Next Article in Journal
ω-3 Long Chain Polyunsaturated Fatty Acids as Sensitizing Agents and Multidrug Resistance Revertants in Cancer Therapy
Previous Article in Journal
Addendum: Cechová, M. et al. Towards Better Understanding of Pea Seed Dormancy Using Laser Desorption/Ionization Mass Spectrometry. Int. J. Mol. Sci. 2017, 18, 2196
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genome-Wide Identification, Evolutionary and Expression Analyses of the GALACTINOL SYNTHASE Gene Family in Rapeseed and Tobacco

1
College of Agronomy and Biotechnology, Southwest University, Chongqing 400715, China
2
Academy of Agricultural Sciences, Southwest University, Chongqing 400715, China
3
Key Laboratory of Molecular Genetics, China National Tobacco Corporation, Guizhou Academy of Tobacco Science, Guiyang 550081, China
4
Upland Flue-Cured Tobacco Quality and Ecology Key Laboratory of China Tobacco, Guizhou Academy of Tobacco Science, Guiyang 550081, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2017, 18(12), 2768; https://doi.org/10.3390/ijms18122768
Submission received: 30 September 2017 / Revised: 16 November 2017 / Accepted: 17 December 2017 / Published: 20 December 2017
(This article belongs to the Section Biochemistry)

Abstract

:
Galactinol synthase (GolS) is a key enzyme in raffinose family oligosaccharide (RFO) biosynthesis. The finding that GolS accumulates in plants exposed to abiotic stresses indicates RFOs function in environmental adaptation. However, the evolutionary relationships and biological functions of GolS family in rapeseed (Brassica napus) and tobacco (Nicotiana tabacum) remain unclear. In this study, we identified 20 BnGolS and 9 NtGolS genes. Subcellular localization predictions showed that most of the proteins are localized to the cytoplasm. Phylogenetic analysis identified a lost event of an ancient GolS copy in the Solanaceae and an ancient duplication event leading to evolution of GolS4/7 in the Brassicaceae. The three-dimensional structures of two GolS proteins were conserved, with an important DxD motif for binding to UDP-galactose (uridine diphosphate-galactose) and inositol. Expression profile analysis indicated that BnGolS and NtGolS genes were expressed in most tissues and highly expressed in one or two specific tissues. Hormone treatments strongly induced the expression of most BnGolS genes and homologous genes in the same subfamilies exhibited divergent-induced expression. Our study provides a comprehensive evolutionary analysis of GolS genes among the Brassicaceae and Solanaceae as well as an insight into the biological function of GolS genes in hormone response in plants.

1. Introduction

Abiotic stress, such as drought, salinity, low temperature, can seriously affect the normal growth and development of plants, especially the yield of crops [1,2]. In response to environmental changes, plants formed a set of physiological, biochemical, cellular and molecular mechanisms to deal and cope metabolically with water deficit periods [3]. To improve the stress resistances and reduce the possibly damages from abiotic stresses, plants have evolved complex strategies, such as the accumulation of certain small soluble molecules (such as trehalose, proline, tryptophan, polyamines and glycyl betaine), to increase the osmotic pressure in cells [4,5]. Acting as osmoprotectants, raffinose family oligosaccharides (RFOs), including raffinose, stachyose and verbascose, play key roles in plant growth and development. Additionally, the accumulation of RFOs is associated with stressful environmental conditions, especially desiccation tolerance [6,7,8,9,10]. In addition, RFOs also act as signaling molecules during pathogen attack and wounding and are used in carbohydrate catabolism to generate energy during germination [11,12,13,14,15].
Galactinol synthase (GolS; EC: 2.4.1.123), the key enzyme in the production of RFOs, is a small representative class of the eukaryotic glycosyltransferase family of enzymes (GTs; EC: 2.4.x.y) and involved in the biosynthesis of diverse sugars. As a key regulator of carbon-partitioning, it also catalyzes the first step in the RFO biosynthesis, i.e., from UDP-galactose (uridine diphosphate-galactose) and inositol to galactinol (α-d-galactosyl-1-l-myo-inositol) [13,16]. Hence, GolS is regarded as an important regulator of RFOs that imparts the osmoprotectant function to RFOs [4]. Previous researches revealed the positive relationship between GolS activity and plant abiotic stress tolerance. Expression of OsGolS increase in response to cold treatments and osmotic stress in rice (Oryza sativa) seedlings [17]. In Arabidopsis thaliana, AtGolS1, AtGolS2 and AtGolS3 were expressed in mature seeds and are also highly induced by abiotic stresses [4]. AtGolS1 and AtGolS2 are induced by drought and salt stress but not cold stress, while AtGolS3 is only induced by cold stress. In maize (Zea mays), ZmGolS1 has low expression in most tissues, while ZmGolS2 accumulates in germinating seeds subjected to dehydration stress and ZmGolS3 predominated in seeds prior to maturation desiccation [18]. In addition, galactinol and raffinose accumulation are related to plant cell protection against oxidative damage caused by many types of stressors and are positively correlated with the biosynthesis of abscisic acid (ABA) [19,20]. These studies all suggested that GolSs play crucial roles in plant abiotic stresses and hormone response.
So far, GolS genes have been identified and studied in several plant species at genome-wide level and many of GolS genes are up-regulated during abiotic and hormone treatments [4,12,14,15,16,17,18,19,20,21,22,23]. Previous evolutionary research suggested that GolS might have evolved from an ancestral fungal sequence (this gene has been lost in the process of evolution) and the role of GolS has being largely seed–plant–specific, the enzyme has been retained and evolved among angiosperms while not retained or differently evolved among other taxa [24]. In the phylogenetic tree of GolS proteins in flowering plants, the GolS in gymnosperm Pseudotsuga menziesii is the base and separation between monocot and dicot clades is distinct, suggesting the two clades might originated from a common ancestor. Though previous study observed that all GolS in monocots are clustered in a narrow clade whereas the diversification in dicots is greater [24], the evolutionary relationships of GolS proteins and their biological meanings in dicots, especially allopolyploid dicots remain unclear.
Rapeseed (Brassica napus) is one of the most important allopolyploid crops worldwide. It is formed by two progenitor species, Brassica rapa and Brassica oleracea [25,26]. Tobacco (Nicotiana tabacum) is an important allopolyploid plant for evolution and gene function characterization [27,28]. To unravel the evolutionary relationships of GolSs in allopolyploid dicots, we used the rapeseed and tobacco as model plants in this study and identified GolS family members in three Brassicaceae species (B. napus, B. rapa and B. oleracea) and three Solanaceae species (N. tabacum, Solanum lycopersicum and Solanum tuberosum). Then, we investigated the evolutionary relationships among GolS members in Brassicaceae and Solanaceae species and evaluated the predicted three-dimensional structures of representative GolS members from B. napus and N. tabacum. In addition, we determined the spatio-temporal and hormone response expression patterns of these GolS genes. Our study provides a comprehensive evolutionary analysis of GolS genes among the Brassicaceae and Solanaceae and a basis for subsequent studies examining the role of GolS in regulating plant tolerance.

2. Results

2.1. Identification of GolS Genes in Rapeseed and Tobacco

Using seven AtGolS proteins from A. thaliana as the query sequences, we identified 20 BnGolS genes and nine NtGolS genes (Table 1). All of the identified GolS proteins share a common Pfam (PF01501.18), which is a conserved domain of the glycosyl transferase 8 family. The nomenclature for BnGolS (BnGolS1-1 to BnGolS7-2) and NtGolS (NtGolS1-1 to NtGolS 2-6) genes was based on the corresponding AtGolS orthologues (Table 1). Each AtGolS corresponds to multiple GolS homologs in B. napus except AtGolS5, no B. napus homologs was identified, while only AtGolS1 and AtGolS2 orthologues were identified in N. tabacum (Table 1). The number of amino acid residues in the BnGolS proteins ranged from 121 (BnGolS7-1) to 342 (BnGolS1-1 and BnGolS1-4) and the relative molecular weights (MWs) ranged from 13.74 kDa (BnGolS7-1) to 39.26 kDa (BnGolS1-1). Theoretical pI (isoelectric point) values ranged from 4.69 (BnGolS4-1) to 7.03 (BnGolS3-2) and only that of BnGolS3-2 was higher than 7 in B. napus. The NtGolS proteins possessed 223 (NtGolS1-3) to 343 (NtGolS 1-1 and NtGolS1-2) amino acid residues, with relative molecular weights ranging from 25.26 kDa (NtGolS1-3) to 41.40 kDa (NtGolS2-2). Theoretical pI values ranged from 5.18 (NtGolS2-3) to 7.03 (NtGolS2-6) and only NtGolS2-6 exhibited a pI of >7. No transmembrane transport and signal peptide sequences were identified in the GolS proteins from rapeseed or tobacco (Table 1). Subcellular localization predictions for BnGolS and NtGolS suggest that only BnGolS3-2 localizes to mitochondria, while the other 28 GolS proteins all localize to the cytoplasm (Table 1).

2.2. Phylogenetic Analysis of BnGolS and NtGolS Proteins

To reveal the evolutionary relationships among GolS gene family members in B. napus, N. tabacum and A. thaliana, we performed multiple sequence alignment to 29 GolS protein sequences, using AmTrGolS (GolS in Amborella trichopoda) as an outgroup. Based on the lowest values of Akaike Information Criterion (AIC) and Bayesian Information Criterion (BIC) in Modelgenerator (Table S1), the was selected for construction of the maximum likelihood (ML, 100 bootstrap replicates) and Bayesian (BI) trees. According to previous phylogenetic analysis [29], GolS proteins were divided into four groups, including the GolS5 and GolS6 clade (Group 1), the GolS4 and GolS7 clade (Group 2), the GolS1 clade (Group 3) and GolS2 and GolS3 clade (Group 4) (Figure 1A). Group 1 and Group 2 contained 12 BnGolSs, while Group 3 and Group 4 possessed 8 BnGolSs and all NtGolSs. In Group 2, the evolutionary relationship between BnGolS and AtGolS was closer than that between NtGolS and AtGolS, which is consistent with the evolutionary relationship of these species. Since Arabidopsis and B. napus both belong to Brassicaceae, Rosidae and their split time might be 13 million years ago (MYA) [30]. Tobacco belongs to Solanaceae, Asterids, it’s ancestor had been diverged from ancestor of Rosidae about 83–123 MYA [31]. Phylogenetic analyses of 83 protein-coding and rRNA genes from the plastid genome for 86 species of seed plants [32] also proven the closer relationship between Arabidopsis and B. napus. All BnGolS and NtGolS proteins except for BnGolS7-1 possessed DxD and HxxGxxKPW motifs, the conserved motifs of GT members [33,34]. Both of the BnGolS6 and NtGolS2 subfamilies have six members, suggesting a notable expansion of these gene subfamilies compared with the other BnGolS and NtGolS subfamilies. Gene loss was also observed for GolS5 in B. napus and GolS3/4/5/6/7 in N. tabacum, since no orthologues of these proteins were found in the aforementioned species.
To further investigate the evolutionary relationships among GolS proteins in Brassicaceae and Solanaceae, a larger ML tree was constructed based on 70 GolS proteins from A. thaliana, B. napus, B. rapa, B. oleracea, N. tabacum, S. lycopersicum, S. tuberosum, O. sativa and Z. mays (AmTrGolS as an outgroup) (Figure 2). Considering that GolS formed a unique clade within the GT family [35,36], we also divided the 70 GolS members into four groups based on previous study [29] and gene copy number comparison (Figure 3). Group 4 (GolS2 and GolS3 clade) was the largest clade, containing 22 GolS members, including 10 GolSs from Brassicaceae and 12 GolSs from Solanaceae species, while group 3 was the smaller clade (GolS1), including 9 GolSs from Brassicaceae and 5 GolSs from Solanaceae species. Group 2 was the Brassicaceae-specific group, which contained 15 GolSs (GolS4 and GolS7 clade), the 15 GolSs only found in the Brassicaceae. Group 1 (GolS5 and GolS6 clade) contained 14 GolSs from Brassicaceae, 5 GolSs from monocots and no GolS from Solanaceae. We were unable to identify orthologues of AtGolS5 in the three Brassica species and orthologues of the AtGolS3/4/5/6/7 in the three Solanaceae species, indicating that gene loss of GolS is more pronounced in Solanaceae than in Brassica species. Each of the AtGolSs and its orthologues from the three Brassica species consistently formed a Brassicaceae cluster first, suggesting that GolS members in B. napus, B. oleracea and B. rapa, are highly homologous. Similarly, close evolutionary relationships were also found among the GolS families from N. tabacum, S. lycopersicum and S. tuberosum. Orthologues of AtGolS further clustered into a larger clade (Figure 2). For example, BnGolS1-1 and Bra021388 and BnGolS1-4 and Bol002593 formed two small Brassica clusters in terms of evolutionary time, then formed the larger Brassica GolS1 cluster with BnGolS1-2, Bra004474, BnGolS1-3 and Bol002519 and then the Brassicaceae GolS1 cluster formed with AtGolS1. Finally, the large GolS1 clade was formed by grouping the Brassicaceae and Solanaceae GolS1 branches. The relationship among GolS genes is in accordance with the evolutionary relationship among the Brassicaceae and Solanaceae families. In addition, the group 1 GolS members from monocots clustered in a clade with GolS5/6 members from Brassicaceae species, which are dicots. This clade represents the most ancient GolS members, which have been lost in the Solanaceae. Members of Brassicaceae in this clade (AtGolS5/6) clustered together and no orthologues of AtGolS5 were found in the three Brassica species. The GolS2/3 members from Brassicaceae clustered together to form a large clade with the GolS2 members from Solanaceae, suggesting that members of GolS3 may be either newly produced genes in Brassicaceae or may have been lost in Solanaceae (Figure 3). In Group 2, GolS4/7 are the most likely newly evolved Brassicaceae-specific genes, since GolS members in this clade were only identified in A. thaliana and three Brassica species.

2.3. Gene Structure and Conserved Motif Analyses

To confirm the exon/intron structures of AtGolS, BnGolS and NtGolS genes, we aligned the cDNAs with their corresponding genomic sequences. The gene structures among BnGolS genes are highly conserved and have high similarity with AtGolSs. However, the structures of NtGolS genes were noticeably different from each other and from the homologous AtGolS genes (Figure 1B). Interestingly, most of the NtGolS genes have one or two introns that are longer than their exons but the reverse trend was found in most of the BnGolS genes, including BnGolS2-1, BnGolS2-2 and BnGolS7-2. All of the AtGolSs genes and most of the BnGolS genes are composed of four exons and three introns, with a few exceptions: BnGolS7-1 only has one exon and no introns and BnGolS3-1 and BnGolS3-2 each have five exons and four introns (Figure 1B). The exons and introns of duplicated NtGolS genes varied substantially.
To better understand the structural diversity of BnGolS and NtGolS proteins, online MEME/MAST tools were used to analyze the conserved motifs [37]. Fifteen putative protein motifs were predicted using the MEME program and uploaded into MAST for motif detection. We found that motif 1 was observed in all 36 GolS proteins, motif 2 was observed in all GolS proteins except BnGolS7-1 and motif 3 was observed in most GolS proteins with the exception of BnGolS7-1 and NtGolS1-3 (Figure 1C). Motifs 4, 5, 6 and 11 were also observed in most GolS proteins. Motif 7 was detected in 30 GolS proteins and displays a specific peptide sequence –APSAA– in the C-terminus but was not found in AtGolS7, BnGolS4-2, BnGolS4-4, BnGolS7-1, BnGolS7-2 and NtGolS2-2. Motifs 8 and 12 were only detected in the N-terminus of proteins from the BnGolS6 clade, while proteins from other clades have motif 10 (Figure 1C). Motif 9 was found in the BnGolS1, BnGolS2 and BnGolS3 branches and the remaining motifs were identified in individual proteins. BnGolS1-1 and BnGolS1-4 had the most motifs, each containing 11 motifs, while BnGolS7-1 had the fewest, containing only 2 (Figure 1C). According to the InterProScan annotation, we found that motifs 1–4 are conserved and important to maintain binding activity of GolS proteins, especially the motif 2 (HxxGxxKPW) and motif 3 (DxD).

2.4. Chromosome Location and Duplication Analysis

In the Brassica database, we determined that the positions of 20 BnGolS genes are located on 11 chromosomes in B. napus. No tandem duplication events occurred in BnGolS gene family. No BnGolS genes were observed on chromosomes A02, A03, A06, A07, A10, C02, C03, C05, C06, or C07 and only one BnGolS gene was identified on each of the A01, A04, A05, A08 and C01 chromosomes (Figure 4A). Two BnGolS genes were found on chromosomes C04 and C09, four BnGolS genes were found on chromosome C08 and five BnGolS genes were found on chromosome A09. Among the four members of the BnGolS4 subfamily, BnGolS4-1 and BnGolS4-3 and BnGolS4-2 and BnGolS4-4 were found to be closely linked on their respective chromosomes (Figure 4A), which might be the result of a tandem duplication event during evolution. Unlike the BnGolS genes, only four NtGolS genes were identified on known chromosomes, while the remaining five NtGolS genes were found on a scaffold which has not yet been assembled into a chromosome (Table 1). Unlike N. tabacum, B. napus experienced an additional whole genome triplication (WGT) event in its evolutional history. By comparing the synteny of GolS genes in A. thaliana, B. napus, B. oleracea and B. rapa, we found that some genes were preserved and some were either duplicated or lost (Figure 4B). We determined that AtGolS1 has two homologous syntenic gene pairs in B. oleracea, B. rapa and four in B. napus, two of which are located on the A genome and two on the C genome. AtGolS7 also has two homologous syntenic genes in B. oleracea and B. rapa but no syntenic homologous genes in B. napus. AtGolS2 and AtGolS3 only possesses one homologous syntenic gene each in both B. oleracea and B. rapa, that were also syntenic with homologous genes in the B. napus genomes A and C, respectively. We were unable to identify any syntenic homologous genes for the AtGolS4 genein Brassica species (Figure 4B).
We discovered that most of the syntenic GolS genes are clustered on chromosome 1 in A. thaliana, however, in Brassica most of the syntenic GolS genes are located on different chromosomes. AtGolSs have good syntenic relationships with Brassica GolS genes due to either genome rearrangement or gene loss after a genome triplication event. This indicates that the expansion pattern of the GolS gene family might be a consequence of mesopolyploidy in Brassica evolution [38]. We discovered 5 syntenic genes in A. thaliana, corresponding to 7 genes in B. oleracea and B. rapa and 8 genes in B. napus. These results suggest that some GolS members were lost after the genome triplication event in Brassica species and some new GolS members were also presented throughout its evolutionary history, which might be necessary for Brassica crops to cope with abiotic stresses that occur due to a change in the environmental conditions. To explore the selective pressures on these GolS genes, we determined the non-synonymous/synonymous mutation ratio (Ka/Ks) for BnGolS and NtGolS genes. Previous report showed that the patterns of codon usage bias in dicot plants are conservative [39], suggesting the Ka/Ks results in our study could reflect selection pressure on the evolution of three plant species. The ratios for all BnGolS and NtGolS genes were less than 1, indicating that all BnGolS and NtGolS genes were subject to purifying selection (Table 2) [40]. In B. napus, BnGolS1, BnGolS2, BnGolS4 and BnGolS7 in the A genome displayed a higher Ka/Ks ratio than their corresponding homologs in the C genome. However, the opposite was observed for BnGolS3 and BnGolS6, suggesting that different evolutionary pressures were experienced on homologous genes after the whole genome triplication (WGT) in B. napus. The Ka/Ks ratio for BnGolS1-1 and BnGolS1-4 was significantly higher than that of other BnGolS family members, especially compared with BnGolS1-2 and BnGolS1-3 (Table 2), implying that BnGolS1-1 and BnGolS1-4 evolved faster than did the other BnGolS members and therefore, they are likely to have different biological functions. All of the NtGolS genes from N. tabacum had low Ka/Ks ratios, the highest of which (NtGolS2-6) was still lower than the lowest ratio of the BnGolS genes (BnGolS2-6) (Table 2). The Ka/Ks ratio of BnGolS was larger than that of NtGolS, indicating that the evolutionary pressures on BnGolS genes were greater than those on NtGolS genes.

2.5. Structure Prediction and Homology Modeling

To generate the 3D structures for BnGolS and NtGolS, structures for two representative GolS proteins (BnGolS1-2 and NtGolS1-1) were predicted using I-TASSER [41]. The best template used for the 3D structure predication was the rabbit muscle glycogenin [42]. The modeled structure for BnGolS1-2 has 11 α-helices and 6 β-strands (Figure 5A), while that of NtGolS1-1 has 12 α-helices and 7 β-strands (Figure 5B). The DxD motif is located between β3 and β4 in BnGolS1-2 (Figure 5A) and between β4 and β5 in NtGolS1-1(Figure 5B). Template-structure alignment resulted in a better structure match but the two GolS1 sequences had low levels of similarity with the template sequence. In the threading aligned region, the sequence identity between BnGolS1-2 or NtGolS1-1 and the 1ll0B template is 0.27 in the threading region and 0.23 across the whole template, in both cases. The coverage of threading alignment was 0.75 for BnGolS1-2 and 0.76 for NtGolS1-1. The normalized Z-scores of the threading alignments were both greater than 1 (1.76), meaning a good alignment was obtained for both BnGolS1-2 and NtGolS1-1 and the two predicted protein structures displayed the same overall folds. Compared with the template model, the unaligned sequences that were used to construct the models were located at the N-terminal start, C-terminal tail and a long middle sequence located on the periphery of the protein structure core.

2.6. Binding Pocket Prediction

Considering it is very difficult to predict the function of conserved motifs without a crystal structure or explicitly defined structural motifs, we applied a method that expounds the function of predicted conserved sequences by mapping conserved residues onto a sufficiently close known crystal structure [24]. Since the modeled 3D structures of the two GolS1 proteins are extremely similar (Figure 5A,B), only BnGolS1-2 was analyzed further. The important substrate-binding residues of BnGolS1-2 were predicted using the crystal structures 1zdgA and 3rmwA as templates [43]. According to this model, Asp125, Asp127, His263 and Cys265 are involved in Mn2+ binding, while Phe33, Leu34, Gly36, Asn37, Tyr40, Ile105, Asp125, Gly126, Asp127, Asn191, Ala189, Gly193, Ala220, Glu221, Gln222, His263, Cys265, Ala266 and Lys270 are involved in UDP-sugar binding. GolS proteins are known to catalyze the synthesis of galactinol from UDP-galactose and myo-inositol but no inositol-binding site was identified using I-TASSER. To support the UDP-galactose binding site and to predict the binding site for inositol, low-resolution protein structures were analyzed by the blind molecular docking method BSP-SLIM [44]. We found that UDP-galactose and inositol are both positioned near the conserved DxD motif (Figure 6A,B), which is buried deep inside the binding pocket (Figure 6C,D). Furthermore, two Asp residues play a key role in the binding of the ligands in UDP-galactose and inositol. We found that the crucial inositol-binding residues are Phe33, Try40, Ile105, Asp125 and Asp127 (Figure 6A), while the important UDP-galactose binding residues are Phe33, Try40, Ile105, Lys109, Try123, Asp125, Asp127, Try163, Ile165, Gln222 and Lys270 (Figure 6B). The Lys270 residue involved in UDP-galactose binding is part of the HxxGxxKPW motif. These results suggest that the conserved DxD motif plays a crucial role in the binding of micro-molecules in the galactinol catalytic pocket, whereas the HxxGxxKPW motif generally participates in macro-molecular binding. Interestingly, we found that the core area is surrounded by abundant loose coil structures, which maintain the stability of the protein and provides a stable environment for the synthesis of galactinol by preventing the hydrolysis of galactinol during the synthesis process.

2.7. Tissue-Specific and Hormone-Induced Expression Patterns of BnGolS and NtGolS

To begin the investigation into the biological functions of BnGolS and NtGolS enzymes, we assessed the expression profiles of the corresponding GolS genes in various tissues during multiple developmental stages. In B. napus, the BnGolS families displayed divergent expression patterns. Four BnGolS1 members were primarily expressed in hypocotyls, while BnGolS1-1 and BnGolS1-3 were highly expressed in seeds at 40 days after flowering (Figure 7A). BnGolS2 members were strongly expressed in seeds and silique pericarps, especially at 40 days after flowering (Figure 7A) and high expression of BnGolS3 members was observed in mature leaves and silique pericarps at 40 days after flowering. One BnGolS4 member (BnGolS4-1) displayed abundant expression in roots. BnGolS6 and BnGolS7 members were not expressed in most tissues and were only weakly expressed in specific tissues during a specific development stage (Figure 7A). For example, BnGolS6-4 was weakly expressed in seeds at 10 days after flowering and BnGolS7-2 was weakly expressed in seeds at 30 and 40 days after flowering. In N. tabacum, NtGolS2-2, NtGolS2-3 and all of the NtGolS1 members displayed constitutive expression in all tissues tested (Figure 7B). NtGolS1-1 and NtGolS1-2 were highly expressed in stems and dry capsules, while NtGolS1-3 was highly expressed in mature flowers and senescent flowers. NtGolS2-2, NtGolS2-3, NtGolS2-4 and NtGolS2-6 displayed high expression in dry capsules, while NtGolS2-1 and NtGolS2-5 had no detectable expression (Figure 7B). These results show that GolS orthologues between B. napus and N. tabacum have different tissue-specific expression patterns and that the expression profiles of the duplicated GolS genes in a small GolS gene family diverged.
To investigate the roles of the GolS genes in response to hormone treatments, rapeseed seedlings were subjected to brassinosteroid (BR), α-naphthaleneacetic acid (NAA), salicylic acid (SA), 6-benzyladenine (6BA), abscisic acid (ABA)and methyl jasmonate (MeJA) treatments. Six hormone inducible marker genes all showed up-regulation in response to corresponding hormone treatments (Figure S1), suggesting the physiological status of seedling plants were suitable for detecting the hormone inducible expression patterns to GolS genes. Although BnGolS2s, BnGolS3s and BnGolS1-2 were strongly induced by more than one hormones, some of the remaining genes were specifically activated by a single hormone (Figure 7C). Moreover, BnGolS1-2 was significantly up-regulated in response to NAA and SA treatments, particularly by 4 and 12 h of treatment. Two BnGolS2 members were up-regulated in response to BR, NAA, SA and ABA treatments and were repressed in response to 6BA and JA treatments but the inducible expression was at one or two stages of hormone treatments in this study regardless of whether up-regulated or repressed. BnGolS3-1 expression was reduced in response to NAA, SA, 6BA and JA treatments, while BnGolS3-2 expression was reduced in response to 6BA and JA treatments (Figure 7C). Although BnGolS6-5 expression was extremely low in all tissues examined, its expression was up-regulated in response to NAA and SA treatments. In addition, the remaining GolS genes were slightly induced or repressed in response to other hormone treatments but the degree of expression variation was not as dramatic as it was for the abovementioned genes (Figure 7C).

3. Discussion

3.1. Structural Characteristics of the BnGolS and NtGolS Gene Family

As representative members of the glycosyltransferase family, GolS genes have been lost from plant genomes over evolutionary time to varying degrees [24]. In this study, we identified 20 BnGolS and 9 NtGolS genes in B. napus and N. tabacum, respectively. We found that most of the GolS family members have three or few introns and most homologous genes have a similar gene structure, implied the similar function of homologous genes. In addition, most of the BnGolS and NtGolS proteins have a pI of <7, which may explain why GolS is enzymatically active in acidic conditions. GolS proteins localized to the cytoplasm, which suggests a potential osmoprotectant function for GolS in this study [24]. The conserved structural pocket in most of GolS proteins contains two conserved GT8 motifs, DxD and HxxGxxKPW, which are located very close to each other (Figure 5A,B). This suggests that these two motifs may function coordinately during the enzymatic reaction. The DxD motif is a conserved motif in diverse GT families [33] and is thought to interact with Mn2+ and mediate binding to NDP-sugar (Nucleotide diphosphate-sugar) donors [42,45]. The HxxGxxKPW motif was identified by Arabidopsis GAUT and GATL proteins, suggested that it would be part of the catalytic site [34]. The H residue in the HxxGxxKPW motif is thought to bind to Mn2+ and the G and K residues interact with the NDP-sugar donor [42]. The pentapeptide in the disordered loops at the C-terminus (APSAA) was found in most of the BnGolS and NtGolS proteins, with the exception of BnGolS4-2, BnGolS4-4, BnGolS7-1, BnGolS7-2 and NtGolS2-2. More than 30 amino acids in the C-terminus of these genes were missing when compared with other members. This illustrates that sequence loss has occurred in some members of the GolS family to varying degrees. Although no evidence was found in the literature that supports a function for the APSAA sequence in binding to UDP-galactose or myo-inositol, upon nucleotide sugar binding, some reorganized GT enzymes would adopt a new conformation within this disordered loop region to create a pocket for the acceptor molecule to bind [46,47]. We speculate that the C-terminus of the GolS proteins provides a stable environment for substrate binding based on the 3D structure and surface coverage of GolS proteins.

3.2. Phylogenetic and Evolutionary Relationship of GolS Proteins

GolS evolved to meet the RFOs biosynthesis requirements during abiotic stress [4,10,11,15,16], especially in higher plants [24]. In this study, GolS members in B. napus, N. tabacum and Arabidopsis were divided into four groups. Average expression levels of GolS genes in Group 2–4 from B. napus were significant higher than those in Group 1 (Figure 1A and Figure 7A, p value = 1.98 × 10−5, Student’s t-test), suggesting that Group 2–4 members might be more important than Group 1 for galactinol synthesis. BnGolS4-2/4 was more closely related to AtGolS7 than AtGolS4, while BnGolS4-1/3 was more closely related to AtGolS4 than to AtGolS7, indicating that the biological function of BnGolS4-2/4 might be diverged after WGT, which could be supported by the results of Ka/Ks values of BnGolS4-2/AtGolS4 and BnGolS4-4/AtGolS4 (Table 2). Generally, each A. thaliana gene has three syntenic copies in B. rapa and B. oleracea [48,49] and six syntenic copies in B. napus, since it was generated by the progenitor species of B. rapa and B. oleracea 7500–12,500 years ago [26]. However, 1.5–2 copies of each GolS gene were identified in B. rapa and B. oleracea and 2–6 copies were identified in B. napus [26,48,50] because of genome shrinkage or gene loss after WGT. In this study, 8 syntenic copies (from 20 copies) of GolS genes in B. napus and 7 syntenic copies (from 10–11 copies) of GolS genes were found in ancient chromosomes of B. rapa and B. oleracea ancestors (Figure 3B), suggesting that genome shrinkage or gene loss occurred in the GolS family of the three Brassica crops. Furthermore, no orthologues of AtGolS5 were found in the three Brassica species, indicating that AtGolS5 may have arisen via duplication of the gene encoding AtGolS6 specifically in A. thaliana. Monocots have fewer GolS copies than do Solanaceae and Brassicaceae [17,18] and Solanaceae have fewer copies than Brassicaceae. Considering the evolutionary history of plant species (Figure 3), we speculate that only one ancient copy of GolS was present in the last common ancestor of both monocots and dicots. In monocots, only one GolS copy remained, while three ancient dicot-specific GolS copies evolved from the pan-eudicot palaeohexaploidy event that occurred early in eudicot evolution. During evolution, an ancient GolS copy (corresponding to AtGolS5/6) was lost in the Solanaceae and an additional GolS copy evolved through gene duplication in Brassicaceae after the split between the Solanaceae and Brassicaceae families, resulting in 4 ancestral GolS copies in the Brassicaceae. Sixty-two percent of duplicated genes in Arabidopsis show differences in their expression patterns, reflecting divergence in their biological functions [51]. A similar phenomenon was observed among the duplicated members of the GolS family, which may have caused the divergence of GolS functions in response to different abiotic stress conditions.

3.3. Expression Profile of GolS Genes

Several studies report that galactinol primarily accumulates during seed maturation and in leaves [4,52,53]. Accumulation of BnGolS-1 (named BnGolS1-1 in this study) in seeds during maturation is crucial for acquiring desiccation tolerance or dry weight deposition [54]. However, in this study, BnGolS1-1sightly expressed in developing seeds, its homologous genes (BnGolS1-2, BnGolS1-3, BnGolS2-1 and BnGolS2-2) strongly expressed in seed maturation to meet the acquisition. It has been reported that the important function of RFOs in phloem export and carbon storage were preserved by highly expression of GolS in leaves, which is crucial for plant protection against cold stress condition and oxidative damage, which is also crucial for plant protection against cold stress and oxidative damage [55]. Additionally, the reduction of GolS expression in leaves leads to an improvement in drought tolerance [4]. In this study, BnGolS3 subfamily members were highly expressed in leaves, while no highly expressed GolS genes were observed in tobacco leaves but NtGolS1-3 was highly expressed in mature and senescent flowers. This illustrates that GolS genes in rapeseed are more important to resistant to abiotic stress in leaves than GolS genes in tobacco. Strong expression of GolS genes in seeds was previously associated with desiccation tolerance in developing seeds [6,53]. In this study, BnGolS1 members were found to be highly expressed in the cotyledon during germination, suggesting that these genes might be crucial for seed germination. The expression patterns we detected also imply that BnGolS2 and BnGolS3 members might be involved in silique pericarp development. Most NtGolS members were highly expressed in dry capsules, suggesting that they play a role in seed maturation. BnGolS and NtGolS were highly expressed in seeds, silique pericarps and dry capsules [56], which is similar to ZmGolS genes that are predominantly expressed in seeds [18].
It has been reported that the expression of GolS genes increased in response to various abiotic stresses in plants [4,17,18,57]. Expression levels of GolS genes were up-regulated by stimulation with plant hormones [4,58]. Among the four BnGolS1 members, only BnGolS1-2 shows significant up-regulation in response to NAA and SA treatments, while other BnGolS1 members were insensitive to these treatments. By contrast, BnGolS3-1 shows significant down-regulation in response to NAA treatment. NAA promotes plant growth and prevents leaf abscission and that SA induces the production of certain proteins. Therefore, we propose that the synthesis of galactinol or raffinose might be induced by NAA and SA. Previous studies showed that OsGolS and AtGolS3 are insensitive to ABA treatments [17] and that the expression of AtGolS1 and AtGolS2 are weakly induced by ABA treatment [4]. Transpiration in leaves is reduced by ABA treatments through stomatal closure, which is important for the plant responses to abiotic stress and pathogen attack. Our results showed that most of the BnGolS genes were slightly induced by ABA treatments (Figure 7C), indicating that these genes might be associated with abiotic stress and ABA-mediated signaling. These highly expressed genes were repressed upon treatments with MeJA and 6-BA. Since JA not only has a similar effect as ABA and 6-BA but also affects the transport of carbohydrates and other organic compounds, this may suppress the biosynthesis and transport of galactinol or raffinose. The expression of BnGolS2 members was more sensitive to hormone treatments, especially to short-term hormone induction, which suggests that plants can increase ROF biosynthesis to enhance stress tolerance by rapidly up-regulating the expression of BnGolS2 members in response to short-term stress treatments.
In summary, we found that GolS genes have different expression patterns that are induced by various plant hormones and that several of these genes have important roles in plant environmental adaptation.

4. Materials and Methods

4.1. Identification of GolS Genes

Genomic, coding and protein sequences from A. thaliana were downloaded from the Arabidopsis Information Resource (TAIR, http://www.arabidopsis.org), those from B. napus, B. rapa and B. oleracea were retrieved from the Brassica Database (BRAD, http://brassicadb.org/brad), those from N. tabacum, S. lycopersicum and S. tuberosum were downloaded from the Solanaceae Database (https://solgenomics.net) and those from maize, rice and Amborella trichopoda were obtained from Phytozome 12.0 (www.phytozome.net).
To identify GolS genes in the above-mentioned species, seven GolS proteins from Arabidopsis were used as the query sequences in a reciprocal Basic Local Alignment Search Tool Protein (BLASTP) analysis [59,60] at a threshold E-value of 1e-5 and a minimum alignment coverage of 50%. Then, all the GolS protein sequences were analyzed with Pfam scan (http://www.ebi.ac.uk/Tools/pfa/pfamscan/) to confirm the presence of a GolS-related domain (PF01501.18, Glyco_transf_8).

4.2. Protein Sequence Analysis

The molecular weight (MW, kDa) and theoretical isoelectric point (pI) were calculated using the ProtParam tool in ExPASy (http://web.expasy.org/protparam/, [61]). The transmembrane transport peptides were predicted by the Tmpred tool in ExPASy (http://www.ch.embnet.org/software/TMPRED_form.html/, [62]) using default parameters and signal peptides were predicted by the SignalP 4.1 (http://www.cbs.dtu.dk/services/SignalP/, [63]) server using default parameters. The subcellular localization of each protein was predicted with the MultiLoc2.0 server (http://abi.inf.uni-tuebingen.de/Services/MultiLoc2/, [64]) using the predictive method: MultiLoc2-HighRes (Plant), 10 Localization.

4.3. Chromosomal Locations, Gene Structure and Protein Motif Identification Analysis

The chromosomal locations of BnGolS genes and the duplicated relationship between A. thaliana and Brassica species were obtained from the Brassica Database and those of NtGolS genes were obtained from the Solanaceae Database (https://solgenomics.net/organism/Nicotiana_tabacum/genome/, [28]). The Gene Structure Display Server (GSDS 2.0: http://gsds.cbi.pku.edu.cn/, [65]) was used to determine the exon/intron structures of BnGolS and NtGolS. The Multiple Expectation Maximization for Motif Elicitation program (MEME 4.12.0, http://alternate.meme-suite.org, [37]) was used to identify the conserved motif of GolS proteins in B. napus and N. tabacum with the following parameter settings: the minimal and maximal motif width were set to 6 and 100 amino acids, respectively and the number of motifs was 15. Only motifs with an e-value of <1 × 10−10 were kept for further motif analysis (other parameters were set to default). Then the MAST 4.12.0 was used to search for the detected motifs in the protein database [66].

4.4. Phylogenetic and Evolution Analysis

Multiple sequence alignment was conducted using MUSCLE (http://www.ebi.ac.uk/Tools/msa/muscle/, [67]). The ML tree was constructed using the online PhyML server (PhyML 3.0, http://www.atgc-montpellier.fr/phyml/, [68,69]) with the JTT + G substitution model. Model of rate heterogeneity: Discrete Gamma; Number of rate categories: 4; Gamma distribution parameter α: 0.43) and 100 bootstrap replicates and was used to infer evolutionary relationships among the sequences. The model used for ML and Bayesian trees were calculated using ModelGenerator v0.85 (http://mcinerneylab.com/software/modelgenerator/). There are 56 nucleotides and 96 amino acid substitution models in the Model generator. The optimal model will be selected based on the lowest values of AIC and BIC [70]. Then, CDS sequence alignments were used to calculate the synonymous mutation rate (Ks) and non-synonymous mutation rate (Ka) and evolutionary constraint (Ka/Ks) for collinear gene pairs using the KaKs_calculator 2.0 with the GY-HKY method [71].
To validate the two ML trees, the corresponding neighbor joining (NJ) and BI trees were constructed using the same alignment of GolS proteins that was prepared with MEGA7 [72] (Figures S2A and S3A) and a new alignment prepared with MrBayes 3.2.6 (Figures S2B and S3B, http://mrbayes.sourceforge.net), respectively. A bootstrap analysis was conducted with 1000 replicates for the NJ tree construction. To generate the BI tree, the JTT + G substitution model was selected and run for 5,000,000 generations, sampling every 1000 generations with default parameters. By comparing the topologies of the NJ, BI and ML trees, the ML tree with best topology was chosen for further evolutionary analyses (Table S2).

4.5. Three-Dimensional Structure Prediction of GolS Proteins

The three-dimensional (3D) structures were predicted with the online I-TASSER program (http://zhanglab.ccmb.med.umich.edu/I-TASSER/, [41]). To identify structurally similar templates in the Protein Data Bank (PDB), the query sequence was first subjected to multiple rounds of threading using LOMETS [73], which is a meta-threading server with nine locally installed threading programs. Rabbit muscle glycogenin (PDB ID: 1ll0) [42] was the top structural template for our queried proteins. Then, the generated 3D model was aligned to the templates using TM-align and BSLIM was used to dock ligands to the protein structure [74,75]. The constructed model was examined and visualized by Chimera 1.2 (https://www.cgl.ucsf.edu/chimera/).

4.6. Plant Materials and Treatment

Seeds from rapeseed cultivar “ZS11” and tobacco cultivar “K326” were germinated in plant growth chambers (a photoperiod of 16 h at 25/18 °C day/night, 60% humidity, 250 µmoles/m2/s, PGC Flex; Conviron, Winnipeg, MB, Canada) and then transplanted into the field in Chongqing, China. To determine the expression patterns of GolS family members in different tissues and organs of B. napus, hypocotyls (Hy) and cotyledons (Co) were harvested at 72 h after seed germination. Roots (Ro), stems (St), mature leaves (Le) and buds (Bu) from B. napus were collected at the initial blooming stage, while seeds (Se) and silique pericarps (SP) were collected at 10, 21, 30 and 40 days after flowering. In tobacco, roots (Ro), stems (St), young leaves (LeY), mature leaves (LeM), senescent leaves (LeS), young flowers (YF), mature flowers (MF), senescent flowers (SF) and dry capsules (DC) were sampled from tobacco plants at 75 days after transplanting (DAT).
To determine the expression patterns of BnGolS genes in leaves in response to hormone stresses, 4-week-old B. napus seedlings were grown in a plant growth chamber under the conditions introduced above [76]. For hormone treatments, the leaves of 4-week-old seedlings were treated with 10 μM BR, 50 μM NAA, 1 μM SA, 100 μM MeJA, or 50 μM ABA (GenTel, Beijing, China), prepared in ethanol, or with 50 μM 6-BA, which was dissolved in 0.1 M NaOH using 0.1% Triton X-100. Leaves from both treated and control plants were sampled at 0.5, 4, 12 and 24 h after treatments. Leaves from plants treated with 0.1% Triton X-100 in ethanol or NaOH were used as controls. For all samples used for qRT-PCR, three biological replicates were collected, each comprising the second youngest leaves of three independent plants. All samples were immediately frozen in liquid nitrogen and stored at −80 °C until RNA isolation.

4.7. RNA Isolation and Quantitative Real-Time PCR (qRT-PCR) Analysis

Total RNA was extracted from all samples using an RNAprep Pure Plant Kit (Tiangen, Beijing, China) and cDNA was synthesized from 1 μg of total RNA using a PrimeScript RT Master Mix Kit (TaKaRa, Dalian, China) according to the manufacturer’s instructions. Primers for the 16 BnGolS genes and 6 external control genes in B. napus used in qRT-PCR detection were obtained from the qPrimerDB database (real-time quantitative PCR Primer Database, http://biodb.swu.edu.cn/qprimerdb) [77] and primers for the nine NtGolS genes were designed using Geneious 10 (Biomatters, Auckland, New Zealand) (Table S3). To monitor the physiological status of plants under different hormone treatments, six hormone inducible marker genes were used as the external controls. All of the qRT-PCR reactions were implemented as described in the MIQE guidelines [78]. Two independent biological replicates, each with three technical replicates, were implemented for each sample. Relative expression levels were calculated using the 2−ΔΔCt method, with BnACT7 and NtEF-1α as internal controls [79].

Supplementary Materials

Supplementary materials can be found at www.mdpi.com/1422-0067/18/12/2768/s1.

Acknowledgments

This work was funded by the Program of China National Tobacco Corporation (110201603009), National Natural Science Foundation of China (31571701), Key Special Program of China National Tobacco Corporation (TS-02-20110014), National Key Research and Development Plan (2016YFD0101007), National High Technology Research and Development Programs of China (2013AA102602), and Programme of Introducing Talents of Discipline to Universities (B12006). The authors are very grateful to Kathleen Farquharson, the Science Editor for The Plant Cell, for her critical reading of the manuscript.

Author Contributions

Bo Lei and Kun Lu conceived and designed the experiments; Yonghai Fan, Mengna Yu, Miao Liu, Rui Zhang Wei Sun and Mingchao Qian performed the sampling and experiments; Yonghai Fan, Huichun Duan, Wei Chang, Jinqi Ma, Cunmin Qu and Kai Zhang analyzed the data; contributed to data analysis and interpretation; Yonghai Fan and Kun Lu wrote the paper. All authors reviewed the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, W.; Vinocur, B.; Altman, A. Plant responses to drought, salinity and extreme temperatures: Towards genetic engineering for stress tolerance. Planta 2003, 218, 1–14. [Google Scholar] [CrossRef] [PubMed]
  2. Theocharis, A.; Clément, C.; Barka, E.A. Physiological and molecular changes in plants grown at low temperatures. Planta 2012, 235, 1091–1105. [Google Scholar] [CrossRef] [PubMed]
  3. Fang, Y.; Xiong, L. General mechanisms of drought response and their application in drought resistance improvement in plants. Cell. Mol. Life Sci. 2015, 72, 673–689. [Google Scholar] [CrossRef] [PubMed]
  4. Taji, T.; Ohsumi, C.; Iuchi, S.; Seki, M.; Kasuga, M.; Kobayashi, M.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Important roles of drought- and cold-inducible genes for galactinol synthase in stress tolerance in Arabidopsis thaliana. Plant J. 2002, 29, 417–426. [Google Scholar] [CrossRef] [PubMed]
  5. Levitt, J. Response of plants to environmental stresses. In Water, Radiation, Salt and Other Stresses; Academic Press: Cambridge, MA, USA, 1980; Volume 2. [Google Scholar]
  6. Castillo, E.M.; de Lumen, B.O.; Reyes, P.S.; de Lumen, H.Z. Raffnose synthase and galactinol synthase in developing seeds and leaves of legumes. Agric. Food Chem. 1990, 38, 351–355. [Google Scholar] [CrossRef]
  7. Chen, Y.; Burris, J.S. Role of carbohydrates in desiccation tolerance and membrane behavior in maturing maize seed. Crop Sci. 1990, 30, 971–975. [Google Scholar] [CrossRef]
  8. Koster, K.L. Glass formation and desiccation tolerance in seeds. Plant Physiol. 1991, 96, 302–304. [Google Scholar] [CrossRef] [PubMed]
  9. Leprince, O.; Hendry, G.A.F.; McKersie, B.D. The mechanisms of desiccation tolerance in developing seeds. Seed Sci. Res. 1993, 3, 231–246. [Google Scholar] [CrossRef]
  10. Peterbauer, T.; Lahuta, L.B.; Blochl, A.; Mucha, J.; Jones, D.A.; Hedley, C.L.; Gorecki, R.J.; Ricgter, A. Analysis of the raffinose family oligosaccharide pathway in pea seeds with contrasting carbohydrate composition. Plant Physiol. 2001, 127, 1764–1772. [Google Scholar] [CrossRef] [PubMed]
  11. Buckeridge, M.S.; Dietrich, S.M.C. Mobilisation of the raffinose family oligosaccharides and galactomannan in germinating seeds of Sesbania marginalta Benth (Leguminosae-Faboideae). Plant Sci. 1996, 117, 33–43. [Google Scholar] [CrossRef]
  12. Downie, B.; Gurusinghe, S.; Dahal, P.; Thacker, R.R.; Snyder, J.C.; Nonogaki, H.; Yim, K.; Fukanaga, K.; Alvarado, V.; Bradford, K.J. Expression of a galactinol synthase gene in tomato seeds is up-regulated before maturation desiccation and again after imbibition whenever radicle protrusion is prevented. Plant Physiol. 2003, 131, 1347–1359. [Google Scholar] [CrossRef] [PubMed]
  13. Peterbauer, T.; Richter, A. Biochemistry and physiology of raffinose family oligosaccharides and galactosyl cyclitols in seeds. Seed Sci. Res. 2001, 11, 185–197. [Google Scholar]
  14. Lahuta, L.B.; Goszczynska, J. Inhibition of raffinose family oligosaccharides and galactosyl pinitols breakdown delays germination of winter vetch (Vicia villosa Roth) seeds. Acta Soc. Bot. Pol. 2009, 78, 203–208. [Google Scholar] [CrossRef]
  15. Cho, S.M.; Kang, E.Y.; Kim, M.S.; Yoo, S.J.; Im, Y.J.; Kim, Y.C.; Yang, K.Y.; Kim, K.Y.; Kim, K.S.; Choi, Y.S.; et al. Jasmonate-dependent expression of a galactinol synthase gene is involved in priming of systemic fungal resistance in Arabidopsis thaliana. Botany 2010, 88, 452–461. [Google Scholar] [CrossRef]
  16. Liu, J.J.; Odegard, W.; de Lumen, B.O. Galactinol synthase from kidney bean cotyledon and zucchini leaf. Plant Physiol. 1995, 109, 505–511. [Google Scholar] [CrossRef] [PubMed]
  17. Takahashi, R.; Joshee, N.; Kitagawa, Y. Induction of chilling resistance by water stress, and cDNA sequence analysis and expression of water stress regulated genes in rice. Plant Mol. Biol. 1994, 26, 339–352. [Google Scholar] [CrossRef] [PubMed]
  18. Zhao, T.Y.; Thacker, R.; Corum, J.W., III; Snyder, J.C.; Meeley, R.B.; Obendorfd, R.L.; Downiea, B. Expression of the maize galactinol synthase gene family: (I) Expression of two different genes during seed development and germination. Physiol. Plant. 2004, 121, 634–646. [Google Scholar] [CrossRef]
  19. Lahuta, L.B.; Górecki, R.J.; Gojlo, E.; Horbowicz, M. Effect of exogenous abscisic acid on accumulation of raffinose family oligosaccharides and galactosyl cyclitols in tiny vetch seeds (Vicia hirsuta [L.] S.F. Gray). Acta Soc. Bot. Pol. 2004, 73, 277–283. [Google Scholar] [CrossRef]
  20. Nishizawa, A.; Yabuta, Y.; Yoshida, E.; Maruta, T.; Yoshimura, K.; Shigeoka, S. Arabidopsis heat shock transcription factor A2 as a key regulator in response to several types of environmental stress. Plant J. 2006, 48, 535–547. [Google Scholar] [CrossRef] [PubMed]
  21. Wang, Z.; Zhu, Y.; Wang, L.; Liu, X.; Liu, Y.; Phillips, J.; Deng, X. A wrky transcription factor participates in dehydration tolerance in Boea hygrometrica by binding to the W-box elements of the galactinol synthase (BhGolS1) promoter. Planta 2009, 230, 1155–1166. [Google Scholar] [CrossRef] [PubMed]
  22. Pillet, J.; Egert, A.; Pieri, P.; Lecourieux, F.; Kappel, C.; Charon, J.; Gomes, E.; Keller, F.; Delrot, S.; Lecourieux, D. VvGOLS1 and VvHsfA2 are involved in the heat stress responses in grapevine berries. Plant Cell Physiol. 2012, 53, 1776–1792. [Google Scholar] [CrossRef] [PubMed]
  23. Dos Santos, T.B.; Budzinski, I.G.; Marur, C.J.; Petkowicz, C.L.; Pereira, L.F.; Vieira, L.G. Expression of three galactinol synthase isoforms in Coffea arabica L. and accumulation of raffinose and stachyose in response to abiotic stresses. Plant Physiol. Biochem. 2011, 49, 441–448. [Google Scholar] [CrossRef] [PubMed]
  24. Sengupta, S.; Mukherjee, S.; Parween, S.; Majumder, A.L. Galactinol synthase across evolutionary diverse taxa: Functional preference for higher plants. FEBS Lett. 2012, 586, 1488–1496. [Google Scholar] [CrossRef] [PubMed]
  25. Nagaharu, U.; Nagaharu, N. Genome analysis in Brassica with special reference to the experimental formation of B. napus and peculiar mode of fertilization. Jpn. J. Bot. 1935, 7, 389–452. [Google Scholar]
  26. Chalhoub, B.; Denoeud, F.; Liu, S.; Parkin, I.A.; Tang, H.; Wang, X. Early allopolyploid evolution in the post-Neolithic Brassica napus oilseed genome. Science 2014, 345, 950–953. [Google Scholar] [CrossRef] [PubMed]
  27. Sierro, N.; Battey, J.N.; Ouadi, S.; Bakaher, N.; Bovet, L.; Willig, A.; Goepfert, S.; Peitsch, M.C.; Ivanov, N.V. The tobacco genome sequence and its comparison with those of tomato and potato. Nat. Commun. 2014, 5, 3833. [Google Scholar] [CrossRef] [PubMed]
  28. Edwards, K.D.; Fernandez-Pozo, N.; Drake-Stowe, K.; Humphry, M.; Evans, A.D.; Bombarely, A.; Allen, F.; Hurst, R.; White, B.; Kernodle, S.P.; et al. A reference genome for Nicotiana tabacum enables map-based cloning of homologous loci implicated in nitrogen utilization efficiency. BMC Genom. 2017, 18, 448–456. [Google Scholar] [CrossRef] [PubMed]
  29. Zhou, J.; Yang, Y.; Yu, J.; Wang, L.; Yu, X.; Ohtani, M.; Kusano, M.; Saito, K.; Demura, T.; Zhuge, Q. Responses of Populus trichocarpa galactinol synthase genes to abiotic stresses. J. Plant Res. 2014, 127, 347–358. [Google Scholar] [CrossRef] [PubMed]
  30. Guo, Y.; Hans, H.; Christian, J.; Molina, C. Mutations in single FT- and TFL1-paralogs of rapeseed (Brassica napus L.) and their impact on flowering time and yield components. Front. Plant Sci. 2014, 5, 282–290. [Google Scholar] [CrossRef] [PubMed]
  31. Sato, S.; Tabata, S.; Hirakawa, H.; Asamizu, E.; Shirasawa, K.; Isobe, S.; Kaneko, T.; Nakamura, Y.; Shibata, D.; Aoki, K. The tomato genome sequence provides insights into fleshy fruit evolution. Nature 2012, 485, 635–641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Moore, M.J.; Soltis, P.S.; Bell, C.D.; Burleigh, J.G.; Soltis, D.E. Phylogenetic analysis of 83 plastid genes further resolves the early diversification of eudicots. Proc. Natl. Acad. Sci. USA 2010, 107, 4623–4628. [Google Scholar] [CrossRef] [PubMed]
  33. Wiggins, C.A.; Munro, S. Activity of the yeast MNN1 α-1,3-mannosyltransferase requires a motif conserved in many other families of glycosyltransferases. Proc. Natl. Acad. Sci. USA 1998, 95, 7945–7950. [Google Scholar] [CrossRef] [PubMed]
  34. Sterling, J.D.; Atmodjo, M.A.; Inwood, S.E.; Kumar Kolli, V.S.; Quigley, H.F.; Hahn, M.G.; Mohnen, D. Functional identification of an Arabidopsis pectin biosynthetic homogalacturonan galacturonosyltransferase. Proc. Natl. Acad. Sci. USA 2006, 103, 5236–5241. [Google Scholar] [CrossRef] [PubMed]
  35. Yin, Y.; Chen, H.; Hahn, M.G.; Mohnen, D.; Xu, Y. Evolution and function of the plant cell wall synthesis-related glycosyltransferase family 8. Plant Physiol. 2010, 153, 1729–1746. [Google Scholar] [CrossRef] [PubMed]
  36. Yin, Y.; Mohnen, D.; Genelio-Albersheim, I.; Xu, Y.; Hahn, M.G. Glycosyltransferases of the GT8 family. In Annual Plant Reviews: Plant Polysaccharides, Biosynthesis and Bioengineering; Ulvskov, P., Ed.; Wiley-Blackwell: Oxford, UK, 2011; Volume 41, pp. 167–212. [Google Scholar]
  37. Bailey, T.L.; Williams, N.; Misleh, C.; Li, W.W. MEME: Discovering and analyzing DNA and protein sequence motifs. Nucleic Acids Res. 2006, 34, 369–373. [Google Scholar] [CrossRef] [PubMed]
  38. Liang, Y.; Wan, N.; Cheng, Z.; Mo, Y.; Liu, B.; Liu, H.; Raboanatahiry, N.; Yin, Y.; Li, M. Whole-Genome identification and expression pattern of the vicinal oxygen chelate family in rapeseed (Brassica napus L.). Front. Plant Sci. 2017, 8, 735–745. [Google Scholar] [CrossRef] [PubMed]
  39. Kawabe, A.; Miyashita, N.T. Patterns of codon usage bias in three dicot and four monocot plant species. Genes Genet. Syst. 2003, 78, 343–352. [Google Scholar] [CrossRef] [PubMed]
  40. Nekrutenko, A.; Makova, K.D.; Li, W.H. The Ka/Ks ratio test for assessing the protein-coding potential of genomic regions: An empirical and simulation study. Genome Res. 2002, 12, 198–202. [Google Scholar] [CrossRef] [PubMed]
  41. Zhang, Y. I-TASSER server for protein 3D structure prediction. BMC Bioinform. 2008, 9, 40–46. [Google Scholar] [CrossRef] [PubMed]
  42. Gibbons, B.J.; Roach, P.J.; Hurley, T.D. Crystal structure of the autocatalytic initiator of glycogen biosynthesis, glycogenin. Mol. Biol. 2002, 319, 463–477. [Google Scholar] [CrossRef]
  43. Yang, J.; Roy, A.; Zhang, Y. BioLiP: A semi-manually curated database for biologically relevant ligand-protein interactions. Nucleic Acids Res. 2013, 41, 1096–1103. [Google Scholar] [CrossRef] [PubMed]
  44. Lee, H.; Zhang, Y. BSP-SLIM: A blind low-resolution ligand-protein docking approach using theoretically predicted protein structures. Proteins 2012, 80, 93–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Persson, K.; Ly, H.D.; Dieckelmann, M.; Wakarchuk, W.W.; Withers, S.G.; Strynadka, N.C.J. Crystal structure of the retaining galactosyltransferase LgtC from Neisseria meningitidis in complex with donor and acceptor sugar analogs. Nat. Struct. Biol. 2001, 8, 166–175. [Google Scholar] [CrossRef] [PubMed]
  46. Boix, E.; Zhang, Y.; Swaminathan, G.J.; Brew, K.; Acharya, K.R. Structural basis of ordered binding of donor and acceptor substrates to the retaining glycosyltransferase α-1,3-galactosyltransferase. Biol. Chem. 2002, 277, 28310–28318. [Google Scholar] [CrossRef] [PubMed]
  47. Ramakrishnan, B.; Balaji, P.V.; Qasba, P.K. Crystal structure of β-1,4galactosyltransferase complex with UDP–Gal reveals an oligosaccharide binding site. Mol. Biol. 2002, 318, 491–502. [Google Scholar] [CrossRef]
  48. Cheng, F.; Wu, J.; Wang, X. Genome triplication drove the diversification of Brassica plants. Hortic. Res. 2014, 1, 14–24. [Google Scholar] [CrossRef] [PubMed]
  49. Liu, S.; Liu, Y.; Yang, X.; Tong, C.; Edwards, D.; Parkin, I.A. The Brassica oleracea genome reveals the asymmetrical evolution of polyploid genomes. Nat. Commun. 2014, 5, 3930. [Google Scholar] [CrossRef] [PubMed]
  50. Wang, X.; Wang, H.; Wang, J.; Sun, R.; Wu, J.; Liu, S. The genome of the mesopolyploid crop species Brassica rapa. Nat. Genet. 2011, 43, 1035–1039. [Google Scholar] [CrossRef] [PubMed]
  51. Blanc, G.; Wolfe, K.H. Widespread paleopolyploidy in model plant species inferred from age distributions of duplicate genes. Plant Cell 2004, 17, 1667–1678. [Google Scholar] [CrossRef] [PubMed]
  52. Blackman, S.A.; Obendorf, R.L.; Leopold, A.C. Maturation proteins and sugars in desiccation tolerance of developing soybean seeds. Plant Physiol. 1992, 100, 225–230. [Google Scholar] [CrossRef] [PubMed]
  53. Brenac, P.; Horbowicz, M.; Downer, S.M.; Dickerman, A.M.; Smith, M.E.; Obendorf, R.L. Raffinose accumulation related to desiccation tolerance during maize (Zea mays L.) seed development and maturation. Plant Physiol. 1997, 5, 150–158. [Google Scholar] [CrossRef]
  54. Li, X.; Zhuo, J.; Jing, Y.; Liu, X.; Wang, X.F. Expression of a GALACTINOL SYNTHASE gene is positively associated with desiccation tolerance of Brassica napus seeds during development. J. Plant Physiol. 2011, 168, 1761–1770. [Google Scholar] [CrossRef] [PubMed]
  55. Nishizawa, A.; Yabuta, Y.; Shigeoka, S. The contribution of carbohydrates including raffinose family oligosaccharides and sugar alcohols to protection of plant cells from oxidative damage. Plant Signal. Behav. 2008, 3, 1016–1018. [Google Scholar] [CrossRef]
  56. Chabrillange, N.; Dussert, S.; Engelmann, F.; Doulbeau, S.; Hamon, S. Desiccation tolerance in relation to soluble sugar contents in seeds of ten coffee (Coffea L.) species. Seed Sci. Res. 2000, 10, 393–396. [Google Scholar] [CrossRef]
  57. Unda, F.; Canam, T.; Preston, L.; Mansfield, S.D. Isolation and characterization of galactinol synthases from hybrid poplar. Exp. Bot. 2012, 63, 2059–2069. [Google Scholar] [CrossRef] [PubMed]
  58. Blöchl, A.; March, G.G.; Sourdioux, M.; Peterbauer, T.; Richter, A. Induction of raffinose oligosaccharide biosynthesis by abscisic acid in somatic embryos of alfalfa (Medicago sativa L.). Plant Sci. 2005, 168, 1075–1082. [Google Scholar] [CrossRef]
  59. Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Basic local alignment search tool. Mol. Biol. 1990, 215, 403–410. [Google Scholar] [CrossRef]
  60. Altschul, S.F.; Madden, T.L.; Schäffer, A.A.; Zhang, J.; Zhang, Z.; Miller, W.; Lipman, D.J. Gapped BLAST and PSI-BLAST: A new generation of protein database search programs. Nucleic Acids Res. 1997, 25, 3389–3402. [Google Scholar] [CrossRef] [PubMed]
  61. Gasteiger, E.; Gattiker, A.; Hoogland, C.; Ivanyi, I.; Appel, R.D.; Bairoch, A. ExPASy: The proteomics server for in-depth protein knowledge and analysis. Nucleic Acids Res. 2003, 31, 3784–3788. [Google Scholar] [CrossRef] [PubMed]
  62. Hofmann, K.; Stoffel, W. TMbase—A database of membrane spanning proteins segments. Biol. Chem. 1993, 374, 166. [Google Scholar]
  63. Henrik, N. Protein function prediction. Mol. Biol. 2017, 1611, 59–73. [Google Scholar]
  64. Blum, T.; Briesemeister, S.; Kohlbacher, O. MultiLoc2: Integrating phylogeny and gene ontology terms improves subcellular protein localization prediction. BMC Bioinform. 2009, 10, 274. [Google Scholar] [CrossRef] [PubMed]
  65. Hu, B.; Jin, J.; Guo, A.; Zhang, H.; Luo, J.; Gao, G. GSDS 2.0: An upgraded gene feature visualization server. Bioinformatics 2015, 31, 1296–1297. [Google Scholar] [CrossRef] [PubMed]
  66. Bailey, T.L.; Gribskov, M. Combining evidence using p–values: Application to sequence homology searches. Bioinformatics 1998, 14, 48–54. [Google Scholar] [CrossRef] [PubMed]
  67. Edgar, R.C. MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004, 32, 1792–1797. [Google Scholar] [CrossRef] [PubMed]
  68. Guindon, S.; Dufayard, J.F.; Lefort, V.; Anisimova, M.; Hordijk, W.; Gascuel, O. New algorithms and methods to estimate maximum-likelihood phylogenies: Assessing the performance of PhyML 3.0. Syst. Biol. 2010, 59, 307–321. [Google Scholar] [CrossRef] [PubMed]
  69. Lefort, V.; Longueville, J.E.; Gascuel, O. SMS: Smart model selection in PhyML. Mol. Biol. Evol. 2017, 34, 2422–2424. [Google Scholar] [CrossRef] [PubMed]
  70. Thomas, M.K.; Christopher, J.C.; Melissa, M.P.; Thomas, J.N.; James, O.M. Assessment of methods for amino acid matrix selection and their use on empirical data shows that ad hoc assumptions for choice of matrix are not justified. BMC Evol. Biol. 2006, 6, 29. [Google Scholar] [CrossRef]
  71. Wang, D.; Zhang, Y.; Zhang, Z.; Zhu, J.; Yu, J. KaKs_Calculator 2.0: A toolkit incorporating gamma-series methods and sliding window strategies. Genom. Proteom. Bioinform. 2010, 8, 77–80. [Google Scholar] [CrossRef]
  72. Sudhir, K.; Glen, S.; Koichiro, T. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 2016. [Google Scholar] [CrossRef]
  73. Wu, S.; Zhang, Y. LOMETS: A local meta-threading-server for protein structure prediction. Nucleic Acids Res. 2007, 35, 3375–3382. [Google Scholar] [CrossRef] [PubMed]
  74. Roy, A.; Kucukural, A.; Zhang, Y. I-TASSER: A unified platform for automated protein structure and function prediction. Nat. Protoc. 2010, 5, 725–738. [Google Scholar] [CrossRef] [PubMed]
  75. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera-a visualization system for exploratory research and analysis. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [PubMed]
  76. Jian, H.; Lu, K.; Yang, B.; Wang, T.; Zhang, L.; Zhang, A.; Wang, J.; Liu, L.; Qu, C.; Li, J. Genome-Wide analysis and expression profiling of the SUC and SWEET gene families of sucrose transporters in oilseed rape (Brassica napus L.). Front. Plant Sci. 2016, 7, 14–64. [Google Scholar] [CrossRef] [PubMed]
  77. Lu, K.; Li, T.; He, J.; Chang, W.; Zhang, R.; Liu, M.; Yu, M.; Fan, Y.; Ma, J.; Sun, W.; et al. qPrimerDB: A thermodynamics-based gene-specific qPCR primer database for 147 organisms. Nucleic Acids Res. 2017. [Google Scholar] [CrossRef] [PubMed]
  78. Bustin, S.A.; Benes, V.; Garson, J.A.; Hellemans, J.; Huggett, J.; Kubista, M.; Mueller, R.; Nolan, T.; Pfaffl, M.W.; Shipley, G.L.; et al. The MIQE guidelines: Minimum Information for publication of quantitative real-time PCR experiments. Clin. Chem. 2009, 55, 611–622. [Google Scholar] [CrossRef] [PubMed]
  79. Lu, K.; Xiao, Z.; Jian, H.; Peng, L.; Qu, C.; Fu, M.; He, B.; Tie, L.; Liang, Y.; Xu, X.; et al. A combination of genome-wide association and transcriptome analysis reveals candidate genes controlling harvest index-related traits in Brassica napus. Sci. Rep. 2016, 6, 36452. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Phylogenetic, gene structure and conserved motifs of GolS proteins in A. thaliana, B. napus and N. tabacum. (A) Amino acid sequences of AtGolS, BnGolS and NtGolS were aligned using MUSCLE. The phylogenetic tree was constructed with the online PhyML server with bootstrap analysis (100 replicates) and displayed using FigTree v1.4.0. The 36 GolS proteins from A. thaliana, B. napus and N. tabacum (GolS in A. trichopoda as an outgroup) clustered into four distinct groups; (B) Gene structures were generated by the Gene Structure Display Server. Exons (CDS) and introns are shown with green wedges and black lines, respectively. Numbers 0, 1 and 2 represent the introns in phases 0, 1 and 2, respectively. The scale bar represents 1.0 kb. At: A. thaliana; Bn: B. napus; Nt: N. tabacum; AmTr: A. trichopoda. (C) Conserved motifs in AtGolS, BnGolS and NtGolS proteins were identified by MEME. A colored box indicates the different motifs that are numbered along the bottom.
Figure 1. Phylogenetic, gene structure and conserved motifs of GolS proteins in A. thaliana, B. napus and N. tabacum. (A) Amino acid sequences of AtGolS, BnGolS and NtGolS were aligned using MUSCLE. The phylogenetic tree was constructed with the online PhyML server with bootstrap analysis (100 replicates) and displayed using FigTree v1.4.0. The 36 GolS proteins from A. thaliana, B. napus and N. tabacum (GolS in A. trichopoda as an outgroup) clustered into four distinct groups; (B) Gene structures were generated by the Gene Structure Display Server. Exons (CDS) and introns are shown with green wedges and black lines, respectively. Numbers 0, 1 and 2 represent the introns in phases 0, 1 and 2, respectively. The scale bar represents 1.0 kb. At: A. thaliana; Bn: B. napus; Nt: N. tabacum; AmTr: A. trichopoda. (C) Conserved motifs in AtGolS, BnGolS and NtGolS proteins were identified by MEME. A colored box indicates the different motifs that are numbered along the bottom.
Ijms 18 02768 g001
Figure 2. Phylogenetic relationship among GolS proteins in plants. The ML tree was generated with bootstrap analysis (100 replicates) using aligned GolS protein sequences from A. thaliana, B. napus, B. rapa, B. oleracea, N. tabacum, S. lycopersicum, S. tuberosum, O. sativa and Z. mays (GolS in A. trichopoda as an outgroup) using the online PhyML server. The tree was displayed with FigTree v1.4.0. GolS proteins in the phylogenetic tree clustered into four groups (Group 1, Group 2, Group 3 and Group 4). At: A. thaliana; Bn: B. napus; Bra: B. rapa; Bol: B. oleracea; Nt: N. tabacum; Sl: S. lycopersicum; St: S. tuberosum; Os: O. sativa; Zm: Z. mays; AmTr: A. trichopoda.
Figure 2. Phylogenetic relationship among GolS proteins in plants. The ML tree was generated with bootstrap analysis (100 replicates) using aligned GolS protein sequences from A. thaliana, B. napus, B. rapa, B. oleracea, N. tabacum, S. lycopersicum, S. tuberosum, O. sativa and Z. mays (GolS in A. trichopoda as an outgroup) using the online PhyML server. The tree was displayed with FigTree v1.4.0. GolS proteins in the phylogenetic tree clustered into four groups (Group 1, Group 2, Group 3 and Group 4). At: A. thaliana; Bn: B. napus; Bra: B. rapa; Bol: B. oleracea; Nt: N. tabacum; Sl: S. lycopersicum; St: S. tuberosum; Os: O. sativa; Zm: Z. mays; AmTr: A. trichopoda.
Ijms 18 02768 g002
Figure 3. Inferred origin and evolutionary relationship of GolS genes and their copy number change among nine plants. The digits represent the number of GolS genes in plants species. Triangle represents the original genes of GolS in plants, while the circle represents the duplicated genes. The blue line indicates that B. napus is formed by B. rape and B. oleracea.
Figure 3. Inferred origin and evolutionary relationship of GolS genes and their copy number change among nine plants. The digits represent the number of GolS genes in plants species. Triangle represents the original genes of GolS in plants, while the circle represents the duplicated genes. The blue line indicates that B. napus is formed by B. rape and B. oleracea.
Ijms 18 02768 g003
Figure 4. Distribution of BnGolS gene family members along the B. napus chromosomes and synteny map of GolS genes from A. thaliana B. napus, B. rapa and B. oleracea. (A) Chromosomal information for BnGolS genes was obtained from the Brassica database and was mapped to B. napus chromosomes. Syntenic relationships are indicated with connecting lines; (B) Genes located within the B. napus genome that are syntenic with genes of A. thaliana, B. rapa and B. oleracea are indicated by connecting lines.
Figure 4. Distribution of BnGolS gene family members along the B. napus chromosomes and synteny map of GolS genes from A. thaliana B. napus, B. rapa and B. oleracea. (A) Chromosomal information for BnGolS genes was obtained from the Brassica database and was mapped to B. napus chromosomes. Syntenic relationships are indicated with connecting lines; (B) Genes located within the B. napus genome that are syntenic with genes of A. thaliana, B. rapa and B. oleracea are indicated by connecting lines.
Ijms 18 02768 g004
Figure 5. 3D Structure predictions for BnGolS and NtGolS. BnGolS1-2 and NtGolS1-1 were selected as the representative GolS proteins from B. napus and N. tabacum, respectively. The models were predicted by I-TASSER and the rabbit muscle glycogenin structure (PDB ID1ll0) was used as the template for the 3D structure predication. The conserved DXD and HxxGxxPW motifs are marked on the 3D structure in red. Green represents α-helices, yellow represents β-strands and navy blue represents random coils. (A) Modeled3D structure of BnGolS1-2; (B) Modeled 3D structure of NtGolS1-1; (C) Template model of 1ll0B. Structural images were generated with Chimera 1.2.
Figure 5. 3D Structure predictions for BnGolS and NtGolS. BnGolS1-2 and NtGolS1-1 were selected as the representative GolS proteins from B. napus and N. tabacum, respectively. The models were predicted by I-TASSER and the rabbit muscle glycogenin structure (PDB ID1ll0) was used as the template for the 3D structure predication. The conserved DXD and HxxGxxPW motifs are marked on the 3D structure in red. Green represents α-helices, yellow represents β-strands and navy blue represents random coils. (A) Modeled3D structure of BnGolS1-2; (B) Modeled 3D structure of NtGolS1-1; (C) Template model of 1ll0B. Structural images were generated with Chimera 1.2.
Ijms 18 02768 g005
Figure 6. Docking of ligands onto the modeled BnGolS1-2 protein structure. (A) Docking results with inositol; (B) Docking results with UDP-galactose; (C,D) Surface representation of BnGolS1-2 showing that the ligands are buried deep in the binding pocket. In BnGolS1-2 proteins, the binding positions for UDP-galactose and inositol are close to the DxD motif. The inositol and UDP-galactose binding sites are represented in blue and the DXD and HxxGxxPW motifs are marked on the surface in red. Images were generated with Chimera 1.2.
Figure 6. Docking of ligands onto the modeled BnGolS1-2 protein structure. (A) Docking results with inositol; (B) Docking results with UDP-galactose; (C,D) Surface representation of BnGolS1-2 showing that the ligands are buried deep in the binding pocket. In BnGolS1-2 proteins, the binding positions for UDP-galactose and inositol are close to the DxD motif. The inositol and UDP-galactose binding sites are represented in blue and the DXD and HxxGxxPW motifs are marked on the surface in red. Images were generated with Chimera 1.2.
Ijms 18 02768 g006
Figure 7. Tissue-specific and hormone-induced expression patterns of BnGolS and NtGolS genes. (A) Expression profiles of BnGolS genes in B. napus; (B) Expression profiles of NtGolS genes in N. tabacum; (C) Expression profiles of BnGolS genes in response to hormone treatments. The color bar to the right of the figures represents the log2 expression value and the green color represents the low or no expression in (A) and (B) while it represents the down-regulation in hormone treatment in (C).
Figure 7. Tissue-specific and hormone-induced expression patterns of BnGolS and NtGolS genes. (A) Expression profiles of BnGolS genes in B. napus; (B) Expression profiles of NtGolS genes in N. tabacum; (C) Expression profiles of BnGolS genes in response to hormone treatments. The color bar to the right of the figures represents the log2 expression value and the green color represents the low or no expression in (A) and (B) while it represents the down-regulation in hormone treatment in (C).
Ijms 18 02768 g007
Table 1. Characteristics of the twenty BnGolS and nine NtGolS proteins identified in this study.
Table 1. Characteristics of the twenty BnGolS and nine NtGolS proteins identified in this study.
Gene NameGene IDOrthologues in ArabidopsisChromosome LocationNumber of IntronsNumber of Amino AcidsMolecular Weights (Mw)Isoelectric Point (pI)Subcellular Localization
BnGolS1-1BnaA04g26930DAT2G47180 (AtGolS1)chrA04: 19029593–19030985334239.265.23cytoplasmic
BnGolS1-2BnaA05g00720DAT2G47180 (AtGolS1)chrA05: 400739–402104333838.786.34cytoplasmic
BnGolS1-3BnaC04g00330DAT2G47180 (AtGolS1)chrC04: 214830–216198333838.885.95cytoplasmic
BnGolS1-4BnaC04g51460DAT2G47180 (AtGolS1)chrC04: 48728752–48730072334239.155.52cytoplasmic
BnGolS2-1BnaA09g15290DAT1G56600 (AtGolS2)chrA09: 8894797–8896969333037.866.76cytoplasmic
BnGolS2-2BnaCnng63310DAT1G56600 (AtGolS2)chrCnn: 63178898–63180749333037.916.40cytoplasmic
BnGolS3-1BnaA09g48480DAT1G09350 (AtGolS3)chrA09: 32476358–32477888433038.096.66cytoplasmic
BnGolS3-2BnaC08g50010DAT1G09350 (AtGolS3)chrC08: 4447550–4449386433038.087.03mitochondrial
BnGolS4-1BnaA09g14180DAT1G60470 (AtGolS4)chrA09: 8110757–8112355333438.344.69cytoplasmic
BnGolS4-2BnaA09g14190DAT1G60470 (AtGolS4)chrA09: 8119504–8120800329734.695.10cytoplasmic
BnGolS4-3BnaC09g14710DAT1G60470 (AtGolS4)chrC09: 11303590–11304914333438.374.70cytoplasmic
BnGolS4-4BnaC09g14720DAT1G60470 (AtGolS4)chrC09: 11312323–11313549329734.685.11cytoplasmic
BnGolS6-1BnaA08g14430DAT4G26250 (AtGolS6)chrA08: 12234029–12235330333738.774.89cytoplasmic
BnGolS6-2BnaA09g41310DAT4G26250 (AtGolS6)chrA09: 28893974–28895377333638.434.87cytoplasmic
BnGolS6-3BnaAnng12190DAT4G26250 (AtGolS6)chrAnn: 13192498–13193777333838.855.13cytoplasmic
BnGolS6-4BnaC08g12130DAT4G26250 (AtGolS6)chrC08: 17362334–17362334333738.744.94cytoplasmic
BnGolS6-5BnaC08g33920DAT4G26250 (AtGolS6)chrC08: 32207643–32209044333638.514.92cytoplasmic
BnGolS6-6BnaC08g50120DAT4G26250 (AtGolS6)chrC08: 4493099–4494386333738.795.00cytoplasmic
BnGolS7-1BnaA01g22350DAT1G60450 (AtGolS7)chrA01: 14694496–14694861012113.776.37cytoplasmic
BnGolS7-2BnaC01g28520DAT1G60450 (AtGolS7)chrC01: 26460047–26467911333137.875.05cytoplasmic
NtGolS1-1Nitab4.5_0000222g0170AT2G47180 (AtGolS1)chr19: 69267186–69268901334339.245.63cytoplasmic
NtGolS1-2Nitab4.5_0001013g0090AT2G47180 (AtGolS1)chr22: 93811889–93813765334339.345.51cytoplasmic
NtGolS1-3Nitab4.5_0003324g0150AT2G47180 (AtGolS1)Scaffold0003324: 299742–303780322325.266.52cytoplasmic
NtGolS2-1Nitab4.5_0000136g0290AT1G56600 (AtGolS2)chr17: 88626458–88628277229333.506.95cytoplasmic
NtGolS2-2Nitab4.5_0000178g0340AT1G56600 (AtGolS2)chr17: 192992786–192996672436041.405.94cytoplasmic
NtGolS2-3Nitab4.5_0001617g0060AT1G56600 (AtGolS2)Scaffold0001617: 393036–394255233337.885.18cytoplasmic
NtGolS2-4Nitab4.5_0003044g0080AT1G56600 (AtGolS2)Scaffold0003044: 139690–141464232136.576.71cytoplasmic
NtGolS2-5Nitab4.5_0008397g0020AT1G56600 (AtGolS2)Scaffold0008397: 82555–87780231635.915.64cytoplasmic
NtGolS2-6Nitab4.5_0011298g0020AT1G56600 (AtGolS2)Scaffold0011298: 15196–17559334239.157.07cytoplasmic
Table 2. Nucleotide substitution rates for BnGolS and NtGolS genes.
Table 2. Nucleotide substitution rates for BnGolS and NtGolS genes.
A. thaliana IDGene NameBnGolS or NtGolS IDKaKsKa/Ks
TWO-COPY LOCI
AT1G56600BnGolS2-1BnaA09g15290D0.07692670.5177060.148591
AT1G56600BnGolS2-2BnaCnng63310D0.08420080.4925350.170954
AT1G09350BnGolS3-1BnaA09g48480D0.04800830.3613380.132862
AT1G09350BnGolS3-2BnaC08g50010D0.05133240.3925960.130751
AT1G60450BnGolS7-1BnaA01g22350D0.1136690.593310.191585
AT1G60450BnGolS7-2BnaC01g28520D0.09219330.4128720.223298
FOUR-COPY LOCI
AT2G47180BnGolS1-1BnaA04g26930D0.9633561.121970.858632
AT2G47180BnGolS1-2BnaA05g00720D0.05342710.4891770.109218
AT2G47180BnGolS1-3BnaC04g00330D0.0551520.4691880.117548
AT2G47180BnGolS1-4BnaC04g51460D0.9691391.10670.875705
AT1G60470BnGolS4-1BnaA09g14180D0.03682750.3134530.11749
AT1G60470BnGolS4-2BnaA09g14190D0.1319780.6307960.209224
AT1G60470BnGolS4-3BnaC09g14710D0.03507850.283710.123642
AT1G60470BnGolS4-4BnaC09g14720D0.1340250.6077960.22051
SIX-COPY LOCI
AT4G26250BnGolS6-1BnaA08g14430D0.09608650.8526380.112693
AT4G26250BnGolS6-2BnaA09g41310D0.1185290.7068390.167689
AT4G26250BnGolS6-3BnaAnng12190D0.09200220.7934890.115946
AT4G26250BnGolS6-4BnaC08g12130D0.09725590.8401120.115765
AT4G26250BnGolS6-5BnaC08g33920D0.1221010.7793040.156679
AT4G26250BnGolS6-6BnaC08g50120D0.08759040.8074850.108473
THREE-COPY LOCI
AT2G47180NtGolS1-1Nitab4.5_0000222g01700.1238793.517590.035217
AT2G47180NtGolS1-2Nitab4.5_0001013g00900.1170143.71930.031461
AT2G47180NtGolS1-3Nitab4.5_0003324g01500.1787673.640080.049111
SIX-COPY LOCI
AT1G56600NtGolS2-1Nitab4.5_0000136g02900.1584663.792290.041786
AT1G56600NtGolS2-2Nitab4.5_0000178g03400.187823.559220.05277
AT1G56600NtGolS2-3Nitab4.5_0001617g00600.1544973.646050.042374
AT1G56600NtGolS2-4Nitab4.5_0003044g00800.2494953.168280.078748
AT1G56600NtGolS2-5Nitab4.5_0008397g00200.1731053.476010.0498
AT1G56600NtGolS2-6Nitab4.5_0011298g00200.2588473.189280.081161

Share and Cite

MDPI and ACS Style

Fan, Y.; Yu, M.; Liu, M.; Zhang, R.; Sun, W.; Qian, M.; Duan, H.; Chang, W.; Ma, J.; Qu, C.; et al. Genome-Wide Identification, Evolutionary and Expression Analyses of the GALACTINOL SYNTHASE Gene Family in Rapeseed and Tobacco. Int. J. Mol. Sci. 2017, 18, 2768. https://doi.org/10.3390/ijms18122768

AMA Style

Fan Y, Yu M, Liu M, Zhang R, Sun W, Qian M, Duan H, Chang W, Ma J, Qu C, et al. Genome-Wide Identification, Evolutionary and Expression Analyses of the GALACTINOL SYNTHASE Gene Family in Rapeseed and Tobacco. International Journal of Molecular Sciences. 2017; 18(12):2768. https://doi.org/10.3390/ijms18122768

Chicago/Turabian Style

Fan, Yonghai, Mengna Yu, Miao Liu, Rui Zhang, Wei Sun, Mingchao Qian, Huichun Duan, Wei Chang, Jinqi Ma, Cunmin Qu, and et al. 2017. "Genome-Wide Identification, Evolutionary and Expression Analyses of the GALACTINOL SYNTHASE Gene Family in Rapeseed and Tobacco" International Journal of Molecular Sciences 18, no. 12: 2768. https://doi.org/10.3390/ijms18122768

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