Next Article in Journal
Phyto-SERM Constitutes from Flemingia macrophylla
Next Article in Special Issue
Arbuscular Mycorrhizal Fungi and Plant Growth-Promoting Pseudomonads Increases Anthocyanin Concentration in Strawberry Fruits (Fragaria x ananassa var. Selva) in Conditions of Reduced Fertilization
Previous Article in Journal
Effects of Fe3O4 Magnetic Nanoparticles on A549 Cells
Previous Article in Special Issue
Protein Synthesis Inhibition Activity by Strawberry Tissue Protein Extracts during Plant Life Cycle and under Biotic and Abiotic Stresses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genetic Analysis of Health-Related Secondary Metabolites in a Brassica rapa Recombinant Inbred Line Population

1
Laboratory of Genetics, Wageningen University, Wageningen, 6700 AH, The Netherlands
2
Department of Biotechnology, College of Agriculture, Bu-Ali Sina University, Hamedan 65174, Iran
3
Faculty of Agriculture, Department of Genetics, Cairo University, Giza, Gamaa St. 12613, Egypt
4
Natural Product Laboratory, Institute of Biology, Leiden University, Leiden, Sylviusweg 72, 2333BE, The Netherlands
5
Plant Breeding Institute, Christian Albrechts University, Kiel, Olshausenstrasse 40, D-24098, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2013, 14(8), 15561-15577; https://doi.org/10.3390/ijms140815561
Submission received: 27 May 2013 / Revised: 9 July 2013 / Accepted: 17 July 2013 / Published: 25 July 2013
(This article belongs to the Special Issue Molecular Research in Plant Secondary Metabolism)

Abstract

:
The genetic basis of the wide variation for nutritional traits in Brassica rapa is largely unknown. A new Recombinant Inbred Line (RIL) population was profiled using High Performance Liquid Chromatography (HPLC) and Nuclear Magnetic Resonance (NMR) analysis to detect quantitative trait loci (QTLs) controlling seed tocopherol and seedling metabolite concentrations. RIL population parent L58 had a higher level of glucosinolates and phenylpropanoids, whereas levels of sucrose, glucose and glutamate were higher in the other RIL population parent, R-o-18. QTL related to seed tocopherol (α-, β-, γ-, δ-, α-/γ- and total tocopherol) concentrations were detected on chromosomes A3, A6, A9 and A10, explaining 11%–35% of the respective variation. The locus on A3 co-locates with the BrVTE1gene, encoding tocopherol cyclase. NMR spectroscopy identified the presence of organic/amino acid, sugar/glucosinolate and aromatic compounds in seedlings. QTL positions were obtained for most of the identified compounds. Compared to previous studies, novel loci were found for glucosinolate concentrations. This work can be used to design markers for marker-assisted selection of nutritional compounds in B. rapa.

Graphical Abstract

1. Introduction

Brassica rapa is a valuable source of health-promoting metabolites, like antioxidants, vitamins or glucosinolates. Plants, in general, produce an amazing diversity of low molecular mass natural compounds [1]. Over 100,000 metabolites have been detected [2], for which the structures of close to 50,000 have been elucidated [3]. Many of these compounds are part of secondary metabolic pathways, which are not directly involved in or essential for the central metabolic processes of the plant, but they play very important roles in ecological interactions, such as plant defense against pathogens and herbivores and the response to abiotic factors. This is in contrast to primary metabolites, such as carbohydrates, vitamins, amino and organic acids, which are found in all plants and are directly involved in normal growth, development and reproduction.
Plant secondary metabolites are of nutritional value for humans, due to their anticancer and health-promoting properties. Tocopherols, for instance, are essential nutrients that humans can only obtain via food. The α-, β-, γ- and δ-tocopherols produced by plants are jointly known as vitamin E. Seeds generally provide the bulk of tocopherols to the human diet [4,5]. They are lipid-soluble amphipathic molecules that can act as antioxidants [68]. Tocopherol content and composition can be determined accurately by High Performance Liquid Chromatography (HPLC) [9]. α-tocopherol is the most interesting component, as it is selectively taken up in the human liver, and its biological activity is 2–50 times higher than that of the others [4]. Generally, no β-tocopherol is found in B. napus and only very small amounts (<1%) of δ-tocopherol [10]. The average contributions of γ-tocopherol and α-tocopherol to total tocopherol in rape seed oil are 65% and 35%, respectively. The α- to γ-tocopherol concentration ratios in rapeseed vary from 0.32 to 1.40 mg kg−1, depending on growth condition and genotype [11]. Since γ-tocopherol has a ten-fold lower biological activity than α-tocopherol, increasing the α-tocopherol fraction is a potentially interesting breeding goal in order to improve the nutritional value of crop plants, like rapeseed [10] and B. rapa. Next to increasing α-tocopherol, to improve nutritional purposes, another interesting goal is to increase δ-and γ-tocopherol concentrations, to improve oil stability. γ-tocopherol is known to be the direct precursor of α-tocopherol [12]. The enzyme, γ-tocopherol methyltransferase (gTMT), catalyzes the conversion from δ- to β- or from γ- to α-tocopherol [13,14]. Some five to seven different quantitative trait loci (QTLs) with additive and/or epistatic effects have been mapped for γ-tocopherol and total tocopherol content and for the α-/γ-tocopherol ratio, in a doubled haploid (DH) population of B. napus [15]. Recently, about 50 QTL and associated markers for tocopherol content and composition have been detected in another DH population of B. napus and its reconstructed F2 [16]. B. rapa is also related to the reference plant species, Arabidopsis thaliana, in which fourteen QTLs affecting seed tocopherol content and composition have been identified, in two recombinant inbred line (RIL) populations [4]. The genes of the tocopherol synthesis pathways have been identified and cloned by mutation studies and genomic-based approaches in A. thaliana and Synechocystis sp. PCC 6803. [5,6,14,1725]. Furthermore, the first Brassica gene involved in tocopherol biosynthesis has been cloned, the B. napus VTE4.a1 gene encoding a γ-tocopherol methyltransferase [26].
Next to tocopherol, there are many more secondary plant metabolites in the plant metabolome that are suggested to have a nutritional effect. Of particular interest are glucosinolates, sulfur containing plant metabolites with anti-carcinogenic properties [27,28] that form a group of more than 100 plant secondary metabolites present primarily in the Brassicaceae family. Each plant species contains a blend of different glucosinolates [29,30]. This blend is largely responsible for the typical flavor and odor of Brassicaceae species plant products. There are significant differences within the Brassicaceae crop species for their glucosinolate profiles [31]. Glucosinolates are grouped into three chemical classes: aliphatic, indolic and aromatic, according to whether their amino acid precursor is methionine, tryptophan or an aromatic amino acid (tyrosine or phenylalanine), respectively [32,33]. Aliphatic glucosinolates are the most prominent glucosinolates found in Brassica vegetables [34]. The concentration and chemical structure can vary considerably, depending on the genotype, stage of development, tissue type and environmental conditions [35,36]. More than 90 different aliphatic glucosinolates have been identified among plants [30] of which up to 16 are found in B. rapa [3740]. QTL mapping of leaf aliphatic glucosinolate loci has been carried out in two doubled haploid (DH) populations of B. rapa, which identified 16 loci controlling aliphatic glucosinolate concentration [39]. So far, 102 genes putatively involved in glucosinolate biosynthesis have been identified by comparative genomic analyses in B. rapa as the orthologues of 52 of such genes in A. thaliana [41].
To get unambiguous structural information about a metabolite, Nuclear Magnetic Resonance (NMR), and, particularly, proton NMR (1H NMR analysis), probably among the most common methods, as it is a non-destructive method and can simultaneously detect all proton-bearing compounds [42]. Although it has a lower sensitivity compared to Mass Spectrometry (MS) [43], 1H NMR spectroscopy has previously been used to uncover qualitative and quantitative differences of various cultivars of B. rapa. Different cultivars could be distinguished by elucidated metabolites, for instance, several organic and amino acids, carbohydrates, adenine, indole acetic acid (IAA), phenylpropanoids, flavonoids and glucosinolates [44].
We have used this technique to analyze the genetic variation for a range of (secondary) metabolites in B. rapa seedlings of a recently developed RIL population [45]. In addition, a targeted approach, to detect tocopherols, was used to analyze variation for these compounds in seeds of the same population. As the complete genome sequence of B. rapa is available [46], our analysis will simplify the identification of candidate genes that can be used for genetic modification or marker-assisted breeding for improved nutritional quality of B. rapa.

2. Results and Discussion

2.1. Seed Tocopherol Concentrations

We analyzed seeds of the parental lines, L58 and R-o-18, and all individual lines of the L58 × R-o-18 RIL population [45] for tocopherol content (Table 1, Figure 1). L58 showed higher levels than R-o-18 for α-, γ- and total tocopherol. Some lines showed a very high α-tocopherol concentration in comparison to the other components. Transgression beyond the parental values was observed for all measured tocopherols, except δ-tocopherol (Figure 1), suggesting both parents to contain both positive and negative alleles of genes involved in tocopherol biosynthesis. An example of this is the contrasting alleles found at the two major QTLs for α and total tocopherol, respectively, on A6 and A9. This observation also indicates a potential for improvement of vitamin E content and tocopherol composition through classical breeding, by combining both positive alleles in one genotype.
Correlation analysis revealed that total tocopherol concentration was highly positively correlated with the concentrations of α- and γ-tocopherol (Table 2).

2.2. QTL Analysis of Seed Tocopherol Concentrations

Significant variation was observed for all tocopherol components, as indicated by the broad sense heritability (Table 3). Each tocopherol component was subjected to QTL analysis, and QTL related to α-, β-, γ-, δ-, α-/γ- or total tocopherol concentrations were detected on chromosomes A3, A6, A9 and A10 (Table 3, Figure 2). About 45% of the phenotypic variance for α-tocopherol was explained by two QTLs (Al1 and Al2, respectively, on chromosomes A9 and A6). Two QTLs were found for total tocopherol (Toc1 and Toc2), explaining almost 42% of the tocopherol variance. Toc2 co-located with Al1, but Toc1 did not co-locate with Al2, although both mapped to A6. Instead, it co-located with the Ga1 locus for γ-tocopherol. The QTL for δ-tocopherol (De2) mapped to the same region of A9 to which also Al1 and Toc2 were mapped. This region also contains a strong seed coat color QTL (SC1) [45]. The seed color locus, SC1, probably encodes for the CCR1 gene, a gene involved in lignin biosynthesis [47]. Since there is no reason to suggest a common biochemical basis of biosynthesis of tocopherol and the flavonoids contributing to seed color, a close linkage of different genes, rather than one common gene with pleiotropic effects, is the most likely explanation for this co-location.
As the α-tocopherol concentration is highly positively correlated to the total tocopherol concentration and two of their respective QTLs (Al1 and Toc2) map to the same position, the concentration of α-tocopherol, and not of the intermediate γ-tocopherol, appears to give the major contribution to the overall tocopherol concentration. However, the second Toc locus (Toc1) co-locates with the Ga1 QTL for γ-tocopherol on A6. This means that QTL for both tocopherols with the highest concentrations make a major contribution to the genetic variation for total tocopherol concentrations. The absence of a significant correlation between α-, γ- and δ tocopherol concentrations and the finding that these are controlled by different QTL indicates their independent genetic regulation, which is in agreement with findings of Marwede et al. [15] in canola (B. napus). Thus, with three independent loci controlling α- and γ-tocopherol, it should be possible to enhance the concentration of both. This will have a negative effect on δ-tocopherol concentration though, since the co-locating De2 and Al1 loci have opposite allele effects. As there are RILs with separated contrasting alleles in this population, we could verify this expectation. The tocopherol analyses of these lines confirm our prediction. A similar antagonistic effect was seen for soybean, where overexpression of the AtVTE3 gene, encoding the tocopherol biosynthetic enzyme, 2-methyl-6-phytylbenzoquinol methyltransferase, causes a decrease in seed β- and δ-tocopherol with a proportionate increase in α- and γ-tocopherol [5]. The combination of Al1 and Toc2 alleles from the R-o-18 parent leads to the highest tocopherol concentration in this population.

2.3. NMR Results of Seedling Metabolites Detection

To further assess the variation in metabolites present in the B. rapa RIL population, we performed NMR analysis on young seedlings. Usually, an NMR spectrum consists of hundreds of signals. Among these, 17 compounds in the organic/amino acid, sugar/glucosinolate and aromatic regions of the NMR spectra could be annotated by 1H-NMR and confirmed their structures using 2D NMR spectroscopy. 1H-NMR data of RIL seedling metabolites were subjected to principal component analysis (PCA) (Figure 2). The score plot of the 1H-NMR spectra showed that the two parental lines were quite distinct, especially in principal component 2 (PC2), which was mainly composed of progoitrin, phenylpropanoids and organic compounds. PC1 mostly corresponded to neoglucobrassicin.
L58 had a higher concentration of glucosinolates and phenylpropanoids, whereas the concentrations of sucrose, glucose and glutamate were higher in R-o-18. The major phenylpropanoid was sinapoyl glucose. Correlation analysis showed that the concentrations of several seedling metabolites were highly positively correlated (Table 4).

2.4. QTL Analysis of Seedling Metabolites

Genetic analysis of 238 signals detected in the NMR spectra enabled the identification of QTL for 146 signals (Table 5, Figure 3). A strong QTL for a compound belonging to the phenylpropanoids was mapped on A7, explaining 43% of the phenotypic variance. Six QTL contributing to variation for alanine, asparagine, glutamine, isoleucine, threonine and valine were detected, explaining up to 37% of the variance. QTL analysis of the glucosinolate NMR signals detected several significant loci, with the most significant one on A9 for neoglucobrassicin. In total, six QTLs for glucosinolates (progoitrin and neoglucobrassicin) were mapped to A3, A5, A9 and A10, with the ones mapping to A3 and A5 possibly co-locating. Previously, five B. rapa QTLs related to progoitrin were mapped to chromosomes A1, A3, A4, A8 and A10 [39] using a DH population made from different parents compared to the parents we used to generate the tested RIL population. The authors used forty-day-old leaves for metabolite analysis, while we used young seedlings, which may even still carry glucosinolates originally present in the seed. Therefore, the differences in population and sampled material are considerable, which are the most likely reasons for the differences in detected loci. In any cases, the QTL on A5 and A9 for progoitrin concentration are new loci that have not been reported previously. Extensive studies on aliphatic glucosinolates in A. thaliana previously identified genes encoding AOP (2-oxoglutarate-dependent dioxygenase) and MAM (methyl-thioalkylmalate synthase), controlling the modification of side-chain moiety and elongation, respectively, as important factors contributing to genetic variation for glucosinolate concentration and composition [4852]. The regulation of aliphatic glucosinolate biosynthesis enzymes is controlled in Arabidopsis by the R2R3 myb-like transcription factors, MYB28 and MYB29 [53]. The B. rapa orthologues of MYB28 were mapped on A3, A9 and A2, and the orthologues of MYB29 were mapped on A10 and A3 [41]. The progoitrin QTL presented on A3 with a peak position at 95 cM, co-located with the map positions of MYB28/MYB29; there is also a MAM gene in this region.
QTL for the essential amino acids, isoleucine and valine, are co-located on A3 and A4. The isoleucine biosynthesis pathway runs almost parallel to valine biosynthesis, except for its first steps, which involve a threonine deaminase and dehydratase. These loci possibly correspond to the genes encoding the biosynthetic threonine dehydrates (TD) isozyme, similar to what has been isolated from tomato and potato [54,55]. Arabidopsis gene AT3G10050, the threonine dehydratase biosynthetic gene, has the syntenic paralog in B. rapa on A3, where isoleucine and valine QTL co-located [56]. Non-essential amino acids, such as alanine, asparagine and glutamine, are equally important as the essential amino acids in our body. Eight QTL for non-essential amino acids were identified in this RIL population. These were all independent, except for one QTL on A7, which was shared between alanine and glutamine. At this same region on A7, also, one of the glutamate QTL was mapped. As glutamate is the substrate for glutamine synthesis and the α-amino group of glutamate can be transferred to pyruvate to form alanine [57], this locus is likely to contain a gene involved in the regulation of all three compounds, which is most likely in the upstream, common part of their biosynthesis pathway.

3. Experimental Section

3.1. Plant Material

The RIL population was derived from a cross between two genotypes belonging to two distinct morphotypes, Cai Xin and Yellow Sarson; both early flowering and self-compatible. The Cai Xin parent is L58, a vegetable type originating from China (B. rapa ssp. parachinensis). The other parent, R-o-18, is a doubled haploid Yellow Sarson oil type line (B. rapa ssp. trilocularis) originating from India. This population has been described [45].

3.2. Seed Preparation for HPLC

F7 seeds derived from one plant per RIL of the L58 × R-o-18 population were used for tocopherol measurement (two replicate plants per line with two technical replicates from each plant). For the tocopherol extraction, 10–40 mg seeds were ground in 2-mL reaction tubes with a Geno/Grinder 2000 (SPEX-Sample Prep, Metuchen, NJ, USA) using n-heptane and 3.0–4.0 mm metal beads. The samples were incubated at −20 °C for 2 h. Further applications and HPLC analyses were performed as described [5860]. Quantification of the tocopherols was done by fluorescence detection (excitation at λ = 290 nm, emission at λ = 328 nm). To identify the individual tocopherols, the retention times were compared with standard substances from Merck’s tocopherol kit (Merck, Darmstadt, Germany). Total tocopherol content was calculated as the sum of α-, β-, γ- and δ-tocopherol.

3.3. Seedling Preparation for NMR Analysis

Thirty seeds per RIL of the B. rapa L58 × R-o-18 were used. Seeds were surface sterilized with 70% ethanol (v/v) for 30 s, followed by agitation for 5 min in sodium hypochlorite (2.0% active chlorite). After three rinses in sterile distilled water, 30 seeds of each individual (for every experiment) were placed in 15 × 90 mm petri dishes, each containing 20–25 mL half strength MS salts and vitamins, without sucrose and solidified with 0.8% (w/v) agar. Petri dishes were placed vertically in a growth chamber maintained at 25 °C with a 16 h light/8 h dark photoperiod at a light intensity of 60 mEm−2 s−1. Five-day-old seedlings without roots were harvested and freeze-dried.
20 mg seedlings (dry weight) were extracted with a mixture of 500 μL methanol-d4 and 500 μL D2O (KH2PO4 buffer, pH 6.0) containing 0.05% TSP (trimethyl silyl propionic acid sodium salt, w/v) by ultra-sonication for 20 min. After centrifugation, 800 μL supernatant was transferred to an NMR tube. 1H NMR spectra were recorded at 25 °C on a 600 MHz Bruker AV600 spectrometer equipped with a cryoprobe, operating at a proton NMR frequency of 600.13 MHz. CD3OD was used as the internal lock. Each 1H NMR spectrum consisted of 128 scans using the following parameters: TD = 51,200, spectrum width = 16.02 ppm, 0.25 Hz/point, pulse width (PW) = 30° (6.6 μs), acquisition time = 1.70 s. and relaxation delay (RD) = 2.0 s. A pre-saturation sequence was used to suppress the residual H2O signal with low power selective irradiation at the H2O frequency at μ 4.869 (2915.9 Hz) by 60.59 dB during the recycle delay. Free Induction Decays (FIDs) were Fourier transformed with LB = 0.3 Hz, and the spectra were zero filled to 32 K points. The resulting spectra were manually phased, baseline corrected and calibrated to TMSP at 0.0 ppm, using Topspin (version 2.1, Bruker).
The 1H NMR spectra were automatically reduced to an ASCII file. Spectral intensities were scaled to the internal standard (TSP) area and reduced to integrated regions of equal width (0.04 ppm) corresponding to the region of δ 0.3–δ 10.0. The regions of δ 4.75–δ 4.90 and δ 3.28–δ 3.34 were excluded from the analysis, because of the residual signals of HDO and CD3OD, respectively. Bucketing was performed by AMIX software (Bruker). Principal component analysis (PCA) was performed with the SIMCA-P software (v. 12.0, Umetrics, Umea, Sweden) with scaling based on the Pareto method.

3.4. QTL Analysis

The genetic map was constructed using JoinMap 4.0 [45,61]. MAPQTL6.0 [61] was used for QTL analysis. First, the interval mapping procedure was performed to detect major QTL. For each trait, a 1000× permutation test was performed to calculate the LOD threshold corresponding to a genome-wide false discovery rate of 5% (p < 0.05). Markers with LOD scores equal to or exceeding the threshold were used as cofactors in multiple-QTL-model (MQM) mapping. If new QTLs were detected, the linked markers were added to the co-factor list, and the MQM analysis was repeated. If the LOD value of a marker dropped below the threshold in the new model, it was removed from the cofactor list, and the MQM analysis was rerun. This procedure was repeated, until the cofactor list became stable. The final LOD score for each trait was determined by restricted MQM (rMQM) mapping. In some cases, rMQM mapping showed that some cofactors should be on the same linkage group, but at slightly different positions. In that case, the new marker was selected as a cofactor and the whole procedure was repeated.

4. Conclusions

The detected genotypic variation in tocopherol seed concentration and seedling metabolites in the RIL population under study allowed the detection of several QTLs for these compounds. The loci we detected can be used to establish diagnostic markers for marker-assisted selection for improved nutritional quality (mainly tocopherol and glucosinolate concentrations). The further analysis of these QTLs affecting metabolic processes will increase our knowledge about the regulatory control of biosynthetic pathways.

Acknowledgements

This work was financially supported by a personal grant to Hedayat Bagheri from the Ministry of Science, Research and Technology of Iran and by the IOP Genomics project IGE050010 on Brassica Vegetable Nutrigenomics. We thank Corrie Hanhart, Jens Hermann, Inge van Oorschot and Joost Keurentjes for assistance with the trait analyses and Maarten Koornneef for critical reading of the manuscript.

Conflict of Interest

The authors declare no conflict of interest.

References

  1. Pichersky, E.; Gang, D.R. Genetics and biochemistry of secondary meatbolites in plants: An evolutionary perspective. Trends Plant Sci 2000, 5, 439–445. [Google Scholar]
  2. Wink, M. Plant breeding: Importance of plant secondary metabolites for protection against pathogens and herbivores. Theor. Appl. Genet 1988, 75, 225–233. [Google Scholar]
  3. De Luca, V.; St-Pierre, B. The cell and developmental biology of alkaloid biosynthesis. Trends Plant Sci 2000, 5, 168–173. [Google Scholar]
  4. Gilliland, L.U.; Magallanes-Lundback, M.; Hemming, C.; Supplee, A.; Koornneef, M.; Bentsink, L.; Dellapenna, D. Genetic basis for natural variation in seed vitamin E levels in Arabidopsis thaliana. Proc. Natl. Acad. Sci 2006, 103, 18834–18841. [Google Scholar]
  5. Van Eenennaam, A.L.; Lincoln, K.; Durrett, T.P.; Valentin, H.E.; Shewmaker, C.K.; Thorne, G.M.; Jiang, J.; Baszis, S.R.; Levering, C.K.; Aasen, E.D.; et al. Engineering vitamin E content: From Arabidopsis mutant to soy oil. Plant Cell 2003, 15, 3007–3019. [Google Scholar]
  6. Porfirova, S.; Bergmüller, E.; Tropf, S.; Lemke, R.; Dörmann, P. Isolation of an Arabidopsis mutant lacking vitamin E and identification of a cyclase essential for all tocopherol biosynthesis. Proc. Natl. Acad. Sci 2002, 99, 12495–12500. [Google Scholar]
  7. DellaPenna, D.; Pogson, B. Vitamin synthesis in plants: Toco-pherols and carotenoids. Annu. Rev. Plant Biol. 2006, 57, 711–738. [Google Scholar]
  8. 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]
  9. Thies, W. Entwicklung von Ausgangsmaterial mit erhöhten α- oder γ Tocopherolgehalten im Samenöl für die Körnerrapszüchtung. I. Quantitative Bestimmung der Tocopherole durch HPLC (Development of starting material with enhanced alpha- or gamma-tocopherol contents in the seed oil for the breeding of new rapeseed cultivars). Angew. Bot 1997, 71, 62–67. [Google Scholar]
  10. Goffman, F.D.; Becker, H.C. Diallel analysis for tocopherol contents in seeds of rapeseed. Crop Sci 2001, 41, 1072–1079. [Google Scholar]
  11. Goffman, F.D.; Becker, H.C. Phenotypic variability for tocopherol content and composition in seeds of winter rapeseed (Brassica napus. L) (in German). Vortr. Pflanzenzüchtung 1998, 42, 105–106. [Google Scholar]
  12. Schultz, G. Biosynthesis of α-tocopherol in chloroplasts of higher plants. Fett/Lipid 1990, 92, 86–90. [Google Scholar]
  13. Soll, J.; Kemmerling, M.; Schultz, G. Tocopherol and plastoquinone synthesis in spinach chloroplasts subfractions. Arch. Biochem. Biophys 1980, 204, 544–550. [Google Scholar] [Green Version]
  14. Bergmüller, E.; Porfirova, S.; Dörmann, P. Characterization of an Arabidopsis mutant deficient in γ-tocopherolmethyltransferase. Plant Mol. Biol 2003, 52, 1181–1190. [Google Scholar]
  15. Marwede, V.; Gul, M.K.; Becker, H.C.; Ecke, W. Mapping of QTL controlling tocopherol content in winter oilseed rape. Plant Breed 2005, 124, 20–26. [Google Scholar]
  16. Wang, X.; Zhang, C.; Li, L.; Fritsche, S.; Endrigkeit, J.; Zhang, W.; Long, Y.; Jung, C.; Meng, J. Unraveling the genetic basis of seed tocopherol content and composition in rapeseed (Brassica napus L.). PLoS One 2012, 7, e50038. [Google Scholar]
  17. Norris, S.R.; Shen, X.; DellaPenna, D. Complementation of the Arabidopsis pds1 mutation with the gene encoding p-hydroxyphenylpyruvate dioxygenase. Plant Physiol 1998, 117, 1317–1323. [Google Scholar]
  18. Collakova, E.; DellaPenna, D. Isolation and functional analysis of homogentisate phytyltransferase from Synechocystis sp. PCC 6803 and Arabidopsis. Plant Physiol 2001, 127, 1113–1124. [Google Scholar]
  19. Cheng, Z.; Sattler, S.; Maeda, H.; Sakuragi, Y.; Bryant, D.A.; DellaPenna, D. Highly divergent methyltransferases catalyze a conserved reaction in tocopherol and plastoquinone synthesis in cyanobacteria and photosynthetic eukaryotes. Plant Cell 2003, 15, 2343–2356. [Google Scholar]
  20. Motohashi, R.; Ito, T.; Kobayashi, M.; Taji, T.; Nagata, N.; Asami, T.; Yoshida, S.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Functional analysis of the 37 kDa inner envelope membrane polypeptide in chloroplast biogenesis using a Ds-tagged Arabidopsis pale-green mutant. Plant J 2003, 34, 719–731. [Google Scholar]
  21. Mène-Saffrané, L.; DellaPenna, D. Biosynthesis, regulation and functions of tocochromanols in plants. Plant Physiol. Biochem 2010, 48, 301–309. [Google Scholar]
  22. Cela, J.; Chang, C.; Munné-Bosch, S. Accumulation of γ- rather than α-tocopherol alters thylene signaling gene expression in the vte4 mutant of Arabidopsis thaliana. Plant Cell Physiol 2011, 52, 1389–1400. [Google Scholar]
  23. 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]
  24. Zhang, C.; Cahoon, R.E.; Hunter, S.C.; Chen, M.; Han, J.; Cahoon, E.B. Genetic and biochemical basis for alternative routes of tocotrienol biosynthesis for enhanced vitamin E antioxidant production. Plant J 2013, 73, 628–639. [Google Scholar]
  25. Martinis, J.; Glauser, G.; Valimareanu, S.; Kessler, F. A chloroplast ABC1-like kinase regulates Vitamin E metabolism in Arabidopsis thaliana. Plant Physiol 2013. doi:10,1104/pp.113218644. [Google Scholar]
  26. 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]
  27. Kohlmeier, L.; Su, L. Cruciferous vegetable consumption and colorectal cancer risk: Meta-analysis of the epidemiological evidence. FASEB J 1997, 11, 369. [Google Scholar]
  28. Hayes, J.; Kelleher, M.; Eggleston, I. The cancer chemopreventive actions of phytochemicals derived from glucosinolates. Eur. J. Nutr 2008, 47, 73–88. [Google Scholar]
  29. Fenwick, G.R.; Heaney, R.K. Glucosinolates and their breakdown products in cruciferous crops, food, and feeding stuffs. Food Chem 1983, 11, 249–271. [Google Scholar]
  30. Fahey, J.W.; Zalcmann, A.M.; Talalay, P. The chemical diversity and distribution of glucosinolates and isothiocyanates among plants. Phytochemistry 2001, 56, 5–61. [Google Scholar]
  31. Ciska, E.; Martyniak-Przybyszewska, B.; Kozlowska, H. Content of glucosinolates in cruciferous vegetables grown at the same site for two years under different climatic conditions. J. Agric. Food Chem 2000, 48, 2862–2867. [Google Scholar]
  32. Giamoustaris, A.; Mithen, R. Genetics of aliphatic glucosinolates. IV. Side-chain modification in Brassica oleracea. Theor. Appl. Genet 1996, 93, 1006–1010. [Google Scholar]
  33. Sønderby, I.E.; Geu-Flores, F.; Halkier, B.A. Biosynthesis of glucosinolate gene discovery and beyond. Trends Plant Sci 2010, 15, 283–296. [Google Scholar]
  34. Mithen, R.; Faulkner, K.; Magrath, R.; Rose, P.; Williamson, G.; Marquez, J. Development of isothiocyanate-enriched broccoli and its enhanced ability to induce phase 2 detoxification enzymes in mammalian cells. Theor. Appl. Genet 2003, 106, 727–734. [Google Scholar]
  35. Cartea, M.E.; Velasco, P. Glucosinolates in Brassica foods: Bioavailability in food and significance for human health. Phytochem. Rev 2008, 7, 213–229. [Google Scholar]
  36. Lee, J.G.; Bonnema, G.; Zhang, N.; Kwak, J.H.; de Vos, R.C.; Beekwilder, J. Evaluation of glucosinolate variation in a collection of turnip (Brassica rapa) germplasm by the analysis of intact and desulfo glucosinolates. J. Agric. Food Chem 2013, 61, 3984–3993. [Google Scholar]
  37. He, H.; Fingerling, G.; Schnitzler, W.H. Glucosinolate contents and patterns in different organs of Chinese cabbages, Chinese kale (Brassica alboglabra bailey) and Choy sum (Brassica campestris L. ssp. chinensis var. Utilis). J. Appl. Bot 2000, 74, 21–25. [Google Scholar]
  38. Padilla, G.; Cartea, M.E.; Velasco, P.; de Haro, A.; Ordas, A. Variation of glucosinolates in vegetable crops of Brassica rapa. Phytochemistry 2007, 68, 536–545. [Google Scholar]
  39. Lou, P.; Zhao, J.; He, H.; Hanhart, C.; Del Carpio, D.P.; Verkerk, R.; Custers, J.; Koornneef, M.; Bonnema, G. Quantitative trait loci for glucosinolate accumulation in Brassica rapa leaves. New Phytol 2008, 179, 1017–1032. [Google Scholar]
  40. Bo, Y.; Quiros, C.F. Survey of glucosinolate variation in leaves of Brassica rapa crops. Genet. Resour. Crop Evol 2010, 57, 1079–1089. [Google Scholar]
  41. Wang, H.; Sun, S.; Liu, B.; Cheng, F.; Sun, R.; Wang, X. Glucosinolate biosynthetic genes in Brassica rapa. Gene 2011, 487, 135–142. [Google Scholar]
  42. Choi, Y.H.; Kim, H.K.; Linthorst, H.J.M.; Hollander, J.G.; Lefeber, A.W.M.; Erkelens, C.; Nuzillard, J.M.; Verpoorte, R. NMR Metabolomics to revisit the Tobacco Mosaic virus Infection in Nicotiana tabacum Leaves. J. Nat. Prod 2006, 69, 742–748. [Google Scholar]
  43. Moco, S.; Bino, R.J.; de Vos, R.C.H.; Vervoort, J. Metabolomics technologies and metabolite identification. Trac-Trend Anal. Chem 2007, 26, 855–866. [Google Scholar]
  44. Abdel-Farid, I.B.; Kim, H.K.; Choi, Y.H.; Verpoorte, R. Metabolic characterization of Brassica rapa leaves by NMR spectroscopy. J. Agric. Food Chem 2007, 55, 7936–7943. [Google Scholar]
  45. Bagheri, H.; El-Soda, M.; van Oorschot, I.; Hanhart, C.; Bonnema, G.; Jansen-van den Bosch, T.; Mank, R.; Keurentjes, J.J.B.; Mang, L.; Wu, J.; Koornneef, M.; Aarts, M.G.M. Genetic analysis of morphological traits in a new, versatile, rapid-cycling Brassica rapa recombinant inbred line population. Front. Plant Sci 2012, 3, 183. [Google Scholar]
  46. Wang, X.; Wang, H.; Wang, J.; Sun, R.; Wu, J.; Liu, S.; Bai, Y.; Mun, J-H.; Bancroft, I.; Cheng, F.; et al. The genome of the mesopolyploid crop species Brassica rapa. Nat. Genet. 2011, 43, 1035–1039. [Google Scholar]
  47. Liu, L.; Stein, A.; Wittkop, B.; Sarvari, P.; Li, J.; Yan, X.; Dreyer, F.; Frauen, M.; Friedt, W.; Snowdon, R.J. A knockout mutation in the lignin biosynthesis gene CCR1 explains a major QTL for acid detergent lignin content in Brassica napus seeds. Theor. Appl. Genet 2012, 124, 1573–1586. [Google Scholar]
  48. Kliebenstein, D.J.; Gershenzon, J.; Mitchell-Olds, T. Comparative quantitative trait loci mapping of aliphatic, indolic and benzylic glucosinolate production in Arabidopsis thaliana leaves and seeds. Genetics 2001, 159, 359–370. [Google Scholar]
  49. Kroymann, J.; Textor, S.; Tokuhisa, J.G.; Falk, K.L.; Bartram, S.; Gershenzon, J.; Mitchell-Olds, T. A gene controlling variation in Arabidopsis glucosinolate composition is part of the methionine chain elongation pathway. Plant Physiol 2001, 127, 1077–88. [Google Scholar]
  50. Field, B.; Cardon, G.; Traka, M.; Botterman, J.; Vancanneyt, G.; Mithen, R. Glucosinolate and amino acid biosynthesis in Arabidopsis. Plant Physiol 2004, 135, 828–839. [Google Scholar]
  51. Heidel, A.J.; Clauss, M.J.; Kroymann, J.; Savolainen, O.; Mitchell-Olds, T. Natural variation in MAM within and between populations of Arabidopsis lyrata determines glucosinolate phenotype. Genetics 2006, 173, 1629–1636. [Google Scholar]
  52. Textor, S.; de Kraker, J.W.; Hause, B.; Gershenzon, J.; Tokuhisa, J.G. MAM3 catalyzes the formation of all aliphatic glucosinolate chain lengths in Arabidopsis. Plant Physiol 2007, 144, 60–71. [Google Scholar]
  53. Hirai, M.Y.; Sugiyama, K.; Sawada, Y.; Tohge, T.; Obayashi, T.; Suzuki, A.; Araki, R.; Sakurai, N.; Suzuki, H.; Aoki, K.; et al. Omics-based identification of Arabidopsis Myb transcription factors regulating aliphatic glucosinolate biosynthesis. Proc. Natl. Acad. Sci 2007, 104, 6478–6483. [Google Scholar]
  54. Samach, A.; Harevent, D.; Gutfinger, T.; Ken-Dror, S.; Lifschitz, E. Biosynthetic threonine deaminase gene of tomato: Isolation, structure, and upregulation in floral organs. Proc. Natl. Acad. Sci 1991, 88, 2678–2682. [Google Scholar]
  55. Hofgen, R.; Streber, W.; Pohlenz, H. Antisense gene expression as a tool for evaluating molecular herbicide targets. Pest Sci 1995, 43, 175–177. [Google Scholar]
  56. Cheng, F.; Wu, J.; Fang, L.; Wang, X. Syntenic gene analysis between Brassica rapa and other Brassicaceae Species. Front. Plant Sci 2012, 3, 198. [Google Scholar]
  57. Forde, B.G.; Lea, P.J. Glutamate in plants: Metabolism, regulation, and signalling. J. Exp. Bot 2007, 58, 2339–2358. [Google Scholar]
  58. Schledz, M.; Seidler, A.; Beyer, P.; Neuhaus, G. A novel phytyltransferase from Synechocystis sp. PCC 6803 involved in tocopherol biosynthesis. FEBS. Lett 2001, 499, 15–20. [Google Scholar]
  59. Dähnhardt, D.; Falk, J.; Appel, J.; van der Kooij, T.A.; Schulz-Friedrich, R.; Krupinska, K. The hydroxyphenylpyruvate dioxygenase from Synechocystis sp. PCC 6803 is not required for plastoquinone biosynthesis. FEBS Lett 2002, 523, 177–181. [Google Scholar]
  60. Falk, J.; Andersen, G.; Kernebeck, B.; Krupinska, K. Constitutive overexpression of barley 4-hydroxyphenylpyruvate dioxygenase in tobacco results in elevation of the vitamin E content in seeds but not in leaves. FEBS Lett 2003, 540, 35–40. [Google Scholar]
  61. Kyazma-Genetic linkage analysis software. Available online: http://www.kyazma.nl (accessed on 19 April 2012).
Figure 1. Frequency distributions of non-normalized data of tocopherols in seeds of the L58 × R-o-18 RIL population, including 160 genotypes. The vertical axis indicates the number of lines per trait value class and the horizontal axis the different trait value classes. From left to right and top to the bottom: α-tocopherol (mg/g); β-tocopherol (mg/g); γ-tocopherol (mg/g); δ-tocopherol (mg/g); total tocopherol (mg/g); α/γ tocopherol ratio. The parental values are the mean of three replicates, indicated with L for L58 and R for R-o-18.
Figure 1. Frequency distributions of non-normalized data of tocopherols in seeds of the L58 × R-o-18 RIL population, including 160 genotypes. The vertical axis indicates the number of lines per trait value class and the horizontal axis the different trait value classes. From left to right and top to the bottom: α-tocopherol (mg/g); β-tocopherol (mg/g); γ-tocopherol (mg/g); δ-tocopherol (mg/g); total tocopherol (mg/g); α/γ tocopherol ratio. The parental values are the mean of three replicates, indicated with L for L58 and R for R-o-18.
Ijms 14 15561f1aIjms 14 15561f1b
Figure 2. Principal component analysis (PCA) analysis of seedling metabolites based on Nuclear Magnetic Resonance (NMR) signals detected in the L58 × R-o-18 RIL population, including 160 genotypes. Numbers in the figure are line numbers of the RIL population. Parental values are indicated with a red triangle.
Figure 2. Principal component analysis (PCA) analysis of seedling metabolites based on Nuclear Magnetic Resonance (NMR) signals detected in the L58 × R-o-18 RIL population, including 160 genotypes. Numbers in the figure are line numbers of the RIL population. Parental values are indicated with a red triangle.
Ijms 14 15561f2
Figure 3. A clustered heat map showing the LOD profiles of QTL identified for the indicated metabolite signals. Columns indicate the 10 B. rapa chromosomes, in centimorgans, ascending from the left to right; rows indicate individual compound signal LOD profiles. A color scale is used to indicate the QTL significance corresponding to the LOD score. Positive values (red and black) represent a positive effect on the compound concentration by the L58 allele; negative values (blue and green) represent a positive effect by the R-o-18 allele. The width of a bar indicates the significance interval of the QTL that was calculated by restricted multiple-QTL-model (rMQM) in MAPQTL6. Hierarchical clustering, shown on the left, reflects the correlation between compound concentrations, based on the QTL profiles.
Figure 3. A clustered heat map showing the LOD profiles of QTL identified for the indicated metabolite signals. Columns indicate the 10 B. rapa chromosomes, in centimorgans, ascending from the left to right; rows indicate individual compound signal LOD profiles. A color scale is used to indicate the QTL significance corresponding to the LOD score. Positive values (red and black) represent a positive effect on the compound concentration by the L58 allele; negative values (blue and green) represent a positive effect by the R-o-18 allele. The width of a bar indicates the significance interval of the QTL that was calculated by restricted multiple-QTL-model (rMQM) in MAPQTL6. Hierarchical clustering, shown on the left, reflects the correlation between compound concentrations, based on the QTL profiles.
Ijms 14 15561f3
Table 1. Analysis of tocopherol concentrations in both parents and the L58 × R-o-18 Recombinant Inbred Line (RIL) population, including 160 genotypes. Tocopherol concentration is given in mg per g of seed.
Table 1. Analysis of tocopherol concentrations in both parents and the L58 × R-o-18 Recombinant Inbred Line (RIL) population, including 160 genotypes. Tocopherol concentration is given in mg per g of seed.
Tocopherolα-β-γ-δ-totalα-/γ-ratio
L58 (Mean value ± SE)0.317 ± 0.0060.0003 ± 0.00010.273 ± 0.0030.004 ± 0.00040.595 ± 0.0101.159 ± 0.008
R-o-18 (Mean value ± SE)0.250 ± 0.0010.0006 ± 0.000080.233 ± 0.0010.019 ± 0.00030.502 ± 0.0021.076 ± 0.003
Max value0.8850.00190.5300.0341.1998.266
Min value0.1610.00010.0860.0030.3250.537
Mean value0.3650.00080.2940.0150.6751.428
Table 2. Pearson correlation analysis of tocopherols in the L58 × R-o-18 RIL population, including 160 genotypes.
Table 2. Pearson correlation analysis of tocopherols in the L58 × R-o-18 RIL population, including 160 genotypes.
Traitα-β-γ-δ-α-/γ-
β-0.134
γ-−0.09−0.22 *
δ-−0.160.1060.15
α-/γ-0.69 **0.253 *−0.65 **−0.23 *
Total0.78 **−0.0240.54 **−0.010.19
α-: α-tocopherol; γ-: γ-tocopherol; δ-: δ-tocopherol; Total: total tocopherol; α-/γ-: α-/γ-tocopherol ratio.
**means significant at p ≤ 0.01;
*significant at p ≤ 0.05.
Table 3. Quantitative trait loci (QTL) related to tocopherol concentration in seeds of the B. rapa L58 × R-o-18 RIL population, including 160 genotypes. “Peak position” indicates the location of the highest LOD score for each QTL. Flanking markers shows marker names flanking the QTL confidence interval based on a one LOD interval. “% Expl. var.” is the percentage of total phenotypic variance explained by individual QTLs. The allelic effect of each QTL is indicated (effect), which is calculated as μA-μB (μ = mean), where A and B are RILs carrying L58 and R-o-18, respectively, alleles at the relevant QTL position. Effects are given in mg/g or without unit (for the ratio of α/γ tocopherol). H2 is broad sense heritability. For all traits, four replicate samples were measured. For β- and δ-tocopherol, values were very small to calculate the difference.
Table 3. Quantitative trait loci (QTL) related to tocopherol concentration in seeds of the B. rapa L58 × R-o-18 RIL population, including 160 genotypes. “Peak position” indicates the location of the highest LOD score for each QTL. Flanking markers shows marker names flanking the QTL confidence interval based on a one LOD interval. “% Expl. var.” is the percentage of total phenotypic variance explained by individual QTLs. The allelic effect of each QTL is indicated (effect), which is calculated as μA-μB (μ = mean), where A and B are RILs carrying L58 and R-o-18, respectively, alleles at the relevant QTL position. Effects are given in mg/g or without unit (for the ratio of α/γ tocopherol). H2 is broad sense heritability. For all traits, four replicate samples were measured. For β- and δ-tocopherol, values were very small to calculate the difference.
TraitQTLLinkage groupLODPeak position (cM)Flanking markersConfidence interval (cM)% Expl. var.EffectH2
α-tocopherolAl1A910.254.4899051|9912525, E3850M151–56.534.5−0.150.89
Al2A63.866.9902204|9940666, 905950|989164961–69.610.7+0.08
β-tocopherolBeA102.680.0E3851M15, BrID1158166.8–8812.1+0.0004-
γ-tocopherolGa1A64.028.4905326|9911105, 899208|996119524–36.514.3+0.080.86
Ga2A103.537.3E3416M30, E3851M126–4812.3−0.06
δ-tocopherolDe1A33.016.0E3835M1, VtE13.5–1912−0.004-
De2A93.063.4904266|9904851, 902257|995564463–65.611.5+0.004
Total tocopherolToc1A68.228.4905326|9911105, 899208|996119524–3627.4+0.160.90
Toc2A94.956.6E3732M3, 899475|995288355–58.815.1−0.12
α/γ tocopherol ratioALGaA95.254.4899051|9912525, E3850M151–55.522.3−0.930.89
Table 4. Pearson correlation analysis of metabolites identified by 1H-NMR signals in the L58 × R-o-18 B. rapa RIL population, including 160 genotypes. Only significant correlations (at p ≤ 0.05) with scores ≥ 0.45 are shown.
Table 4. Pearson correlation analysis of metabolites identified by 1H-NMR signals in the L58 × R-o-18 B. rapa RIL population, including 160 genotypes. Only significant correlations (at p ≤ 0.05) with scores ≥ 0.45 are shown.
Compound signalsFlavonoid3NeoglucobrassicinPhenylpropanoidGlucose alphaGlucose betaProgoitrinCholineAsparagineCitrullineMalateValineThreonineAlanineIsoleucineGlutamine
Neoglucobrassicin0.53
Phenylpropanoid--
Sinapoyl glucose--0.61
Glucose alpha---
Glucose beta---0.62
Progoitrin---0.58-
Choline---0.500.550.68
Asparagine-----0.95-
Citrulline-------0.56
Malate-0.490.66--0.55--0.45
Valine----65.00.540.54---
Threonine---0.56-0.46----0.48
Alanine-----0.450.54-0.52-0.800.47
Isoleucine--------0.60-0.98-0.73
Glutamine---0.540.620.580.55-0.45-0.820.480.750.73
Glutamate---0.680.590.570.57--0.570.840.670.740.690.83
Table 5. Overview of QTL related to seedling metabolites based on NMR signals detected in the B. rapa L58 × R-o-18 RIL population, including 160 genotypes. The peak position indicates the genetic map position of the highest LOD score for each QTL. Flanking markers show marker names flanking the QTL confidence interval based on a one LOD interval. % Expl. var. is the percentage of total phenotypic variance explained by individual QTLs.
Table 5. Overview of QTL related to seedling metabolites based on NMR signals detected in the B. rapa L58 × R-o-18 RIL population, including 160 genotypes. The peak position indicates the genetic map position of the highest LOD score for each QTL. Flanking markers show marker names flanking the QTL confidence interval based on a one LOD interval. % Expl. var. is the percentage of total phenotypic variance explained by individual QTLs.
Compound signalLinkage groupLODPeak position (cM)Flanking markersConfidence interval (cM)% Expl. Var.
ProgoitrinA34.624E3556M6, P2348M16020.8–30.617
A31895E3851M18, E3850M290–102.539

PhenylpropanoidsA5658P2348M220, 903986|991457753–61.512
A2320BrID101239-A5, E3850M916.5–299
A54.669E3732M6, 898692|994526364.8–7110
A64.8101899475|9952883, P2348M20092–11215
A716.5112E3416M2, 905396|9906565110–12543
A92.88899475|9952883, P2348M2003–13.46

NeoglucobrassicinA96.858BRMS-018-A7, 904702|990241555.5–6015
A102.881E3416M2, 905396|990656567–866
A1325E3851M15, BrID1158122–36.59
A12.8138E4051M5, BrID11087133–140.54
A33.316E3835M7, 905044|990129112–16.66
A38.644E3835M1, VtE39–4724
A54.318904714|9903794, BrID101239-A514.6–3114

AlanineA53.871P2348M112, 905015|991588168–7910
A63.472E3851M17, E3416M1370.6–758.5
A92418905396|9906565, E3850M413–2250
A33.6139P2348M294, E4051M3137–147.512
A74.132BrID10119-A7, P2348M18930–34.813.5

AsparagineA314.592E3851M6, E3851M1887–94.737.3
A43.247.7E3851M9, E3856M334–556.5
A54.358P2348M220, 903986|991457756–60.810
A10348904135|9928717, P2348M31737.4–497

GlutamineA14.483P2348M66, E3416M878–868.7
A39.419VtE, P2147M28516.6–2123
A37.243E3835M6, 899062|991154841.3–4719
A4455E3850M10, E3856M647–58.312.7
A53134.5BrID90357, 902420|9939979130–1385.5
A64.1101E3732M6, 898692|994526392–1138.3
A77.635901866|9957577, BrID10107-A732–4716
A94.3107904535|9904729, 903426|991939784–1128.5

IsoleucineA35.322P2147M276, E3416M1419–2515.5
A45.634.2E3851M10, E3850M1030–4816.5

ThreonineA36121902392|9921642, E3835M10111–12519.3
A43.234.2E3851M10, E3850M1031–47.610

ValineA36.422.5VtE, P2147M28519–2515.3
A48.425E3851M10, E3850M1013–3120.3
A73.835901866|9957577, BrID10107-A731.5–46.48.6
A93.30E3732M12, E3416M20–5.57.4
A9484BrID10187, 904535|990472981.5–1118.5
A10456P2348M317, 900140|993864352–599

GlutamateA1583P2348M66, E3416M878–91.511.2
A3943E3416M14, E3835M640.6–4727
A310110903854|9907725, P2348M294106–12030
A73.633.5BrID10119-A7, P2348M18930–44.610

MalateA64.148.5899015|9918455, 902942|991676643–5210
A711.5109BRMS-018-A7, 904702|9902415105–112.633
A103.783BrID11581, E3835M478–88.59

Glucose (alpha, beta)A35.643E3835M6, 899062|991154838–46.420
A33.495E3851M18, E3850M289–10313

Share and Cite

MDPI and ACS Style

Bagheri, H.; El-Soda, M.; Kim, H.K.; Fritsche, S.; Jung, C.; Aarts, M.G.M. Genetic Analysis of Health-Related Secondary Metabolites in a Brassica rapa Recombinant Inbred Line Population. Int. J. Mol. Sci. 2013, 14, 15561-15577. https://doi.org/10.3390/ijms140815561

AMA Style

Bagheri H, El-Soda M, Kim HK, Fritsche S, Jung C, Aarts MGM. Genetic Analysis of Health-Related Secondary Metabolites in a Brassica rapa Recombinant Inbred Line Population. International Journal of Molecular Sciences. 2013; 14(8):15561-15577. https://doi.org/10.3390/ijms140815561

Chicago/Turabian Style

Bagheri, Hedayat, Mohamed El-Soda, Hye Kyong Kim, Steffi Fritsche, Christian Jung, and Mark G. M. Aarts. 2013. "Genetic Analysis of Health-Related Secondary Metabolites in a Brassica rapa Recombinant Inbred Line Population" International Journal of Molecular Sciences 14, no. 8: 15561-15577. https://doi.org/10.3390/ijms140815561

Article Metrics

Back to TopTop