregeana Embryonic Axes: A Study of Survival and Oxidative Stress Metabolism. Physiol. Plant. 2011, 142, 326–338. [Google Scholar] [CrossRef]
  48. Gomes, M.P.; Garcia, Q.S. Reactive Oxygen Species and Seed Germination. Biologia 2013, 68, 351–357. [Google Scholar] [CrossRef]
  49. Araldi, C.G.; Coelho, C.M.M.; Gaziola, S.A.; Azevedo, R.A. Storage Elicits a Fast Antioxidant Enzyme Activity in Araucaria Angustifolia Embryos. Acta Physiol. Plant 2016, 38, 201. [Google Scholar] [CrossRef]
  50. Zhuang, Z.; Shi, F.-H.; Ding, Y.-F.; Shen, Y.-B. Study on dehydration tolerance of Citron Fruit seed. Sci. Agric. JiangSu 2015, 43, 229–231. [Google Scholar] [CrossRef]
  51. Pizarro, M.; Contreras, R.A.; Köhler, H.; Zúñiga, G.E. Desiccation Tolerance in the Antarctic Moss Sanionia Uncinata. Biol. Res. 2019, 52, 46. [Google Scholar] [CrossRef] [PubMed]
  52. Manghwar, H.; Li, J. Endoplasmic Reticulum Stress and Unfolded Protein Response Signaling in Plants. Int. J. Mol. Sci. 2022, 23, 828. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, J.-X.; Howell, S.H. Endoplasmic Reticulum Protein Quality Control and Its Relationship to Environmental Stress Responses in Plants. Plant Cell 2010, 22, 2930–2942. [Google Scholar] [CrossRef]
  54. Al-Whaibi, M.H. Plant Heat-Shock Proteins: A Mini Review. J. King Saud. Univ. Sci. 2011, 23, 139–150. [Google Scholar] [CrossRef]
  55. Cheng, X.; Wang, W.-J.; Cao, S.; Li, R.-X.; Zhang, B.-B.; Sun, A.-Q.; Zhang, C.-Q. Mechanism of DA-6 Treatment Regulating Wheat Seed Vigor After Anthesis. Sci. Agric. Sin. 2021, 54, 1821–1834. [Google Scholar]
  56. Demska, K.; Filip, E.; Skuza, L. Expression of Genes Encoding Protein Disulfide Isomerase (PDI) in Cultivars and Lines of Common Wheat with Different Baking Quality of Flour. BMC Plant Biol. 2018, 18, 294. [Google Scholar] [CrossRef] [PubMed]
  57. Zhang, C.-Q.; Feng, L.-H.; Gu, M.-H.; Liu, Q.-Q. Progress on inheritance and gene cloning for rice grain quality in Jiangsu province. Hereditas 2021, 43, 425–441. [Google Scholar] [CrossRef]
  58. Urade, R. Oxidative Protein Folding in the Plant Endoplasmic Reticulum. Biosci. Biotechnol. Biochem. 2019, 83, 781–793. [Google Scholar] [CrossRef]
  59. Schieber, M.; Chandel, N.S. ROS Function in Redox Signaling and Oxidative Stress. Cur. Biol. 2014, 24, R453–R462. [Google Scholar] [CrossRef] [PubMed]
  60. Arnér, E.S.; Holmgren, A. Physiological Functions of Thioredoxin and Thioredoxin Reductase. Eur. J. Biochem. 2000, 267, 6102–6109. [Google Scholar] [CrossRef]
  61. Obroucheva, N.; Sinkevich, I.; Lityagina, S. Physiological Aspects of Seed Recalcitrance: A Case Study on the Tree Aesculus hippocastanum. Tree Physiol. 2016, 36, 1127–1150. [Google Scholar] [CrossRef]
  62. Amara, I.; Zaidi, I.; Masmoudi, K.; Ludevid, M.D.; Pagès, M.; Goday, A.; Brini, F. Insights into Late Embryogenesis Abundant (LEA) Proteins in Plants: From Structure to the Functions. Am. J. Plant Sci. 2014, 5, 3440–3455. [Google Scholar] [CrossRef]
  63. Liu, H.; Xing, M.; Yang, W.; Mu, X.; Wang, X.; Lu, F.; Wang, Y.; Zhang, L. Genome-Wide Identification of and Functional Insights into the Late Embryogenesis Abundant (LEA) Gene Family in Bread Wheat (Triticum aestivum). Sci. Rep. 2019, 9, 13375. [Google Scholar] [CrossRef]
  64. Huang, R.; Xiao, D.; Wang, X.; Zhan, J.; Wang, A.; He, L. Genome-Wide Identification, Evolutionary and Expression Analyses of LEA Gene Family in Peanut (Arachis hypogaea L.). BMC Plant Biol. 2022, 22, 155. [Google Scholar] [CrossRef] [PubMed]
  65. Geng, W.; Wang, Y.; Zhang, J.; Liu, Z.; Chen, X.; Qin, L.; Yang, L.; Tang, H. Genome-Wide Identification and Expression Analyses of Late Embryogenesis Abundant (LEA) Gene Family in Tobacco (Nicotiana tabacum L.) Reveal Their Function in Abiotic Stress Responses. Gene 2022, 836, 146665. [Google Scholar] [CrossRef]
  66. Li, Z. Identification and Genetic Effect of LEA Genes During Seed Development in Linseed Flax. Master’Thesis, Chinese Academy of Agricultural Sciences, Beijing, China, 2021. Available online: https://kns.cnki.net/kcms2/article/abstract?v=3uoqIhG8C475KOm_zrgu4lQARvep2SAkOTSE1G1uB0_um8HHdEYmZtFg9vXhdECNf9kdrWHWDI9g8wdpBt7Zg4FLKWL-UD4b&uniplatform=NZKPT (accessed on 15 February 2023).
  67. Cheng, J.-Y.; Zhang, D. Correlation Analysis of Gene Expression of LEA Protein and Moisture Content During Embryo Maturation in Nelumbo nucifera. J. Northwest For. Univ. 2020, 35, 75–81+120. [Google Scholar]
  68. Zhang, H.M. Proteomic Analysis of Drought Stress Responses and LEA proteins. Ph. D. Thesis, Northwest A & F University, Beijing, China, 2014. Available online: https://kns.cnki.net/kcms2/article/abstract?v=3uoqIhG8C447WN1SO36whLpCgh0R0Z-i4Lc0kcI_HPe7ZYqSOTP4QvUsX-vsGudGdptCq0hsEzYmYphSeKmAOmJonwPFphA1&uniplatform=NZKPT (accessed on 15 February 2023).
  69. Jin, X.; Cao, D.; Wang, Z.; Ma, L.; Tian, K.; Liu, Y.; Gong, Z.; Zhu, X.; Jiang, C.; Li, Y. Genome-Wide Identification and Expression Analyses of the LEA Protein Gene Family in Tea Plant Reveal Their Involvement in Seed Development and Abiotic Stress Responses. Sci. Rep. 2019, 9, 14123. [Google Scholar] [CrossRef]
  70. Delahaie, J.; Hundertmark, M.; Bove, J.; Leprince, O.; Rogniaux, H.; Buitink, J. LEA Polypeptide Profiling of Recalcitrant and Orthodox Legume Seeds Reveals ABI3-Regulated LEA Protein Abundance Linked to Desiccation Tolerance. J. Exp. Bot. 2013, 64, 4559–4573. [Google Scholar] [CrossRef]
  71. Yang, Q.-H.; Song, S.-Q.; Ye, W.-H.; Yin, S.-H. Relationship between dehydration tolerance and sugar in seeds. Bull. Bot. Res. 2003, 40, 204–210. [Google Scholar]
  72. Thormann, I.; Dulloo, M.; Engels, J. Techniques for Ex Situ Plant Conservation. In Plant Conservation Genetics; Haworth Press: Philadelphia, PA, USA, 2006; pp. 7–36. ISBN 978-1-56022-997-1. [Google Scholar]
  73. Walters, C.; Berjak, P.; Pammenter, N.; Kennedy, K.; Raven, P. Preservation of Recalcitrant Seeds. Science 1984, 339, 915–916. [Google Scholar] [CrossRef] [PubMed]
  74. Sakai, A.; Yoshida, S. Survival of Plant Tissue at Super-Low Temperature VI. Effects of Cooling and Rewarming Rates on Survival. Plant Physiol. 1967, 42, 1695–1701. [Google Scholar] [CrossRef]
  75. Zhao, D.-Q.; Song, J.-P.; Jia, H.-X.; Zhang, X.-H.; Yang, W.-L.; Wu, Y.-H.; Song, Z.; Wang, H.-P. Study on Cryopreservation Method of Allium Crop Germplasm. Plant Genet. Resour. 2022, 23, 1258–1268. [Google Scholar] [CrossRef]
  76. Hirano, T.; Godo, T.; Miyoshi, K.; Ishikawa, K.; Ishikawa, M.; Mii, M. Cryopreservation and Low-Temperature Storage of Seeds of Phaius tankervilleae. Plant Biotechnol. Rep. 2009, 3, 103–109. [Google Scholar] [CrossRef]
  77. de Oliveira, R.S.; Souza, F.V.D.; dos Santos, I.L.; de Souza, S.O.; Aona, L.Y.S.; de Souza, E.H. Cryopreservation and Low-Temperature Storage of Seeds of Tillandsia Species (Bromeliaceae) with Ornamental Potential. 3 Biotech. 2021, 11, 186. [Google Scholar] [CrossRef]
  78. Vishwakarma, P.K.; Vincent, L.; Vasugi, C.; Rajasekharan, P.E. Effect of Cryopreservation on Pollen Viability, Fertility and Morphology of Different Psidium Species. Cryobiology 2021, 98, 112–118. [Google Scholar] [CrossRef] [PubMed]
  79. Rasl, T.; Schalk, M.; Temsch, E.; Kodym, A. Direct Cryopreservation of Winter-Acclimated Buds of Dracocephalum austriacum (Lamiaceae) from Field Material. Plant Cell Tissue Organ Culture 2020, 142, 167–176. [Google Scholar] [CrossRef] [PubMed]
  80. Ninagawa, T.; Eguchi, A.; Kawamura, Y.; Konishi, T.; Narumi, A. A Study on Ice Crystal Formation Behavior at Intracellular Freezing of Plant Cells Using a High-Speed Camera. Cryobiology 2016, 73, 20–29. [Google Scholar] [CrossRef] [PubMed]
  81. Tian, X.-M.; Li, H.-L.; He, Y.; Hong, Q.-M.; Hu, W.-B.; Li, Q. Research Progress on Recalcitrant Seeds Storage of Tropical Crops. Chin. J. Trop. Agric. 2014, 34, 52–58. [Google Scholar]
  82. Kholina, A.B.; Voronkova, N.M. Seed Cryopreservation of Some Medicinal Legumes. J. Bot. 2012, 2012, 186891. [Google Scholar] [CrossRef]
  83. Hardstaff, L.K.; Sommerville, K.D.; Funnekotter, B.; Bunn, E.; Offord, C.A.; Mancera, R.L. Myrtaceae in Australia: Use of Cryobiotechnologies for the Conservation of a Significant Plant Family under Threat. Plants 2022, 11, 1017. [Google Scholar] [CrossRef]
  84. Dussert, S.; Chabrillange, N.; Rocquelin, G.; Engelmann, F.; Lopez, M.; Hamon, S. Tolerance of Coffee (Coffea Spp.) Seeds to Ultra-Low Temperature Exposure in Relation to Calorimetric Properties of Tissue Water, Lipid Composition, and Cooling Procedure. Physiol. Plant 2001, 112, 495–504. [Google Scholar] [CrossRef]
  85. Zeng, L.; He, M.-J.; Gu, Y.-K.; Wu, Y.; Zheng, X.-L. Study on Cryopreservation of Five Kinds of Medicinal Legume Seeds. Biot. Resour. 2017, 39, 42–47. [Google Scholar] [CrossRef]
  86. Xia, F.; Wang, X.; Li, M.; Mao, P. Mitochondrial Structural and Antioxidant System Responses to Aging in Oat (Avena sativa L.) Seeds with Different Moisture Contents. Plant Physiology and Biochemistry 2015, 94, 122–129. [Google Scholar] [CrossRef]
  87. Cheng, C.; Chen, L.; Chou, C.; Liang, J. Investigations of Riboflavin Photolysis via Coloured Light in the Nitro Blue Tetrazolium Assay for Superoxide Dismutase Activity. J. Photochem. 2015, 148, 262–267. [Google Scholar] [CrossRef] [PubMed]
  88. Maehly, A.C.; Chance, B. The Assay of Catalases and Peroxidases. Methods Biochem. Anal. 1954, 1, 357–424. [Google Scholar] [CrossRef] [PubMed]
  89. Hossain, M.A.; Asada, K. Inactivation of Ascorbate Peroxidase in Spinach Chloroplasts on Dark Addition of Hydrogen Peroxide: Its Protection by Ascorbate. Plant Cell Physiol. 1984, 25, 1285. [Google Scholar] [CrossRef]
Figure 1. Dynamic map of water content and germination percentage of S. tonkinensis seeds. (A) Changes of seed water content with dehydration time. (B) Changes of seed germination percentage with water content before and after cryopreservation treatment. Note: Different letters on top of each bar indicate statistically significant differences (p < 0.05).
Figure 1. Dynamic map of water content and germination percentage of S. tonkinensis seeds. (A) Changes of seed water content with dehydration time. (B) Changes of seed germination percentage with water content before and after cryopreservation treatment. Note: Different letters on top of each bar indicate statistically significant differences (p < 0.05).
Plants 12 01842 g001
Figure 2. Subcellular structural changes of plumular axis of S. tonkinensis seeds after dehydration and dehydration followed by cryopreservation. (A) Control. (B) Dehydration for 2 h. (C) Cryopreservation for 24 h in liquid nitrogen after 2 h-dehydration. (D) Dehydration for 24 h. (E) Cryopreservation for 24 h in liquid nitrogen after 24 h-dehydration. NU, nucleus; CW, cell wall; LB, lipid bodies; M, mitochondria; PB, protein bodies; S, starch grain; CY, cytoplasm; NE, nuclear envelope, NAB, nucleolus associated bodies.
Figure 2. Subcellular structural changes of plumular axis of S. tonkinensis seeds after dehydration and dehydration followed by cryopreservation. (A) Control. (B) Dehydration for 2 h. (C) Cryopreservation for 24 h in liquid nitrogen after 2 h-dehydration. (D) Dehydration for 24 h. (E) Cryopreservation for 24 h in liquid nitrogen after 24 h-dehydration. NU, nucleus; CW, cell wall; LB, lipid bodies; M, mitochondria; PB, protein bodies; S, starch grain; CY, cytoplasm; NE, nuclear envelope, NAB, nucleolus associated bodies.
Plants 12 01842 g002
Figure 3. Determination of antioxidant enzyme activity after dehydration and dehydration followed cryopreservation in S. tonkinensis seed plumular axis. (A) Changes of SOD activity with dehydration time before and after cryopreservation. (B) Changes of CAT activity with dehydration time before and after cryopreservation. (C) Changes of APX activity with dehydration time before and after cryopreservation. Note: *, **, and *** mean that the difference between the two groups is less than 0.05 (p < 0.05), 0.01 (p < 0.01), 0.001 (p < 0.001), respectively; ns means that the difference between the two groups is more than 0.05.
Figure 3. Determination of antioxidant enzyme activity after dehydration and dehydration followed cryopreservation in S. tonkinensis seed plumular axis. (A) Changes of SOD activity with dehydration time before and after cryopreservation. (B) Changes of CAT activity with dehydration time before and after cryopreservation. (C) Changes of APX activity with dehydration time before and after cryopreservation. Note: *, **, and *** mean that the difference between the two groups is less than 0.05 (p < 0.05), 0.01 (p < 0.01), 0.001 (p < 0.001), respectively; ns means that the difference between the two groups is more than 0.05.
Plants 12 01842 g003
Figure 4. The profile of the differentially expressed proteins (DEPs) in comparison groups (D2/CK, CD2/CK, D24/CK, and CD24/CK). (A) The number of DEPs identified in comparison groups. (B) Venn diagram analysis of DEPs among different comparison groups.
Figure 4. The profile of the differentially expressed proteins (DEPs) in comparison groups (D2/CK, CD2/CK, D24/CK, and CD24/CK). (A) The number of DEPs identified in comparison groups. (B) Venn diagram analysis of DEPs among different comparison groups.
Plants 12 01842 g004
Figure 5. GO and KEGG enrichment results of DEPs. (AD) GO enrichment analysis of DEPs in D2 vs. CK, CD2 vs. CK, D24 vs. CK, and CD24 vs. CK. (EH) KEGG enrichment analysis of DEPs in D2 vs. CK, CD2 vs. CK, D24 vs. CK, and CD24 vs. CK.
Figure 5. GO and KEGG enrichment results of DEPs. (AD) GO enrichment analysis of DEPs in D2 vs. CK, CD2 vs. CK, D24 vs. CK, and CD24 vs. CK. (EH) KEGG enrichment analysis of DEPs in D2 vs. CK, CD2 vs. CK, D24 vs. CK, and CD24 vs. CK.
Plants 12 01842 g005
Figure 6. Interaction network of DEPs.
Figure 6. Interaction network of DEPs.
Plants 12 01842 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Luo, Y.; Zhang, Y.; Jiang, Y.; Dai, Z.; Li, Q.; Mou, J.; Xu, L.; Deng, S.; Li, J.; Wang, R.; et al. iTRAQ-Based Proteomic and Physiological Analyses Reveal the Mechanisms of Dehydration and Cryopreservation Tolerance of Sophora tonkinensis Gagnep. Seeds. Plants 2023, 12, 1842. https://doi.org/10.3390/plants12091842

AMA Style

Luo Y, Zhang Y, Jiang Y, Dai Z, Li Q, Mou J, Xu L, Deng S, Li J, Wang R, et al. iTRAQ-Based Proteomic and Physiological Analyses Reveal the Mechanisms of Dehydration and Cryopreservation Tolerance of Sophora tonkinensis Gagnep. Seeds. Plants. 2023; 12(9):1842. https://doi.org/10.3390/plants12091842

Chicago/Turabian Style

Luo, Yongjian, Yixin Zhang, Yu Jiang, Zhangyan Dai, Qing Li, Jiaolin Mou, Li Xu, Shiming Deng, Jitao Li, Ru Wang, and et al. 2023. "iTRAQ-Based Proteomic and Physiological Analyses Reveal the Mechanisms of Dehydration and Cryopreservation Tolerance of Sophora tonkinensis Gagnep. Seeds" Plants 12, no. 9: 1842. https://doi.org/10.3390/plants12091842

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop