Next Article in Journal
Hypertensive Nephropathy: Unveiling the Possible Involvement of Hemichannels and Pannexons
Next Article in Special Issue
Evolution and Phylogeny of Soybean Mosaic Virus Based on 143 Complete Genomes
Previous Article in Journal
The Current Developments in Medicinal Plant Genomics Enabled the Diversification of Secondary Metabolites’ Biosynthesis
Previous Article in Special Issue
Conservation and Divergence of the Trihelix Genes in Brassica and Expression Profiles of BnaTH Genes in Brassica napus under Abiotic Stresses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Overexpression of γ-Tocopherol Methyltransferase on α-Tocopherol and Fatty Acid Accumulation and Tolerance to Salt Stress during Seed Germination in Brassica napus L.

State Key Laboratory of Crop Stress Biology for Arid Areas, National Yangling Agricultural Biotechnology & Breeding Center, Shaanxi Key Laboratory of Crop Heterosis and College of Agronomy, Northwest A&F University, Yangling 712100, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(24), 15933; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232415933
Submission received: 28 September 2022 / Revised: 9 December 2022 / Accepted: 12 December 2022 / Published: 14 December 2022
(This article belongs to the Special Issue Molecular Genetics and Plant Breeding 2.0)

Abstract

:
Rapeseed (Brassica napus L.) is an important oil crop and a major source of tocopherols, also known as vitamin E, in human nutrition. Enhancing the quality and composition of fatty acids (FAs) and tocopherols in seeds has long been a target for rapeseed breeding. The gene γ-Tocopherol methyltransferase (γ-TMT) encodes an enzyme catalysing the conversion of γ-tocopherol to α-tocopherol, which has the highest biological activity. However, the genetic basis of γ-TMT in B. napus seeds remains unclear. In the present study, BnaC02.TMT.a, one paralogue of Brassica napus γ-TMT, was isolated from the B. napus cultivar “Zhongshuang11” by nested PCR, and two homozygous transgenic overexpression lines were further characterised. Our results demonstrated that the overexpression of BnaC02.TMT.a mediated an increase in the α- and total tocopherol content in transgenic B. napus seeds. Interestingly, the FA composition was also altered in the transgenic plants; a reduction in the levels of oleic acid and an increase in the levels of linoleic acid and linolenic acid were observed. Consistently, BnaC02.TMT.a promoted the expression of BnFAD2 and BnFAD3, which are involved in the biosynthesis of polyunsaturated fatty acids during seed development. In addition, BnaC02.TMT.a enhanced the tolerance to salt stress by scavenging reactive oxygen species (ROS) during seed germination in B. napus. Our results suggest that BnaC02.TMT.a could affect the tocopherol content and FA composition and play a positive role in regulating the rapeseed response to salt stress by modulating the ROS scavenging system. This study broadens our understanding of the function of the Bnγ-TMT gene and provides a novel strategy for genetic engineering in rapeseed breeding.

1. Introduction

Rapeseed (Brassica napus L., AACC, 2n = 38) is one of the most important sources of edible oils worldwide. It is an allopolyploid species originating from the natural hybridisation between two ancestorial species, namely, B. rapa (AA, 2n = 20) and B. oleracea (CC, 2n = 18) [1]. The nutritional characteristics of rapeseed oil are mainly determined by its fatty acid (FA) composition and vitamin E content. The major FAs present in B. napus seeds (canola type) are oleic acid (C18:1), linoleic acid (C18:2), and linolenic acid (C18:3), with concentrations of 61%, 21%, and 11%, respectively [2]. Among these, C18:2 and C18:3 are polyunsaturated FAs (PUFAs) that cannot be biosynthesised in the human body. These PUFAs play an important role in the protection of eyesight and the prevention of various diseases, such as cancer and cardiovascular and inflammatory diseases [3,4,5]. Vitamin E, which is present in cold-pressed oil, is an essential dietary nutritional element for all mammals, and its deficiency leads to neurological disorders, ataxia, and even death [6,7]. Increasing the quantities of FAs and vitamin E has long garnered interest in rapeseed breeding [8]. Therefore, the characterisation of key genes regulating the biosynthesis of FAs and/or vitamin E in B. napus will be helpful in accurately designing new germplasms or varieties of crops with high PUFA and vitamin E contents using molecular biology techniques, thereby improving human health.
Vitamin E is an essential lipid-soluble antioxidant comprising tocopherols and tocotrienols. Tocopherols are mostly present in the seeds of common oil crops and the green-coloured components of higher-order plants, whereas tocotrienols are mainly found in cereal kernels and certain tropical fruits [9]. According to the number and position of the methyl substituents on the aromatic ring, tocopherols are classified into four forms: α, β, γ, and δ [10]. Among them, α-tocopherol is beneficial to human health, as it has the highest efficacy [11].
In plants, tocopherol biosynthesis occurs in plastids through a combination of two main pathways (Figure 1). The shikimate pathway leads to the synthesis of an aromatic ring from homogentisic acid (HGA) and p-hydroxyphenylpyruvic acid (HPP) by the enzyme HPP dioxygenase (PDS1/HPPD) in the cytoplasm [12,13]. The methylerythritol phosphate (MEP) pathway leads to the generation of a polyprenyl side chain from phytyl diphosphate (PDP) in plastids [14,15]. An additional pathway for PDP production was discovered from the phytol recycling pathway [16,17]. In tocopherol biosynthesis, the two substrates, HGA and PDP, are fused together to produce the first intermediate, 2-methyl-6-phytyl-1,4-benzoquinol (MBPQ). This reaction is mediated by the enzyme HGA phytyl transferase (VTE2/HPT). MPBQ is a substrate of either tocopherol cyclase (VTE1/TC) or MPBQ methyltransferase (VTE3/MPBQ-MT). MPBQ-MT catalyses the formation of 2,3-dimethyl-6-phytyl-1,4-benzoquinol (DMPBQ) from MPBQ. TC converts MPBQ and DMPBQ to δ- and γ-tocopherols, respectively. Finally, the last key enzyme, γ-tocopherol methyltransferase (VTE4/γ-TMT), catalyses the conversion of γ- to α-tocopherol and δ- to β-tocopherol, thereby determining the type of tocopherol [18,19]. The role of the γ-TMT gene has been functionally characterised in Arabidopsis thaliana. The loss of A. thaliana Atγ-TMT function results in reduced levels of α-tocopherol; however, an increased γ-tocopherol content in leaves is observed, and its overexpression leads to the approximate complete conversion of γ- to α-tocopherol and an increase in vitamin E activity in A. thaliana seeds [20,21]. Many useful crops with a high α-tocopherol content can be developed by introducing this gene. Transgenic soybean (Glycine max L.) [22], lettuce (Latuca sativa L.) [23,24], rice (Oryza sativa L.) [25], and perilla (Perilla frutescens L.) [26] with high levels of α-tocopherol have been developed by overexpressing the γ-TMT gene from the model plant A. thaliana (Atγ-TMT). Recently, transgenic soybean [27] and rapeseed (B. juncea L. AABB, 2n = 36) [28] with increased α-tocopherol content were developed by expressing the perilla γ-TMT gene. Generally, tocopherols in most common oilseeds contain minor amounts of α-tocopherol and relatively higher levels of γ-tocopherol [20,29]. Therefore, improving the γ- to α-tocopherol conversion by elevating the expression level of the γ-TMT gene in B. napus seeds would be a promising strategy.
The major tocopherol functions are related to antioxidant properties because of their ability to interact with polyunsaturated acyl groups and protect membrane lipids (especially PUFAs) from oxidative damage by scavenging lipid peroxy radicals and quenching reactive oxygen species (ROS) in photosystem II. Furthermore, they also prevent damage due to the fact of oxidation during membrane lipid peroxidation [10,30]. Tocopherols donate a hydrogen atom to lipid-free radicals, thereby neutralising the radical and terminating the autocatalytic chain reaction of lipid peroxidation and protecting cell membranes [31,32]. In plants, both tocopherol and FA biosynthesis occur in the plastid, and the tocopherol functions as an antioxidant for stabilising PUFAs against lipid oxidation in vegetable oils [32]. The balance between vitamin E and PUFA content mainly determines the susceptibility to lipid peroxidation and storage stability of rapeseed oil. However, there is a paucity of studies examining the correlation between tocopherol and FA components in rapeseed oils. A positive correlation has been demonstrated between linoleic acid (18:2) and α-tocopherol, linolenic acid (18:3) and γ-tocopherol from the vegetable oils of 14 different plant species [33]. In addition, upon studying a collection of maize hybrids, a relationship between FAs, particularly PUFAs, and tocochromanol content was observed [34]. Furthermore, the content of tocopherol and FA components in seeds from three Brassica oil crop species was analysed, and a significantly positive correlation was observed between α-tocopherol and the sum of C18:1 and C18:2, whereas no association was observed between the γ-tocopherol and FA components [35].
In addition to their antioxidant function, recent studies have suggested additional roles for tocopherols in diverse biological processes, including germination [36], cell signalling [37], and biotic [38] and abiotic stresses, such as those related to salt [39,40,41,42,43,44], low temperature [45,46], drought [41,47], and heavy metals [40,43,48,49]. The RNAi-mediated silencing of γ-TMT leads to a deficiency in α-tocopherol and elevated susceptibility to salt stress, whereas the overexpression of γ-TMT increases the salt and heavy metal tolerance in transgenic tobacco [39,43]. In addition, the increased sensitivity to salt stress in vitamin E-deficient mutants of A. thaliana has been observed. Tocopherols play a key role in salt stress tolerance not only by reducing the extent of oxidative stress but also by improving ion homeostasis and the hormonal balance of leaves [42]. Furthermore, the overexpression of the γ-TMT gene in transgenic B. juncea plants enhances the tolerance to abiotic stresses, such as salt, heavy metals, and osmoticums [40].
In the present study, BnaC02.TMT.a, an Atγ-TMT orthologue in B. napus, was isolated using nested PCR and functionally characterised. To explore the biological functions of BnaC02.TMT.a in rapeseed, we measured the contents of tocopherols and FAs and conducted a germination assay on sodium chloride (NaCl) stress medium in transgenic overexpression plants. We showed that the overexpression of BnaC02.TMT.a increased the α- and total tocopherol content, altered the proportion of FA composition, and promoted the germination rate after NaCl treatment. Under conditions involving salt-related stress, the accumulation of ROS is an important factor affecting the inhibition of seed germination, and hydrogen peroxide (H2O2) and superoxide anion (O2-) are the main components of ROS [50,51]. To obtain a deeper insight into the mechanism underlying increased salt tolerance in transgenic plants, we analysed ROS accumulation, which was estimated using diaminobenzidine (DAB) and nitroblue tetrazolium (NBT) staining and measured the H2O2 and O2- content. Our results suggest that BnaC02.TMT.a prevents oxidative damage by ROS scavenging during germination in B. napus seeds under conditions involving salt stress. This study revealed that the overexpression of BnaC02.TMT.a could affect tocopherols, FAs, and the response to salt stress, which will be useful for engineering rapeseed with improved vitamin E content, FA composition, and salt tolerance.

2. Results

2.1. Sequence Analysis of Bnγ-TMT Paralogs

Upon performing the protein BLAST using A. thaliana Atγ-TMT in Brassica napus pan-genome information resource (BnPIR), four Bnγ-TMT paralogs were predicted in the genome of the B. napus cultivar “Zhongshuang11 (ZS11)”: BnaC02T0331100ZS, BnaC02T0197500ZS, BnaA02T0247300ZS, and BnaA02T0154300ZS, which were located on the chromosomes C02 and A02 and designated BnaC02.TMT.a, BnaC02.TMT.b, BnaA02.TMT.a, and BnaA02.TMT.b, respectively. The alignment of the amino acid sequences of the four Bnγ-TMT paralogs revealed high identities ranging from 96.5% to 98.8% (Figure S1). The identity of the protein sequence between the four Bnγ-TMT proteins and Atγ-TMT was at least 88.4%, whereas BnaC02.TMT.a showed a higher identity with Atγ-TMT, reaching 89.3% (Figure S1). Moreover, all Bnγ-TMT paralogs were predicted to possess four highly conserved S-adenosylmethionine (SAM)-binding domains, similar to Atγ-TMT (Figure 2). This implies that all four paralogs of Bnγ-TMT may be functionally conservative. Phylogenetic analysis was performed to investigate the evolutionary relationships between the Bnγ-TMT and 11 γ-TMT proteins in A. thaliana and six other oil crops. As shown in Figure 3, BnaC02.TMT.b was most closely related to Atγ-TMT, whereas BnaC02.TMT.a was closely related to BnaA02.TMT.a. Compared with γ-TMT in other species, all four paralogs were closely related to Atγ-TMT.

2.2. Analysis of the Expression Pattern of Bnγ-TMT

To verify the function of Bnγ-TMT, quantitative reverse transcription PCR (RT-qPCR) was used to explore the spatiotemporal expression pattern of all Bnγ-TMT paralogs in various tissues of the B. napus cultivar “ZS11”. As illustrated in Figure 4, Bnγ-TMT was widely expressed in various tissues, with higher expression levels in the flowers, leaves, and developing seeds (35 days after pollination (DAP)) than in the roots and stems. Notably, during seed development, the Bnγ-TMT expression remained relatively low at the embryogenesis stage from 10 to 15 DAP, decreased slightly from 20 to 26 DAP, and progressively increased at 29 DAP and 32 DAP, reaching a maximum at 35 DAP (Figure 4).

2.3. BnaC02.TMT.a Increased α-Tocopherol Content in B. napus Seeds

To elucidate the role of Bnγ-TMT, we introduced the overexpression construct 35S:BnaC02.TMT.a-6 haemagglutinin (6HA) into the B. napus cultivar “ZS11” plants (Figure 5A). Two representative independent T2 homozygous lines (OE#3 and OE#5) were confirmed by the PCR amplification of BnaC02.TMT.a using the specific primers 35S_P/BnTMT_R (Figure 5B). Furthermore, the expression level of the Bnγ-TMT gene was confirmed in these transgenic plants by RT-qPCR. The expression of Bnγ-TMT was significantly higher in these transgenic plants than in wild-type ZS11 (Figure 5C).
To investigate the function of BnaC02.TMT.a in mediating tocopherol accumulation, we measured the levels of total tocopherol and α-, β-, γ-, and δ-tocopherol in the mature seeds of wild-type ZS11 and the transgenic lines OE#3 and OE#5. The levels of α-tocopherol and total tocopherol were much higher in both of the transgenic lines than those in the wild-type seeds (Figure 5D,E), whereas no apparent differences were observed in the levels of γ- and δ-tocopherol between the wild-type and transgenic plants (Figure S2). In addition, no β-tocopherol was detected in either the control or transgenic plants.

2.4. BnaC02.TMT.a Promoted PUFA Biosynthesis in B. napus Seeds

To elucidate the biological functions of BnaC02.TMT.a in seed FAs, we measured the quantities of the major FA compositions and total FAs of mature seeds between the wild-type ZS11 plants and the BnaC02.TMT.a homozygous transgenic lines OE#3 and OE#5. In the mature seeds of the transgenic lines, a significant decrease was observed in the C18:1 (oleic acid) content along with a significant increase in the amount of C18:2 (linoleic acid) and C18:3 (linolenic acid) (Figure 6A), suggesting that α-tocopherol might affect the conversion of oleic acid to linoleic acid and linoleic acid to linolenic acid. However, no significant differences were observed with respect to the total FA content (Figure S3D) and several seed morphological traits, such as the seed coat colour (Figure S3A), seed size (Figure S3B), and dry weight (Figure S3C), between the mature seeds of the wild-type and transgenic plants.
Three critical stages of seed oil accumulation (24, 28, and 32 DAP) were carefully selected to compare the expression profiles of BnFAD2 and BnFAD3 between the transgenic line OE#3 and the wild-type control. As expected, the expression of BnFAD2 was significantly higher at the seed maturation stage from 28 and 32 DAP in the transgenic seeds than in the wild-type seeds (Figure 6B), and the expression of BnFAD3 was consistently upregulated in the developing seeds of the OE#3 transgenic plants at 24 and 32 DAP (Figure 6B).

2.5. Functional Characterisation of BnaC02.TMT.a in the Seed Germination under Conditions Involving Salt Stress in B. napus

The effects of BnaC02.TMT.a on seed germination under conditions involving salt stress in the wild-type and transgenic lines OE#3 and OE#5 were examined in sterile redistilled water with or without 180 mM NaCl. As illustrated in Figure 7 and Figure 8, OE#3, OE#5, and the control ZS11 showed similar seed germination rates in the medium without NaCl treatment. Notably, under stress conditions involving 180 mM NaCl, the differences in germination between the wild-type and transgenic seeds were apparent (Figure 7 and Figure 8). The seed germination was completely inhibited during the first 40 h after sowing (HAS) and from 40 to 192 HAS, which was significantly higher in the transgenic line than in the controls (Figure 8).
Under control conditions, the ROS levels were similar between the wild-type and transgenic lines OE#3 and OE#5. In contrast, upon treatment with 180 mM NaCl, a lower DAB and NBT staining intensity, especially in cotyledons, was observed with the transgenic plants, suggesting that the ROS levels were significantly suppressed in the transgenic lines OE#3 and OE#5 compared to those in the wild-type ZS11 (Figure 9A). Paralleling the observations described in Figure 9A, under control conditions, no significant differences were observed in the H2O2 and O2 contents between the wild-type and transgenic plants. Compared to the controls, the H2O2 and O2 contents were substantially decreased in both of the transgenic plants exposed to 180 mM NaCl (Figure 9B). Therefore, our results revealed that the overexpression of BnaC02.TMT.a alleviated the oxidative damage to the seeds under conditions involving salt stress by ROS scavenging, thereby promoting seed germination under such conditions.
Our results suggest that Bnγ-TMT plays a positive role in regulating the response of B. napus to salt stress, which might involve complex mechanisms, such as the ROS scavenging system.

3. Discussion

Rapeseed is one of the most important oil supply crops and contains a significant amount of tocopherols, which are beneficial for human health. Enhancing the quality and composition of FAs and tocopherols in seeds has long been a target for rapeseed breeding. In this study, we showed that BnaC02.TMT.a promoted the accumulation of α- and total tocopherol, altered the FA composition in seeds, and enhanced the tolerance to salt stress during seed germination in B. napus.
B. napus is an allotetraploid species formed by the natural hybridisation between B. rapa and B. oleracea [1]. A large number of chromosome duplications, rearrangements, and deletions occurred during the evolutionary processes, resulting in an average of 2–8 paralogs of each A. thaliana locus in the B. napus genome [52]. Consistently, four paralogs to Atγ-TMT were found in the B. napus cultivar “ZS11” genome (Figure 2). It was shown the overexpression of Atγ-TMT increased the α-tocopherol concentration and vitamin E activity with unaltered total tocopherol quantity in seeds, and the Atγ-TMT mutation exhibited an absent α-tocopherol but high levels of γ-tocopherol accumulation in the leaves of A. thaliana [20,21]. We found that the overexpression of BnaC02.TMT.a promoted α-tocopherol accumulation in B. napus seeds (Figure 5D). BnaC02.TMT.a shared a high amino acid identity and a close evolutionary relationship with Atγ-TMT (Figure 2 and Figure 3). In addition, four SAM binding domains involved in methyl transfer reactions [53] were highly conserved between BnaC02.TMT.a and Atγ-TMT (Figure 2). These results suggest that BnaC02.TMT.a may exhibit a conserved role in regulating the accumulation of α-tocopherol in A. thaliana. In rapeseed oil, the major tocopherol form is γ-tocopherol (65%), followed by α-tocopherol (35%). Furthermore, a small proportion of δ-tocopherol (<1%), and β-tocopherol is absent [54]. BnaA.VTE4.a1 (one γ-TMT paralogue from B. napus variety “Express”) is able to increase the α-tocopherol content because of the shift from γ- to α-tocopherol, and the total tocopherol level is not altered in transgenic A. thaliana seeds [55]. Interestingly, we found that the overexpression of BnaC02.TMT.a resulted in higher levels of α-tocopherol (Figure 5D); however, no obvious changes were observed in γ-tocopherol content in B. napus seeds (Figure S2A), which may explain the higher levels of total tocopherol (Figure 5E). The distinct roles of BnaA.VTE4.a1 and BnaC02.TMT.a in γ- and total tocopherol production might be attributed to their amino acid differences and functional divergence (Figure S4). Notably, the functional characterisation of BnaA.VTE4.a1 was performed in a heterologous transgenic system, which may not truly reflect its function in B. napus. A possible reason for the unaltered γ-tocopherol content is that γ-tocopherol could be replenished by other mechanisms in the B. napus transgenic lines.
The levels of α-tocopherol are positively correlated with linoleic acid and PUFAs in vegetable oils [33,34]. Upon examining a collection of B. napus lines with low erucic acid, a significant positive correlation was found between the α-/γ-tocopherol ratio and C18:2 and PUFAs [35]. Accompanied by an increased α-tocopherol level, the FA composition was altered; that is, decreased levels of oleic acid (C18:1) and increased levels of linoleic acid (C18:2) and linolenic acid (C18:3) were observed in 35S:BnaC02.TMT.a-6HA plants (Figure 6A). FAD2 [56,57] and FAD3 [58,59] are essential for the formation of C18:2 and C18:3, respectively, and higher transcript levels of BnFAD2 and BnFAD3 (Figure 6B) may be helpful for the accumulation of C18:2 and C18:3 in 35S:BnaC02.TMT.a-6HA seeds. As previously reported, the association between tocopherol and FAs in oils is mostly attributed to the biochemical function of tocopherol in protecting against lipid oxidation, rendering it crucial for PUFA stability [33,34]. Furthermore, biochemical and genetic evidence indicates that tocopherols have a more limited role in photoprotection than previously assumed [45,46]. In the absence of tocopherols, the conversion of 18:2 to 18:3 in endoplasmic reticulum (ER) lipids is reduced under low temperatures, suggesting some plausible scenarios for the direct or indirect interaction of tocopherols during ER PUFA metabolism. FAD2 and FAD3 are ER-localised FA desaturases. Together with the findings indicating that BnaC02.TMT.a enhanced the expression of FAD2 and FAD3 in 35S:BnaC02.TMT.a-6HA developing seeds (Figure 6B), we hypothesised that in addition to their antioxidant activity, the function of α-tocopherol in regulating FA formation may be attributed to the interaction between α-tocopherol and ER FA metabolism.
The role of tocopherols in abiotic stresses, such as salt [39,40,41,42,43,44], low temperature [45,46], drought [41,47], and heavy metals [40,43,48,49], has been observed. These results, together with our observations of 35S:BnaC02.TMT.a-6HA plants under conditions involving salt stress (Figure 7, Figure 8 and Figure 9), suggest that tocopherols play a crucial role in the alleviation of salt stress. Stresses lead to excessive ROS levels causing oxidative damage [60,61]. H2O2 and O2- are the most common forms of ROS generated during photosynthesis [50]. In this study, the overexpression of BnaC02.TMT.a promoted seed germination (Figure 7 and Figure 8) and reduced the production of H2O2 and O2- in B. napus under conditions involving salt stress (Figure 9), indicating that BnaC02.TMT.a alleviated the negative effects of NaCl toxicity by scavenging ROS. This was reminiscent of the expression of γ-TMT in chloroplasts, accelerating the conversion of γ-tocopherol to α-tocopherol, which contributes to a decrease in susceptibility to salt [43]. Tocopherol-deficient Arabidopsis mutants were particularly sensitive to salt stress, highlighting the role of α-tocopherol in maintaining cellular Na+/K+ homeostasis and hormonal balance [42,47]. The overexpression of Msγ-TMT in alfalfa (Medicago sativa L.) can directly or indirectly affect the biosynthesis of multiple phytohormones that may play a comprehensive role in drought stress tolerance [47]. FAs or FA derivatives function as signalling molecules or hormones, carbon and energy storage materials, and surface layers that protect plants from environmental stresses [62,63,64]. In addition, C18:3 serves as a precursor for the biosynthesis of jasmonic acid, which is involved in the response to various stresses [65,66]. Overall, considering the changes in FA composition, especially with an increase in the proportion of C18:3 (Figure 6A), other possible reasons for the better seed germination performance of the 35S:BnaC02.TMT.a-6HA plants in the present study might, in part, be attributed to signalling cascades connecting tocopherol levels with FAs, hormonal responses, ion homeostasis, and all aspects that warrant further investigation.
Plant growth and crop production are adversely affected by environmental stresses in nature. By uncovering the genes that underlie the tolerance adaptive trait, natural variation has the potential to be introgressed into elite cultivars [67]. Modern analytical approaches, such as genomic prediction, machine learning, and multi-trait gene editing, as well as genome–environment associations, were applied to speed up the identification and deployment of genotypic sources for climate change adaptation [67,68]. Indeed, genotypes from advanced interspecific congruity backcross exhibit promising responses to extreme conditions (i.e., heat and drought), which offers novel perspectives to breed traits of interest in the face of multiple climate changes [69]. Our study revealed that BnaC02.TMT.a acts as a positive regulator of tocopherol biosynthesis and salt stress response, providing a beneficial candidate for improving tocopherol activity and salt stress resistance in rapeseed breeding. Tocopherol-associated traits could be employed in traditional breeding for rapeseed improvement. Furthermore, advanced breeding strategies established in recent years have provided new avenues for harnessing tocopherols biosynthesis to improve the adaptation to stressful environments.

4. Materials and Methods

4.1. Plant Material and Growth Conditions

The B. napus cultivar “Zhongshuang11 (ZS11)” was used in this study. It was obtained through a material transfer agreement from the Oil Crops Research Institute of the Chinese Academy of Agricultural Sciences, Wuhan, China, and selfed for at least 10 generations. The rapeseed seeds were placed in a glass dish with three layers of filter paper soaked with sterile redistilled water and imbibed at 4 °C for one day in the dark. After imbibition, the seeds were transferred to 11 × 11 cm pots and grown in a greenhouse at 22 °C with a long day duration (LD, 16 h light/8 h dark) for six weeks. Subsequently, the plants were vernalised for four weeks at 4 °C under LD conditions in a cold chamber. After vernalisation, the plants were returned to the initial greenhouse conditions for ten weeks until harvest.

4.2. Protein Sequence and Phylogenetic Analysis

The protein sequences of γ-TMT were obtained from the National Center for Biotechnology Information (NCBI) database and B. napus pan-genome information resource (BnPIR) database (http://cbi.hzau.edu.cn/bnapus/index.php, accessed on 6 September 2022). The protein sequence alignment was performed using MUSCLE (http://www.ebi.ac.uk/Tools/msa/muscle/, accessed on 8 September 2022). The S-adenosyl methionine-binding domain was predicted using the Conserved Domain Search program (http://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/Structure/cdd/wrpsb.cgi, accessed on 8 September 2022). The phylogenetic analysis was performed using MEGA7 to confirm the evolutionary relationships between the γ-TMT protein sequences. A bootstrap analysis with 1000 replicates was performed to assess the statistical reliability of the tree topology.

4.3. Gene Cloning and Plasmid Construction

Owing to a high sequence similarity among the four paralogs of Bnγ-TMT, nested PCR was used for cloning BnaC02.TMT.a. Two pairs of PCR primers (specific and nested primers) were designed. The specific primers were designed based on the alignment of four Bnγ-TMT paralogs, and the nested primers were designed based on the full-length coding sequence (CDS) of BnaC02.TMT.a (BnaC02T0331100ZS) without a stop codon derived from the BnPIR database. To obtain full-length BnaC02.TMT.a cDNA, the total RNA was extracted from the developing seeds of the B. napus cultivar “ZS11” using the SteadyPure Plant RNA Extraction Kit (Accurate Biology, Changsha, China) and was subjected to reverse transcription to obtain first-strand cDNA according to the manufacturer’s instructions (TransGen, Beijing, China).
The full-length BnaC02.TMT.a CDS was amplified by nested PCR using specific and nested primers. The first pair of the PCR primers amplified a specific fragment using the primers BnaC02.TMT.a_Fc and BnaC02.TMT.a_Rc. The second pair of primers, called nested primers, was used to clone the full-length CDS of BnaC02.TMT.a. The sequence was ligated into the binary vector pGreen-35S-6HA under the control of the CaMV35S (35S) promoter, which had been digested by Xho1 and Cla1 [70]. The primers used in this study are listed in Supplementary Table S1.

4.4. Generation of B. napus Transgenic Plants

To obtain the transgenic B. napus plants, the hypocotyl segments of the B. napus cultivar “ZS11” were infected with the plasmid 35S:BnaC02.TMT.a-6HA by Agrobacterium tumefaciens strain GV3101 [71]. Transgenic T0 plants were selected on MS agar medium containing 10 μg/mL phosphinothricin, and a successful transformation was confirmed by DNA genotyping according to the PCR amplification performed using gene-specific primers. The transformants were selected by genotyping according to PCR amplification, and homozygous seeds from the T2 generation were used for subsequent experiments after cultivation under the same aforementioned conditions.

4.5. Morphological Observation of Mature Seeds

The T3 plants of the BnaC02.TMT.a homozygous transgenic lines (OE#3 and OE#5) and wild-type ZS11 controls were grown in a greenhouse under the aforementioned conditions. Siliques were harvested from the middle and basal parts of the main inflorescence. Subsequently, mature seeds were randomly selected from the different lines and observed using a stereomicroscope (Olympus SZ 61, Tokyo, Japan).

4.6. Measurement of the FA Content

The T3 plants of the BnaC02.TMT.a homozygous transgenic lines (OE#3 and OE#5) and wild-type ZS11 controls were grown in a greenhouse under the aforementioned conditions. For the FA measurement, seeds were harvested from the middle and basal parts of the main stem of eight individual plants from each line. The FAs were extracted and analysed as previously described [72,73], with minor modifications. First, the seeds were ground in a mortar, and approximately 8 mg of seed powder was heated at 80 °C in a methanol solution containing 2.5% (v/v) H2SO4 for 2 h, after which the total FAs were converted to FA methyl-esters. After cooling to room temperature, the FA methyl-esters were extracted with 1 mL hexane and 2 mL 0.9% (w/v) NaCl, and the organic phase was transferred to autoinjector vials. The gas chromatography (GC-2010 Plus, Shimadzu, Japan) analysis was performed using methyl heptadecanoate as an internal standard. The concentration of each FA species was normalised to that of the internal control.

4.7. HPLC Analysis of the Tocopherol Content

The tocopherol analysis was performed by Suzhou Michy Biology (http://www.michybio.com, accessed on 8 July 2021). The dried seeds from eight individual plants (transgenic lines and wild-type ZS11) were sent to the company. Briefly, 30–80 mg of mature seeds were ground, and the powder was mixed with 1 mL of ethyl alcohol extraction buffer containing 1% butylated hydroxytoluene (BHT) and 1 mL of 100 g/L potassium hydroxide solution. After incubation for 40 min (85 °C) and cooling to room temperature, 2 mL of petroleum ether was added. After mixing and centrifugation, the lower organic layer was recovered, dried under nitrogen, and resuspended in 0.4 mL methanol for the HPLC analysis.
The tocopherol content of the organic extract was determined using an Agilent 1100 HPLC system. A total of 10 μL of the sample was chromatographically analysed using a Compass C18 column (250 × 4.6 mm length, 5 μm particle size) and a solvent system consisting of methyl alcohol as the mobile phase with a flow rate of 1 mL/min. The column temperature was 20 °C, and the sample was scanned at 294 nm. The identification and quantification of tocopherols were performed by comparing the retention time and peak area to tocopherol standards (Supplementary Table S2).

4.8. ROS Staining

As previously described [74], the seeds collected at 96 h after sowing (HAS) were incubated in the dark at 37 °C. The seeds were saturated with 0.1% (w/v) diaminobenzidine (DAB, Sigma-Aldrich, St. Louis, MO, USA) in 10 mM 2-(N-morpholino) ethanesulphonic acid (MES, pH 6.5) for 8 h for the H2O2 staining and with 0.1% (w/v) nitroblue tetrazolium (NBT, Sigma-Aldrich, St. Louis, MO, USA) in 50 mM potassium phosphate buffer (pH 6.4) for 10 min for superoxide staining. Finally, the seeds were destained with 95% (v/v) ethanol.

4.9. ROS Quantification

The H2O2 content was measured using the titanium–peroxide complex method, as described previously, with minor modifications [74,75,76]. A total of 0.1 g (fresh weight) of the seeds collected at 96 HAS were ground in liquid nitrogen and extracted with 1 mL of cooled acetone. After the vortex mixing, the samples were centrifuged for 10 min at 8,000× g at 4 °C. The supernatant was then added to a solution comprising 0.1 mL of 5% (w/v) titanium sulphate and 0.2 mL of ammonia. The mixture was centrifuged for 10 min at 8,000× g. After removing the supernatant, the precipitate was dissolved in 1 mL of 2 M H2SO4. The absorbance was immediately measured at 415 nm. The H2O2 content was calculated using a standard curve generated using known concentrations of H2O2.
The O2 content was quantified as previously described, with minor modifications [74,76]. First, 0.1 g (fresh weight) of the seeds at 96 HAS was ground in liquid nitrogen and added to 1 mL of K-phosphate buffer (65 mM, pH 7.8). The homogenates were centrifuged at 12,000× g for 20 min (4 °C). A total of 1 mL 0.1 M hydroxylamine was added to the supernatant, and the mixture was incubated at 25 °C for 20 min. Then, 1 mL of 7 mM α-naphthalenamine and 1 mL of 58 mM 4-aminobenzenesulfonic acid were added, and the samples were centrifuged for 20 min at 12,000× g at 4°C. Trichloromethane (1 mL) was added to eliminate the pigments, and the mixture was centrifuged at 10,000× g for 5 min. The absorbance of the supernatant was measured at 530 nm. The O2- content was determined using a standard curve generated from known concentrations of NaNO2.

4.10. Germination Test

For the seed germination assay, 50 plump seeds from each treatment group were carefully picked and washed with sterile redistilled water thrice, evenly placed on a glass dish with three layers of filter paper soaked in the same volume of 180 mM NaCl solution or sterile redistilled water as the control, and imbibed at 4 °C for one day in the dark. After imbibition, the seeds were transferred to a growth chamber for germination. Every other day, the germinating seeds were transferred to a new glass dish with three layers of filter paper soaked in the same volume of 180 mM NaCl solution or sterile redistilled water until the seed germination experiment was completed. Three independent biological replicates were obtained for the control and salt treatments of the transgenic lines (OE#3 and OE#5) and ZS11. Seed germination was defined as the complete emergence of the radicle through the seed coat [77]. The number of germinated seeds was calculated every 8 h until the completion of the experiment.

4.11. Gene Expression Analysis

The roots, stems, flowers, and developing seeds for the spatial and temporal expression analysis were harvested from at least eight individual plants grown in different pots. The total RNA was isolated using the SteadyPure Plant RNA Extraction Kit (Accurate Biology, Changsha, China) and reverse-transcribed using EasyScript One-Step gDNA Removal and cDNA Synthesis SuperMix (TransGen, Beijing, China) according to the manufacturer’s instructions. Quantitative reverse transcription RT-PCR (RT-qPCR) was performed in 96-well blocks using SYBR Green Master Mix (Cofitt, Hong Kong, China) with three biological replicates and three technical replicates. Considering the high sequence similarity among the paralogs of Bnγ-TMT, the primers were designed for the conserved regions within groups of paralogs to detect joint gene expression. The relative expression levels were calculated using a modified double delta method [78]. The housekeeping gene encoding glyceraldehyde-3-phosphate dehydrogenase (BnGAPDH), a key enzyme in the process of glycolysis and gluconeogenesis, was considered as a reference, which is widely founds in various organisms and is regarded as an internal control [79]. The primers used for RT-qPCR are listed in Supplementary Table S1.

4.12. Statistical Analysis

A completely randomised design was used in this study. All data were classified using Win-Excel and reported as the mean ± standard deviation (SD). The data were analysed by one-way analysis of variance (ANOVA) using SPSS software (version 17.0, SPSS Inc., Chicago, IL, USA). Significant differences were determined by Student’s t-test (p ≤ 0.05).

5. Conclusions

In summary, we discovered new functions for Bnγ-TMT in the regulation of tocopherol and FA biosynthesis and salt tolerance in rapeseed. Since it is an allopolyploid species, four paralogs of Bnγ-TMT were found in the B. napus genome. The protein sequence and phylogenetic analyses suggested that all four Bnγ-TMT paralogs might exhibit functions similar to those shown by Atγ-TMT for determining tocopherol accumulation. Together with the expression data, these results further suggest that Bnγ-TMT might regulate tocopherol accumulation mainly during the maturation stage in B. napus seeds. Overexpression of BnaC02.TMT.a promotes seed α- and total tocopherol accumulation during seed maturation. Additionally, the fatty acid composition was altered. Our findings suggest that BnaC02.TMT.a promoted the conversion of oleic acid to linoleic acid and subsequently linoleic acid to linolenic acid by upregulating the expression of BnFAD2 and BnFAD3 in rapeseed. Furthermore, BnaC02.TMT.a plays a crucial role in mediating the response to salt stress during seed germination, which involves complex mechanisms, including ROS scavenging capability, signalling cascades connecting tocopherol with FAs, hormonal responses, and ion homeostasis. This study broadens our understanding of the function of the Bnγ-TMT gene and provides a novel strategy for genetic engineering to improve rapeseed breeding.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ijms232415933/s1.

Author Contributions

M.C. conceived and designed the experiments; D.L. and Y.G. conducted the experiments and analysed the data; T.L., M.L., Y.L. and W.Z. conducted parts of the experiments; Y.G. and D.L. wrote the draft of the manuscript; Z.L. and M.C. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by a grant from the Yang Ling Seed Industry Innovation Center (Grant no. K3031122024), the National Natural Science Foundation of China (Grant no. 31801393), and the National Key Research and Development Program of China (Grant no. 2022YFD1200400).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data included in this study are available upon reasonable request by contact with the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

A. thalianaArabidopsis thaliana
ANOVA analysis of variance
BHTbutylated hydroxy toluene
BnPIRBrassica napus pan-genome information resource
B. junceaBrassica juncea
B. napusBrassica napus
B. oleraceaBrassica oleracea
B. rapaBrassica rapa
CDS coding sequence
DAB diaminobenzidine
DAPdays after pollination
DMBPQ2,3-dimethyl-6-phytyl-1,4-benzoquinol
ER endoplasmic reticulum
FAsfatty acids
GAPDHglyceraldehyde-3-phosphate dehydrogenase
GGPPgeranylgeranyl diphosphate
GGRgeranylgeranyl reductase
HA haemagglutinin
HAS h after sowing
HGAhomogentisic acid
H2O2hydrogen peroxide
HPPp-hydroxyphenylpyruvic acid
HPPDHPP dioxygenase
HPTHGA phytyl transferase
LDlong day
MEPmethylerythritol phosphate
MES 2-(N-morpholino) ethanesulfonic acid
MBPQ2-methyl-6-phytyl-1,4-benzoquinol
MPBQ-MTMPBQ methyltransferase
NaCl sodium chloride
NBTnitroblue tetrazolium
NCBINational Center for Biotechnology Information
O2-superoxide anion
PDPphytyl diphosphate
Phytyl-Pphytyl phosphate
PKphytol kinase
PPKphytyl phosphate kinase
PUFAs polyunsaturated FAs
ROSreactive oxygen species
γ-TMTγ-tocopherol methyltransferase
RT-qPCRquantitative reverse transcription PCR
SAM S-adenosylmethionine
SD standard deviation
TATtyrosine aminotransferase
TCtocopherol cyclase
ZS11Zhongshuang11

References

  1. Chalhoub, B.; Denoeud, F.; Liu, S.; Parkin, I.A.P.; Tang, H.; Wang, X.; Chiquet, J.; Belcram, H.; Tong, C.; Samans, B.; et al. Early allopolyploid evolution in the post-Neolithic Brassica napus oilseed genome. Science 2014, 345, 950–953. [Google Scholar] [CrossRef] [Green Version]
  2. Fu, T.D. Variety improvement in rapeseed. Crop. Res. 2007, 3, 159–162. (In Chinese) [Google Scholar]
  3. Nestel, P.; Clifton, P.; Colquhoun, D.; Noakes, M.; Mori, T.A.; Sullivan, D.; Thomas, B. Indications for omega-3 long chain polyunsaturated fatty acid in the prevention and treatment of cardiovascular disease. Heart Lung Circ. 2015, 24, 769–779. [Google Scholar] [CrossRef] [Green Version]
  4. Zhao, J.V.; Schooling, C.M. Role of linoleic acid in autoimmune disorders: A Mendelian randomisation study. Ann. Rheum. Dis. 2019, 78, 711. [Google Scholar] [CrossRef]
  5. Djuricic, I.; Calder, P.C. Beneficial outcomes of omega-6 and omega-3 polyunsaturated fatty acids on human health: An update for 2021. Nutrients 2021, 13, 2421. [Google Scholar] [CrossRef]
  6. Weinstein, S.J.; Wright, M.E.; Lawson, K.A.; Snyder, K.; Mannisto, S.; Taylor, P.R.; Virtamo, J.; Albanes, D. Serum and dietary vitamin E in relation to prostate cancer risk. Cancer Epidemiol. Biomark. Prev. 2007, 16, 1253–1259. [Google Scholar] [CrossRef] [Green Version]
  7. Euch-Fayache, G.E.; Bouhlal, Y.; Amouri, R.; Feki, M.; Hentati, F. Molecular, clinical and peripheral neuropathy study of Tunisian patients with ataxia with vitamin E deficiency. Brain 2014, 137, 402–410. [Google Scholar] [CrossRef]
  8. Hunter, S.; Cahoon, E. Enhancing vitamin E in oilseeds: Unraveling tocopherol and tocotrienol biosynthesis. Lipids 2007, 42, 97–108. [Google Scholar] [CrossRef]
  9. Horvath, G.; Wessjohann, L.; Bigirimana, J.; Jansen, M.; Guisez, Y.; Caubergs, R.; Horemans, N. Differential distribution of tocopherols and tocotrienols in photosynthetic and non-photosynthetic tissues. Phytochemistry 2006, 67, 1185–1195. [Google Scholar] [CrossRef]
  10. Munné-Bosch, S.; Alegre, L. The function of tocopherols and tocotrienols in plants. Crit. Rev. Plant Sci. 2002, 21, 31–57. [Google Scholar] [CrossRef]
  11. DellaPenna, D.; Last, R.L. Progress in the dissection and manipulation of plant vitamin E biosynthesis. Physiol. Plant. 2006, 126, 356–368. [Google Scholar] [CrossRef]
  12. Garcia, I.; Rodgers, M.; Lenne, C.; Rolland, A.; Sailland, A.; Matringe, M. Subcellular localization and purification of a p-hydroxyphenylpyruvate dioxygenase from cultured carrot cells and characterization of the corresponding cDNA. Biochem. J. 1997, 325, 761–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Riewe, D.; Koohi, M.; Lisec, J.; Pfeiffer, M.; Lippmann, R.; Schmeichel, J.; Willmitzer, L.; Altmann, T. A tyrosine aminotransferase involved in tocopherol synthesis in Arabidopsis. Plant J. 2012, 71, 850–859. [Google Scholar] [CrossRef] [PubMed]
  14. Keller, Y.; Bouvier, F.; D’Harlingue, A.; Camara, B. Metabolic compartmentation of plastid prenyllipid biosynthesis evidence for the involvement of a multifunctional geranylgeranyl reductase. Eur. J. Biochem. 1998, 251, 413–417. [Google Scholar] [CrossRef] [PubMed]
  15. Kimura, E.; Abe, T.; Murata, K.; Kimura, T.; Otoki, Y.; Yoshida, T.; Miyazawa, T.; Nakagawa, K. Identification of OsGGR2, a second geranylgeranyl reductase involved in α-tocopherol synthesis in rice. Sci. Rep. 2018, 8, 1870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Valentin, H.E.; Lincoln, K.; Moshiri, F.; Jensen, P.K.; Qi, Q.; Venkatesh, T.V.; Karunanandaa, B.; Baszis, S.R.; Norris, S.R.; Savidge, B.; et al. The Arabidopsis vitamin E pathway gene5-1 mutant reveals a critical role for phytol kinase in seed tocopherol biosynthesis. Plant Cell 2006, 18, 212–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Dorp, K.V.; Hölzl, G.; Plohmann, C.; Eisenhut, M.; Abraham, M.; Weber, A.P.M.; Hanson, A.D.; Dörmann, P. Remobilization of phytol from chlorophyll degradation is essential for tocopherol synthesis and growth of Arabidopsis. Plant Cell 2015, 27, 2846–2859. [Google Scholar]
  18. Mène-Saffrané, L.; DellaPenna, D. Biosynthesis, regulation and functions of tocochromanols in plants. Plant Physiol. Bioch. 2010, 48, 301–309. [Google Scholar] [CrossRef] [PubMed]
  19. Fritsche, S.; Wang, X.; Jung, C. Recent advances in our understanding of tocopherol biosynthesis in plants: An overview of key genes, functions, and breeding of vitamin E improved crops. Antioxidants 2017, 6, 99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Shintani, D.; DellaPenna, D. Elevating the vitamin E content of plants through metabolic engineering. Science 1998, 282, 2098–2100. [Google Scholar] [CrossRef]
  21. Bergmüller, E.; Porfirova, S.; Dörmann, P. Characterization of an Arabidopsis mutant deficient in γ-tocopherol methyltransferase. Plant Mol. Biol. 2003, 52, 1181–1190. [Google Scholar] [CrossRef]
  22. Kim, Y.J.; Seo, H.Y.; Park, T.I.; Baek, S.H.; Shin, W.C.; Kim, H.S.; Kim, J.G.; Choi, Y.E.; Yun, S.J. Enhanced biosynthesis of α-tocopherol in transgenic soybean by introducing γ-TMT gene. Plant Biotechnol. J. 2005, 7, 203–209. [Google Scholar]
  23. Kim, M.J.; Baek, S.H.; Yoo, N.H.; Yun, S.J. Transformation of Arabidopsis gamma-tocopherol methyltransferase into lettuce (Lactuca sativa L.). Korean J. Plant Tiss. Cult. 2000, 27, 435–439. [Google Scholar]
  24. Cho, E.A.; Lee, C.A.; Kim, Y.S.; Baek, S.H.; de Los, R.B.; Yun, S.J. Expression of γ-tocopherol methyltransferase transgene improves tocopherol composition in lettuce (Latuca sativa L.). Mol. Cell 2005, 19, 16–22. [Google Scholar]
  25. Zhang, G.Y.; Liu, R.R.; Xu, G.; Zhang, P.; Li, Y.; Tang, K.X.; Liang, G.H.; Liu, Q.Q. Increased α-tocotrienol content in seeds of transgenic rice overexpressing Arabidopsis γ-tocopherol methyltransferase. Transgenic Res. 2013, 22, 89–99. [Google Scholar] [CrossRef]
  26. Ghimire, B.K.; Seong, E.S.; Lee, C.O.; Lim, J.D.; Lee, J.G.; Yoo, J.H.; Chung, I.M.; Kim, N.Y.; Yu, C.Y. Enhancement of α-tocopherol content in transgenic Perilla frutescens containing the γ-TMT gene. Afr. J. Biotechnol. 2011, 10, 2430–2439. [Google Scholar]
  27. Tavva, V.S.; Kim, Y.; Kagan, I.A.; Dinkins, R.D.; Kim, K.; Collins, G.B. Increased α-tocopherol content in soybean seed overexpressing the Perilla frutescens γ-tocopherol methyltransferase gene. Plant Cell Rep. 2007, 26, 61–70. [Google Scholar] [CrossRef]
  28. Yusuf, M.A.; Sarin, N.B. Antioxidant value addition in human diets: Genetic transformation of Brassica juncea with γ-TMT gene for increased α-tocopherol content. Transgenic Res. 2007, 16, 109–113. [Google Scholar] [CrossRef]
  29. Grusak, M.A.; DellaPenna, D. Improving the nutrient composition of plants to enhance human nutrition and health. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1999, 50, 133–161. [Google Scholar] [CrossRef]
  30. Krieger-Liszkay, A.; Fufezan, C.; Trebst, A. Singlet oxygen production in photosystem II and related protection mechanism. Photosynth. Res. 2008, 98, 551–564. [Google Scholar] [CrossRef]
  31. Liebler, D.C. The role of metabolism in the antioxidant function of vitamin E. Crit. Rev. Toxicol. 1993, 23, 147–169. [Google Scholar] [CrossRef] [PubMed]
  32. Kamal-Eldin, A.; Appelqvist, L.A. The chemistry and antioxidant properties of tocopherols and tocotrienols. Lipids 1996, 31, 671–701. [Google Scholar] [CrossRef] [PubMed]
  33. Kamal-Eldin, A.; Andersson, R. A multivariate study of the correlation between tocopherol content and fatty acid composition in vegetable oils. J. Am. Oil Chem. Soc. 1997, 74, 375–380. [Google Scholar] [CrossRef]
  34. Goffman, F.D.; Böhme, T. Relationship between fatty acid profile and vitamin E content in maize hybrids (Zea mays L.). J. Agric. Food Chem. 2001, 49, 4990–4994. [Google Scholar] [CrossRef]
  35. Li, Y.; Hussain, N.; Zhang, L.; Chen, X.; Ali, E.; Jiang, L. Correlations between tocopherol and fatty acid components in germplasm collections of Brassica oilseeds. J. Agri. Food Chem. 2013, 61, 34–40. [Google Scholar] [CrossRef]
  36. Sattler, S.E.; Gilliland, L.U.; Magallanes-Lundback, M.; Pollard, M.; DellaPenna, D. Vitamin E is essential for seed longevity and for preventing lipid peroxidation during germination. Plant Cell 2004, 16, 1419–1432. [Google Scholar] [CrossRef]
  37. Munné-Bosch, S.; Weiler, E.W.; Alegre, L.; Müller, M.; Düchting, P.; Falk, J. α-Tocopherol may influence cellular signaling by modulating jasmonic acid levels in plants. Planta 2007, 225, 681–691. [Google Scholar] [CrossRef]
  38. Stahl, E.; Hartmann, M.; Scholten, N.; Zeier, J. A role for tocopherol biosynthesis in Arabidopsis basal immunity to bacterial infection. Plant Physiol. 2019, 181, 1008–1028. [Google Scholar] [CrossRef] [Green Version]
  39. Abbasi, A.R.; Hajirezaei, M.; Hofius, D.; Sonnewald, U.; Voll, L.M. Specific roles of α- and γ-tocopherol in abiotic stress responses of transgenic tobacco. Plant Physiol. 2007, 143, 1720–1738. [Google Scholar] [CrossRef] [Green Version]
  40. Yusuf, M.A.; Kumar, D.; Rajwanshi, R.; Strasser, R.J.; Tsimilli-Michael, M.; Govindjee; Sarin, N.B. Overexpression of γ-tocopherol methyl transferase gene in transgenic Brassica juncea plants alleviates abiotic stress: Physiological and chlorophyll a fluorescence measurements. Biochim. Biophys. Acta Bioenerg. 2010, 1797, 1428–1438. [Google Scholar] [CrossRef] [Green Version]
  41. Cela, J.; Chang, C.; Munné-Bosch, S. Accumulation of γ- rather than α-tocopherol alters ethylene signaling gene expression in the vte4 mutant of Arabidopsis thaliana. Plant Cell Physiol. 2011, 52, 1389–1400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Ellouzi, H.; Hamed, K.B.; Cela, J.; Müller, M.; Abdelly, C.; Munné-Bosch, S. Increased sensitivity to salt stress in tocopherol-deficient Arabidopsis mutants growing in a hydroponic system. Plant Signal Behav. 2013, 8, e23136. [Google Scholar] [CrossRef] [Green Version]
  43. Jin, S.; Daniell, H. Expression of γ-tocopherol methyltransferase in chloroplasts results in massive proliferation of the inner envelope membrane and decreases susceptibility to salt and metal-induced oxidative stresses by reducing reactive oxygen species. Plant Biotechnol. J. 2014, 12, 1274–1285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ma, J.; Qiu, D.; Xu, L.; Cui, M.; Gao, H.; Pang, Y.; Qin, Y.; Wang, X. Overexpression of alfalfa γ-tocopherol methyltransferase (γ-TMT) gene increases salt susceptibility of transgenic Arabidopsis in seed germination. Environ. Exp. Bot. 2020, 180, 104264. [Google Scholar] [CrossRef]
  45. Maeda, H.; Song, W.; Sage, T.L.; DellaPenna, D. Tocopherols play a crucial role in low-temperature adaptation and phloem loading in Arabidopsis. Plant Cell 2006, 18, 2710–2732. [Google Scholar] [CrossRef] [Green Version]
  46. Maeda, H.; Sage, T.L.; Isaac, G.; Welti, R.; DellaPenna, D. Tocopherols modulate extraplastidic polyunsaturated fatty acid metabolism in Arabidopsis at low temperature. Plant Cell 2008, 20, 452–470. [Google Scholar] [CrossRef]
  47. Ma, J.; Qiu, D.; Gao, H.; Wen, H.; Wu, Y.; Pang, Y.; Wang, X.; Qin, Y. Over-expression of a γ-tocopherol methyltransferase gene in vitamin E pathway confers PEG-simulated drought tolerance in alfalfa. BMC Plant Biol. 2020, 20, 226. [Google Scholar] [CrossRef]
  48. Collin, V.C.; Eymery, F.; Genty, B.; Rey, P.; Havaux, M. Vitamin E is essential for the tolerance of Arabidopsis thaliana to metal-induced oxidative stress. Plant Cell Environ. 2008, 31, 244–257. [Google Scholar] [CrossRef]
  49. Zhu, Q.; Zhang, J.; Yu, H.; Li, L.; Chen, X.; Jiang, M.; Tan, M. Maize Cd-tolerant ZmVTE4 encoding γ-tocopherol-methyl-transferase alleviated Cd-toxicity through its product α-tocopherol. Environ. Exp. Bot. 2019, 158, 171–179. [Google Scholar] [CrossRef]
  50. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 2002, 7, 405–410. [Google Scholar] [CrossRef]
  51. Miller, G.; Suzuki, N.; Ciftci-Yilmaz, S.; Mittler, R. Reactive oxygen species homeostasis and signalling during drought and salinity stresses. Plant Cell Environ. 2010, 33, 453–467. [Google Scholar] [CrossRef]
  52. Schranz, M.E.; Lysak, M.A.; Mitchell-olds, T. The ABC’s of comparative genomics in the Brassicaceae: Building blocks of crucifer genomes. Trends Plant Sci. 2006, 11, 535–542. [Google Scholar] [CrossRef]
  53. Grillo, M.A.; Colombatto, S. S-adenosylmethionine and its products. Amino Acids 2008, 34, 187–193. [Google Scholar] [CrossRef] [PubMed]
  54. Goffman, F.D.; Velasco, L.; Becker, H.C. Tocopherols accumulation in developing seeds and pods of rapeseed (Brassica napus L.). Lipid/Fett 1999, 101, 400–403. [Google Scholar] [CrossRef]
  55. Endrigkeit, J.; Wang, X.; Cai, D.; Zhang, C.; Long, Y.; Meng, J.; Jung, C. Genetic mapping, cloning, and functional characterization of the BnaX.VTE4 gene encoding a γ-tocopherol methyltransferase from oilseed rape. Theor. Appl. Genet. 2009, 119, 567–575. [Google Scholar] [CrossRef] [PubMed]
  56. Okuley, J.; Lightner, J.; Feldmann, K.; Yadav, N.; Lark, E.; Browse, J. Arabidopsis FAD2 gene encodes the enzyme that is essential for polyunsaturated lipid synthesis. Plant Cell 1994, 6, 147–158. [Google Scholar]
  57. Lee, K.R.; Lee, Y.; Kim, E.H.; Lee, S.B.; Roh, K.H.; Kim, J.B.; Kang, H.C.; Kim, H.U. Functional identification of oleate 12-desaturase and ω-3 fatty acid desaturase genes from Perilla frutescens var. Frutescens. Plant Cell Rep. 2016, 35, 2523–2537. [Google Scholar] [CrossRef]
  58. Shah, S.; Xin, Z.G.; Browse, J. Overexpression of the FAD3 desaturase gene in a mutant of Arabidopsis. Plant Physiol. 1997, 114, 1533–1539. [Google Scholar] [CrossRef] [Green Version]
  59. Peng, Z.; Ruan, J.; Tian, H.; Shan, L.; Meng, J.; Guo, F.; Zhang, Z.; Ding, H.; Wan, S.; Li, X. The family of peanut fatty acid desaturase genes and a functional analysis of four ω-3 AhFAD3 members. Plant Mol. Biol. Rep. 2020, 38, 209–221. [Google Scholar] [CrossRef] [Green Version]
  60. Choudhury, F.K.; Rivero, R.M.; Blumwald, E.; Mittler, R. Reactive oxygen species, abiotic stress and stress combination. Plant J. 2017, 90, 856–867. [Google Scholar] [CrossRef]
  61. Miller, G.; Shulaev, V.; Mittler, R. Reactive oxygen signaling and abiotic stress. Physiol. Plant. 2008, 133, 481–489. [Google Scholar] [CrossRef] [PubMed]
  62. Ohlrogge, J.B.; Jaworski, J.G. Regulation of fatty acid synthesis. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1997, 48, 109–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Hong, J.K.; Choi, H.W.; Hwang, I.S.; Kim, D.S.; Kim, N.H.; Choi, D.S.; Kim, Y.J.; Hwang, B.K. Function of a novel GDSL-type pepper lipase gene, CaGLIP1, in disease susceptibility and abiotic stress tolerance. Planta 2008, 227, 539–558. [Google Scholar] [CrossRef] [PubMed]
  64. Mu, J.; Tan, H.; Zheng, Q.; Fu, F.; Liang, Y.; Zhang, J.; Yang, X.; Wang, T.; Chong, K.; Wang, X.J.; et al. LEAFY COTYLEDON1 is a key regulator of fatty acid biosynthesis in Arabidopsis. Plant Physiol. 2008, 148, 1042–1054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Mueller, M.J. Enzymes involved in jasmonic acid biosynthesis. Physiol. Plant. 1997, 100, 653–663. [Google Scholar] [CrossRef]
  66. Wang, J.; Song, L.; Gong, X.; Xu, J.; Li, M. Functions of jasmonic acid in plant regulation and response to abiotic stress. Int. J. Mol. Sci. 2020, 21, 1446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Cortés, A.J.; López-Hernández, F. Harnessing crop wild diversity for climate change adaptation. Genes 2021, 12, 783. [Google Scholar] [CrossRef] [PubMed]
  68. Cortés, A.J.; López-Hernández, F.; Blair, M.W. Genome-environment associations, an innovative tool for studying heritable evolutionary adaptation in orphan crops and wild relatives. Front. Genet. 2022, 13, 910386. [Google Scholar] [CrossRef]
  69. Burbano-Erazo, E.; León-Pacheco, R.I.; Cordero-Cordero, C.C.; López-Hernández, F.; Cortés, A.J.; Tofiño-Rivera, A.P. Multi-environment yield components in advanced common bean (Phaseolus vulgaris L.) × Tepary bean (P. acutifolius A. Gray) interspecific lines for heat and drought tolerance. Agronomy 2021, 11, 1978. [Google Scholar] [CrossRef]
  70. Liu, C.; Chen, H.; Er, H.L.; Soo, H.M.; Kumar, P.P.; Han, J.; Liou, Y.C.; Yu, H. Direct interaction of AGL24 and SOC1 integrates flowering signals in Arabidopsis. Development 2008, 135, 1481–1491. [Google Scholar] [CrossRef] [Green Version]
  71. Cardoza, V.; Stewart, C.N. Increased Agrobacterium-mediated transformation and rooting efficiencies in canola (Brassica napus L.) from hypocotyl segment explants. Plant Cell Rep. 2003, 21, 599–604. [Google Scholar] [CrossRef] [PubMed]
  72. Poirier, Y.; Ventre, G.; Caldelari, D. Increased flow of fatty acids toward β-oxidation in developing seeds of Arabidopsis deficient in diacylglycerol acyltransferase activity or synthesizing medium-chain-length fatty acids. Plant Physiol. 1999, 121, 1359–1366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Chen, M.; Wang, Z.; Zhu, Y.; Li, Z.; Hussain, N.; Xuan, L.; Guo, W.; Zhang, G.; Jiang, L. The effect of transparent TESTA2 on seed fatty acid biosynthesis and tolerance to environmental stresses during young seedling establishment in Arabidopsis. Plant Physiol. 2012, 160, 1023–1036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Yu, Y.; Wang, J.; Li, S.; Kakan, X.; Zhou, Y.; Miao, Y.; Wang, F.; Qin, H.; Huang, R. Ascorbic acid integrates the antagonistic modulation of ethylene and abscisic acid in the accumulation of reactive oxygen species. Plant Physiol. 2019, 179, 1861–1875. [Google Scholar] [CrossRef] [Green Version]
  75. Kakan, X.; Yu, Y.; Li, S.; Li, X.; Huang, R.; Wang, J. Ascorbic acid modulation by ABI4 transcriptional repression of VTC2 in the salt tolerance of Arabidopsis. BMC Plant Biol. 2021, 21, 112. [Google Scholar] [CrossRef] [PubMed]
  76. Luo, X.; Dai, Y.; Zheng, C.; Yang, Y.; Chen, W.; Wang, Q.; Chandrasekaran, U.; Du, J.; Liu, W.; Shu, K. The ABI4-RbohD/VTC2 regulatory module promotes Reactive Oxygen Species (ROS) accumulation to decrease seed germination under salinity stress. New Phytol. 2021, 229, 950–962. [Google Scholar] [CrossRef] [PubMed]
  77. Assmann, S.M. G protein signaling in the regulation of Arabidopsis seed germination. Sci. STKE 2005, 308, cm11. [Google Scholar] [CrossRef]
  78. Pfaffl, M.W.; Horgan, G.W.; Dempfle, L. Relative expression software tool (REST©) for group-wise comparison and statistical analysis of relative expression results in real-time PCR. Nucleic Acids Res. 2002, 30, e36. [Google Scholar] [CrossRef]
  79. Küpper, H.; Seib, L.O.; Sivaguru, M.; Hoekenga, O.A.; Kochian, L.V. A method for cellular localization of gene expression via quantitative in situ hybridization in plants. Plant J. 2007, 50, 159–175. [Google Scholar] [CrossRef]
Figure 1. Simplified schematic diagram showing the tocopherols biosynthetic pathway in plants: ① shikimic pathway; ② methylerythritol phosphate pathway; ③ phytol recycling pathway. TAT: tyrosine aminotransferase; HPP: p-hydroxyphenylpyruvic acid; PDS1 (HPPD): HPP dioxygenase; HGA: homogentisic acid; VTE5 (PK): phytol kinase; Phytyl-P: phytyl phosphate; VTE6 (PPK): phytyl phosphate kinase; GGPP: geranylgeranyl diphosphate; GGR: geranylgeranyl reductase; PDP: phytyl diphosphate; VTE2 (HPT): HGA phytyl transferase; MPBQ: 2-methyl-6-phytyl-1,4-benzoquinol; VTE3 (MPBQ-MT): MPBQ methyltransferase; SAM: S-adenosylmethionine; DMBPQ: 2,3-dimethyl-6-phytyl-1,4-benzoquinol; VTE1 (TC): tocopherol cyclase; VTE4 (γ-TMT): γ-tocopherol methyltransferase.
Figure 1. Simplified schematic diagram showing the tocopherols biosynthetic pathway in plants: ① shikimic pathway; ② methylerythritol phosphate pathway; ③ phytol recycling pathway. TAT: tyrosine aminotransferase; HPP: p-hydroxyphenylpyruvic acid; PDS1 (HPPD): HPP dioxygenase; HGA: homogentisic acid; VTE5 (PK): phytol kinase; Phytyl-P: phytyl phosphate; VTE6 (PPK): phytyl phosphate kinase; GGPP: geranylgeranyl diphosphate; GGR: geranylgeranyl reductase; PDP: phytyl diphosphate; VTE2 (HPT): HGA phytyl transferase; MPBQ: 2-methyl-6-phytyl-1,4-benzoquinol; VTE3 (MPBQ-MT): MPBQ methyltransferase; SAM: S-adenosylmethionine; DMBPQ: 2,3-dimethyl-6-phytyl-1,4-benzoquinol; VTE1 (TC): tocopherol cyclase; VTE4 (γ-TMT): γ-tocopherol methyltransferase.
Ijms 23 15933 g001
Figure 2. Alignment of amino acid sequences of γ-TMT derived from Arabidopsis thaliana and Brassica napus. The alignment was conducted using the MUSCLE program (http://www.ebi.ac.uk/Tools/msa/muscle/, accessed on 8 September 2022), and the different amino acids are indicated by asterisks. S-adenosylmethionine binding sites are underlined.
Figure 2. Alignment of amino acid sequences of γ-TMT derived from Arabidopsis thaliana and Brassica napus. The alignment was conducted using the MUSCLE program (http://www.ebi.ac.uk/Tools/msa/muscle/, accessed on 8 September 2022), and the different amino acids are indicated by asterisks. S-adenosylmethionine binding sites are underlined.
Ijms 23 15933 g002
Figure 3. Phylogenetic tree analysis of γ-TMT proteins derived from A. thaliana and seven oil crops. A neighbour-joining tree (Jones–Taylor–Thornton model) was generated by MEGA7. A bootstrap analysis with 1000 replicates was performed to assess the statistical reliability of the tree topology. The accession numbers corresponding to the species names are listed as follows: A. thaliana, Atγ-TMT (NP_176677.1); B. napus, BnaC02.TMT.a (BnaC02T0331100ZS), BnaC02.TMT.b (BnaC02T0197500ZS), BnaA02.TMT.a (BnaA02T0247300ZS), and BnaA02.TMT.b (BnaA02T0154300ZS); Zea mays, Zmγ-TMT (NP_001105914.1); Elaeis guineensis, Egγ-TMT (XP_010906229.1); Linum usitatissimum, Luγ-TMT (ACO57634.1); Glycine max, Gmγ-TMT-1 (NP_001304355.1) and Gmγ-TMT-2 (NP_001236725.1); Gossypium raimondii, Grγ-TMT-1 (XP_012470094.1) and Grγ-TMT-2 (XP_012464711.1); Arachis hypogaea, Ahγ-TMT-1 (XP_025699759.1), Ahγ-TMT-2 (XP_025654031.1), and Ahγ-TMT-3 (XP_025654033.1).
Figure 3. Phylogenetic tree analysis of γ-TMT proteins derived from A. thaliana and seven oil crops. A neighbour-joining tree (Jones–Taylor–Thornton model) was generated by MEGA7. A bootstrap analysis with 1000 replicates was performed to assess the statistical reliability of the tree topology. The accession numbers corresponding to the species names are listed as follows: A. thaliana, Atγ-TMT (NP_176677.1); B. napus, BnaC02.TMT.a (BnaC02T0331100ZS), BnaC02.TMT.b (BnaC02T0197500ZS), BnaA02.TMT.a (BnaA02T0247300ZS), and BnaA02.TMT.b (BnaA02T0154300ZS); Zea mays, Zmγ-TMT (NP_001105914.1); Elaeis guineensis, Egγ-TMT (XP_010906229.1); Linum usitatissimum, Luγ-TMT (ACO57634.1); Glycine max, Gmγ-TMT-1 (NP_001304355.1) and Gmγ-TMT-2 (NP_001236725.1); Gossypium raimondii, Grγ-TMT-1 (XP_012470094.1) and Grγ-TMT-2 (XP_012464711.1); Arachis hypogaea, Ahγ-TMT-1 (XP_025699759.1), Ahγ-TMT-2 (XP_025654031.1), and Ahγ-TMT-3 (XP_025654033.1).
Ijms 23 15933 g003
Figure 4. RT-qPCR analysis of the Bnγ-TMT expression in various tissues and developing seeds at different developmental stages in the B. napus cultivar “ZS11”. The RT-qPCR results were normalised against BnGAPDH expression as an internal control, and the expression level of Bnγ-TMT in roots was set to 1. The values indicate the means of three replicates of dilutions of cDNA obtained from three independent RNA extractions. The error bars denote standard deviations.
Figure 4. RT-qPCR analysis of the Bnγ-TMT expression in various tissues and developing seeds at different developmental stages in the B. napus cultivar “ZS11”. The RT-qPCR results were normalised against BnGAPDH expression as an internal control, and the expression level of Bnγ-TMT in roots was set to 1. The values indicate the means of three replicates of dilutions of cDNA obtained from three independent RNA extractions. The error bars denote standard deviations.
Ijms 23 15933 g004
Figure 5. Characterisation of 35S:BnaC02.TMT.a-6HA lines. (A) Schematic diagram of the constitutive expression cassette of the 35S:BnaC02.TMT.a-6HA gene in the binary vector pGreen 2 × 35S used for plant transformation. RB, right border; LB, left border; NOS-pro, nopaline synthase promoter; NOS-ter, nopaline synthase terminator; Basta, glyphosate; 35S-pro, CaMV 35S promoter. (B) PCR-based genotyping of the 35S:BnaC02.TMT.a-6HA transgenic plants using specific primers for the 35S_P/BnTMT_R. Cas, cassette. (C) Expression analysis of Bnγ-TMT in the 35S:BnaC02.TMT.a-6HA transgenic plants using RT-qPCR. The RT-qPCR results were normalised against the expression of BnGAPDH, which was used as an internal control. The values are the means ± SD (n = 3). (D,E) The levels of α-tocopherol and total tocopherol in the mature seeds of the transgenic lines and control plants (ZS11). The asterisks denote statistically significant differences between the wild-type ZS11 and the 35S:BnaC02.TMT.a-6HA transgenic plants (Student’s t-test, p ≤ 0.05). The error bars denote the standard deviation.
Figure 5. Characterisation of 35S:BnaC02.TMT.a-6HA lines. (A) Schematic diagram of the constitutive expression cassette of the 35S:BnaC02.TMT.a-6HA gene in the binary vector pGreen 2 × 35S used for plant transformation. RB, right border; LB, left border; NOS-pro, nopaline synthase promoter; NOS-ter, nopaline synthase terminator; Basta, glyphosate; 35S-pro, CaMV 35S promoter. (B) PCR-based genotyping of the 35S:BnaC02.TMT.a-6HA transgenic plants using specific primers for the 35S_P/BnTMT_R. Cas, cassette. (C) Expression analysis of Bnγ-TMT in the 35S:BnaC02.TMT.a-6HA transgenic plants using RT-qPCR. The RT-qPCR results were normalised against the expression of BnGAPDH, which was used as an internal control. The values are the means ± SD (n = 3). (D,E) The levels of α-tocopherol and total tocopherol in the mature seeds of the transgenic lines and control plants (ZS11). The asterisks denote statistically significant differences between the wild-type ZS11 and the 35S:BnaC02.TMT.a-6HA transgenic plants (Student’s t-test, p ≤ 0.05). The error bars denote the standard deviation.
Ijms 23 15933 g005
Figure 6. Effect of BnaC02.TMT.a on FA metabolism in seeds. (A) Comparison of the FA compositions of major seeds between the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants. The asterisks indicate significant differences in the total FA levels in the seeds (two-tailed paired Student’s t-test, p ≤ 0.05). The error bars denote the standard deviation. (B) Comparison of the expression of genes contributing to the FA modification in the developing seeds of the ZS11 and BnaC02.TMT.a-6HA plants. The asterisks indicate significant differences between the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants (two-tailed paired Student’s t-test, p ≤ 0.05). The error bars denote the standard deviation.
Figure 6. Effect of BnaC02.TMT.a on FA metabolism in seeds. (A) Comparison of the FA compositions of major seeds between the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants. The asterisks indicate significant differences in the total FA levels in the seeds (two-tailed paired Student’s t-test, p ≤ 0.05). The error bars denote the standard deviation. (B) Comparison of the expression of genes contributing to the FA modification in the developing seeds of the ZS11 and BnaC02.TMT.a-6HA plants. The asterisks indicate significant differences between the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants (two-tailed paired Student’s t-test, p ≤ 0.05). The error bars denote the standard deviation.
Ijms 23 15933 g006
Figure 7. Phenotypes of the seed germination at 96 h after sowing (HAS) in the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants under control conditions and after treatment with 180 mM NaCl. The emergence of a visible radicle was used as a morphological marker for seed germination. Bar = 1 cm.
Figure 7. Phenotypes of the seed germination at 96 h after sowing (HAS) in the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants under control conditions and after treatment with 180 mM NaCl. The emergence of a visible radicle was used as a morphological marker for seed germination. Bar = 1 cm.
Ijms 23 15933 g007
Figure 8. Time course analysis of seed germination in the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants under control conditions and after treatment with 180 mM NaCl. Seed germination was scored every eight hours, and the emergence of a visible radicle was used as a morphological marker. The data are expressed as the mean ± SD from three independent experiments evaluating 50 individual seeds.
Figure 8. Time course analysis of seed germination in the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants under control conditions and after treatment with 180 mM NaCl. Seed germination was scored every eight hours, and the emergence of a visible radicle was used as a morphological marker. The data are expressed as the mean ± SD from three independent experiments evaluating 50 individual seeds.
Ijms 23 15933 g008
Figure 9. Effect of BnaC02.TMT.a on ROS accumulation. (A) diaminobenzidine (DAB, Left) and nitroblue tetrazolium (NBT, right) staining of ROS in the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants treated with 180 mM NaCl at 96 h after sowing (HAS). Bar = 1 cm; (B) measurements of H2O2 (Left) and O2 (right) in the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants in response to treatment with 180 mM NaCl at 96 HAS. The asterisks indicate significant differences between the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants under control conditions or after 180 mM NaCl treatment, respectively (two-tailed paired Student’s t-test, p ≤ 0.05). The error bars denote the standard deviation. FW, fresh weight.
Figure 9. Effect of BnaC02.TMT.a on ROS accumulation. (A) diaminobenzidine (DAB, Left) and nitroblue tetrazolium (NBT, right) staining of ROS in the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants treated with 180 mM NaCl at 96 h after sowing (HAS). Bar = 1 cm; (B) measurements of H2O2 (Left) and O2 (right) in the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants in response to treatment with 180 mM NaCl at 96 HAS. The asterisks indicate significant differences between the wild-type ZS11 and 35S:BnaC02.TMT.a-6HA plants under control conditions or after 180 mM NaCl treatment, respectively (two-tailed paired Student’s t-test, p ≤ 0.05). The error bars denote the standard deviation. FW, fresh weight.
Ijms 23 15933 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, Y.; Li, D.; Liu, T.; Liao, M.; Li, Y.; Zhang, W.; Liu, Z.; Chen, M. Effect of Overexpression of γ-Tocopherol Methyltransferase on α-Tocopherol and Fatty Acid Accumulation and Tolerance to Salt Stress during Seed Germination in Brassica napus L. Int. J. Mol. Sci. 2022, 23, 15933. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232415933

AMA Style

Guo Y, Li D, Liu T, Liao M, Li Y, Zhang W, Liu Z, Chen M. Effect of Overexpression of γ-Tocopherol Methyltransferase on α-Tocopherol and Fatty Acid Accumulation and Tolerance to Salt Stress during Seed Germination in Brassica napus L. International Journal of Molecular Sciences. 2022; 23(24):15933. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232415933

Chicago/Turabian Style

Guo, Yuan, Dong Li, Tiantian Liu, Meifang Liao, Yuxin Li, Weitang Zhang, Zijin Liu, and Mingxun Chen. 2022. "Effect of Overexpression of γ-Tocopherol Methyltransferase on α-Tocopherol and Fatty Acid Accumulation and Tolerance to Salt Stress during Seed Germination in Brassica napus L." International Journal of Molecular Sciences 23, no. 24: 15933. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232415933

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