[go: up one dir, main page]

 
 
ijms-logo

Journal Browser

Journal Browser

Biosynthesis and Regulatory Mechanism of Secondary Metabolites in Medicinal Plants

A special issue of International Journal of Molecular Sciences (ISSN 1422-0067). This special issue belongs to the section "Molecular Plant Sciences".

Deadline for manuscript submissions: closed (31 December 2023) | Viewed by 24129

Special Issue Editor


E-Mail Website
Guest Editor
Institute of Medicinal Plant Development, Chinese Academy of Medical Sciences & Peking Union Medical College, Beijing, China
Interests: medicinal plant; bioactive compound; biosynthetic pathway; noncoding RNA; transcription factor; genome; transcriptome; metabolome
Special Issues, Collections and Topics in MDPI journals

Special Issue Information

Dear Colleagues,

Medicinal plants are an important resource for humans. However, compared with model systems and crops, the number of studies on medicinal plants has fallen far behind the amount of research on other topics. Recently, with the increase in demand for medicinal plants and the development and application of high-throughput technologies, the research field of medicinal plants has rapidly expanded. Significant progress has been made in genomics, epigenomics, transcriptomics and metabolomics of medicinal plants. Numerous studies have contributed to the biosynthetic pathway of secondary metabolites, genes encoding key enzymes of the pathway, and regulatory mechanisms of secondary metabolism. This enables the production of secondary metabolites through metabolic engineering and synthetic biology. Moreover, novel technologies and strategies are developing and applying to this research field. This open-access Special Issue of IJMS is devoted to publishing original research and review articles on medicinal plant studies, highlighting recent advances in the biosynthesis and regulatory mechanisms of secondary metabolites, particularly significant discoveries from intensive studies, and the development and application of novel technologies. This issue aims to provide an accessible collection of research that shares the latest innovative results from the research field of medicinal plants to aid further studies on secondary metabolism, medicinal plant improvement, and the production of functionally important secondary metabolites.

Topics of this Special Issue include, but are not limited to:

  • Genomics, epigenomics, transcriptomics and metabolomics of medicinal plants;
  • Biosynthetic pathway of secondary metabolites;
  • Key enzyme genes involved in the biosynthesis of secondary metabolites;
  • Epigenetic regulation of secondary metabolism: microRNA, long noncoding RNA, DNA methylation, RNA methylation, etc.;
  • Transcription factor and regulatory network in medicinal plants;
  • Metabolic engineering and synthetic biology of secondary metabolites;
  • Application of high-throughput sequencing technologies;
  • Databases associated with the biosynthesis and regulation of secondary metabolites;
  • Novel technologies and strategies for secondary metabolism studies.

Prof. Dr. Shanfa Lu
Guest Editor

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. International Journal of Molecular Sciences is an international peer-reviewed open access semimonthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. There is an Article Processing Charge (APC) for publication in this open access journal. For details about the APC please see here. Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • biosynthetic pathway
  • epigenetic regulation
  • high-throughput sequencing
  • medicinal plant
  • metabolic engineering
  • omics
  • regulatory mechanism
  • secondary metabolite
  • synthetic biology
  • transcription factor

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Related Special Issue

Published Papers (13 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Research

15 pages, 10118 KiB  
Article
Overexpression of AtMYB2 Promotes Tolerance to Salt Stress and Accumulations of Tanshinones and Phenolic Acid in Salvia miltiorrhiza
by Tianyu Li, Shuangshuang Zhang, Yidan Li, Lipeng Zhang, Wenqin Song and Chengbin Chen
Int. J. Mol. Sci. 2024, 25(7), 4111; https://doi.org/10.3390/ijms25074111 - 8 Apr 2024
Cited by 1 | Viewed by 1091
Abstract
Salvia miltiorrhiza is a prized traditional Chinese medicinal plant species. Its red storage roots are primarily used for the treatment of cardiovascular and cerebrovascular diseases. In this study, a transcription factor gene AtMYB2 was cloned and introduced into Salvia miltiorrhiza for ectopic expression. [...] Read more.
Salvia miltiorrhiza is a prized traditional Chinese medicinal plant species. Its red storage roots are primarily used for the treatment of cardiovascular and cerebrovascular diseases. In this study, a transcription factor gene AtMYB2 was cloned and introduced into Salvia miltiorrhiza for ectopic expression. Overexpression of AtMYB2 enhanced salt stress resistance in S. miltiorrhiza, leading to a more resilient phenotype in transgenic plants exposed to high-salinity conditions. Physiological experiments have revealed that overexpression of AtMYB2 can decrease the accumulation of reactive oxygen species (ROS) during salt stress, boost the activity of antioxidant enzymes, and mitigate oxidative damage to cell membranes. In addition, overexpression of AtMYB2 promotes the synthesis of tanshinones and phenolic acids by upregulating the expression of biosynthetic pathway genes, resulting in increased levels of these secondary metabolites. In summary, our findings demonstrate that AtMYB2 not only enhances plant tolerance to salt stress, but also increases the accumulation of secondary metabolites in S. miltiorrhiza. Our study lays a solid foundation for uncovering the molecular mechanisms governed by AtMYB2 and holds significant implications for the molecular breeding of high-quality S. miltiorrhiza varieties. Full article
Show Figures

Figure 1

Figure 1
<p>Expression pattern of <span class="html-italic">AtMYB2</span> in <span class="html-italic">S. miltiorrhiza</span> transgenic hairy roots (<b>A</b>) and <span class="html-italic">S. miltiorrhiza</span> transgenic plants (<b>B</b>). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Morphological changes under salt stress. Scale bar = 5 cm; CK, control group; DAT, days after treatment; WT, wild type; pCAMBIA1301, empty vector control pCAMBIA1301 line; <span class="html-italic">OE-AtMYB2-L2/L5/L7</span>, <span class="html-italic">AtMYB2</span> transgenic lines 2/5/7.</p>
Full article ">Figure 3
<p>Accumulation of O<sup>2−</sup> (<b>A</b>,<b>C</b>) and H<sub>2</sub>O<sub>2</sub> (<b>B</b>,<b>D</b>) in all <span class="html-italic">S. miltiorrhiza</span> lines under salt stress. The localization and accumulation of O<sup>2−</sup> (<b>A</b>) and H<sub>2</sub>O<sub>2</sub> (<b>B</b>) were visualized by blue and brown staining. Control, plants grown under normal conditions; Salt, plants subjected to salt stress treatment for 7 days. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p><span class="html-italic">AtMYB2</span> improves antioxidant capacity under salt stress. Content of EL (<b>A</b>), total chlorophyll (<b>B</b>), and MDA (<b>C</b>) on 7 DAT; activities of SOD (<b>D</b>), POD (<b>E</b>), and CAT (<b>F</b>) in all <span class="html-italic">S. miltiorrhiza</span> lines on 7 DAT. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p><span class="html-italic">AtMYB2</span> promotes the accumulation of tanshinones and salvianolic acid. The phenotype of harvested <span class="html-italic">S. miltiorrhiza</span> hairy roots and <span class="html-italic">S. miltiorrhiza</span> plants, scale bar = 5 cm (<b>A</b>,<b>B</b>); the salvianolic acid B (<b>C</b>) and tanshinones (<b>D</b>) extracts from the <span class="html-italic">AtMYB2</span> transgenic hairy root lines; the salvianolic acid B (<b>E</b>) and tanshinones (<b>F</b>) extracts from the <span class="html-italic">AtMYB2</span> transgenic plant lines; content of secondary metabolites in transgenic hairy roots (<b>G</b>) and plants (<b>H</b>). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Transcription levels of genes involved in the salvianolic acid biosynthesis pathway in transgenic <span class="html-italic">Salvia miltiorrhiza</span> plants. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Transcription levels of genes involved in the tanshinone biosynthesis pathway in transgenic <span class="html-italic">Salvia miltiorrhiza</span> plants. MVA pathway, the mevalonate pathway in the cytosol; MEP pathway, 2-C-methyl-d-erythritol-4-phosphate pathway in the plastids. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Hierarchical clustering of DEGs of <span class="html-italic">AtMYB2</span> transgenic lines versus those of pCAMBIA1301 plants (<b>A</b>) and KEGG classification of DEGs of <span class="html-italic">AtMYB2</span> transgenic lines versus those of pCAMBIA1301 plants (<b>B</b>).</p>
Full article ">
16 pages, 9579 KiB  
Article
Metabolic Pathway Engineering Improves Dendrobine Production in Dendrobium catenatum
by Meili Zhao, Yanchang Zhao, Zhenyu Yang, Feng Ming, Jian Li, Demin Kong, Yu Wang, Peng Chen, Meina Wang and Zhicai Wang
Int. J. Mol. Sci. 2024, 25(1), 397; https://doi.org/10.3390/ijms25010397 - 28 Dec 2023
Cited by 3 | Viewed by 1432
Abstract
The sesquiterpene alkaloid dendrobine, widely recognized as the main active compound and a quality control standard of medicinal orchids in the Chinese Pharmacopoeia, demonstrates diverse biological functions. In this study, we engineered Dendrobium catenatum as a chassis plant for the production of dendrobine [...] Read more.
The sesquiterpene alkaloid dendrobine, widely recognized as the main active compound and a quality control standard of medicinal orchids in the Chinese Pharmacopoeia, demonstrates diverse biological functions. In this study, we engineered Dendrobium catenatum as a chassis plant for the production of dendrobine through the screening and pyramiding of key biosynthesis genes. Initially, previously predicted upstream key genes in the methyl-D-erythritol 4-phosphate (MEP) pathway for dendrobine synthesis, including 4-(Cytidine 5′-Diphospho)-2-C-Methyl-d-Erythritol Kinase (CMK), 1-Deoxy-d-Xylulose 5-Phosphate Reductoisomerase (DXR), 2-C-Methyl-d-Erythritol 4-Phosphate Cytidylyltransferase (MCT), and Strictosidine Synthase 1 (STR1), and a few downstream post-modification genes, including Cytochrome P450 94C1 (CYP94C1), Branched-Chain-Amino-Acid Aminotransferase 2 (BCAT2), and Methyltransferase-like Protein 23 (METTL23), were chosen due to their deduced roles in enhancing dendrobine production. The seven genes (SG) were then stacked and transiently expressed in the leaves of D. catenatum, resulting in a dendrobine yield that was two-fold higher compared to that of the empty vector control (EV). Further, RNA-seq analysis identified Copper Methylamine Oxidase (CMEAO) as a strong candidate with predicted functions in the post-modification processes of alkaloid biosynthesis. Overexpression of CMEAO increased dendrobine content by two-fold. Additionally, co-expression analysis of the differentially expressed genes (DEGs) by weighted gene co-expression network analysis (WGCNA) retrieved one regulatory transcription factor gene MYB61. Overexpression of MYB61 increased dendrobine levels by more than two-fold in D. catenatum. In short, this work provides an efficient strategy and prospective candidates for the genetic engineering of D. catenatum to produce dendrobine, thereby improving its medicinal value. Full article
Show Figures

Figure 1

Figure 1
<p>Multigene reconstruction to enhance dendrobine production. (<b>A</b>) Physical map of the multigene constructs used for integration and expression of the synthetic operons. The selected target genes are depicted as light blue boxes. Promotor <span class="html-italic">Prrn</span> is shown in dark yellow and terminator <span class="html-italic">TrbcL</span> in orange. The <span class="html-italic">HPTII</span> selectable marker gene for transformation is represented as a light yellow box. An intercistronic expression element (<span class="html-italic">IEE</span>) was put between <span class="html-italic">STR1</span> and <span class="html-italic">CYP94C1</span> operons to ensure the downstream cistron expression under the same promotor. (<b>B</b>) <span class="html-italic">Not</span> I-digestion analysis of pYLTAC380H-multigene (<span class="html-italic">EV</span>, <span class="html-italic">TG</span>, <span class="html-italic">FG</span>, and <span class="html-italic">SG</span>) constructs. M: DNA ladder marker. (<b>C</b>) Dendrobine content in <span class="html-italic">D. catenatum</span> leaves infiltrated with <span class="html-italic">Agrobacterium tumefaciens</span> carrying multigene constructs. There are three replicates for each sample. (<b>D</b>) The relative expression of each gene in a specific multigene construct. <span class="html-italic">EV</span>: empty vector; <span class="html-italic">TG</span>: two genes; <span class="html-italic">FG</span>: five genes; <span class="html-italic">SG</span>: seven genes. Asterisks indicate significance based on the Student’s <span class="html-italic">t</span>-test. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001; **** <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 2
<p>Functional verification of <span class="html-italic">MCT</span> in dendrobine synthesis. (<b>A</b>) <span class="html-italic">MCT</span> expression was checked by qRT-PCR with samples being collected 6 h after infiltration (n = 3). (<b>B</b>) Transiently infiltrated <span class="html-italic">D. catenatum</span> leaves were harvested (5 dpi) for dendrobine measurement (n = 3). (<b>C</b>) CRISPRi was performed to knock down <span class="html-italic">MCT</span> expression. The kinase-dead version of Cas9 (dCas9) was used to block transcription in the promotor region of <span class="html-italic">MCT</span>. Empty vector without dCas9 served as the control. <span class="html-italic">MCT</span> knock-down was verified by qRT-PCR, with samples being collected 6 h after infiltration (n = 3). (<b>D</b>) Leave samples were collected at 5 dpi and subjected to dendrobine measurement (n = 3). ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001; **** <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 3
<p>Generation and identification of multigene-transgenic plants. (<b>A</b>) Transgenic <span class="html-italic">D. catenatum</span> plantlets of <span class="html-italic">SG</span>-multigene (11-month-old) and <span class="html-italic">EV</span> control (13-month-old). (<b>B</b>) Analysis of the <span class="html-italic">hptII</span> transgene in <span class="html-italic">SG</span>-multigene and <span class="html-italic">EV</span> control transgenic <span class="html-italic">D. catematum</span> plantlets. DNA isolated from wild-type plants was used as the negative control “−”, while the <span class="html-italic">hptII</span> gene in the <span class="html-italic">EV</span> plasmid was used as the positive control “+”. <span class="html-italic">DcActin</span> was PCR amplified to demonstrate an equal amount of loading. <span class="html-italic">nptII</span> was amplified to avoid bacterial contamination. (<b>C</b>) <span class="html-italic">FPPS</span> expression was checked to demonstrate activation of the dendrobine synthesis pathway. <span class="html-italic">EV</span> transgenic <span class="html-italic">D. catenatum</span> served as the control (Ctrl). (<b>D</b>) <span class="html-italic">SG</span>-transgenic <span class="html-italic">D. catenatum</span> grown in pine-bark pots for 10 months. (<b>E</b>) Molecular characterization of the <span class="html-italic">hptII</span> transgene in <span class="html-italic">SG</span>-transgenic <span class="html-italic">D. catenatum</span> grown in pine-bark pots. (<b>F</b>) Molecular characterization of the <span class="html-italic">hptII</span> transgene in <span class="html-italic">SG</span>-transgenic <span class="html-italic">Arabidopsis</span>. (<b>G</b>) Expression analysis of individual genes in <span class="html-italic">SG</span>-transgenic <span class="html-italic">Arabidopsis</span> by qRT-PCR. Scale bar in (<b>A</b>) represents 1 cm. **** <span class="html-italic">p</span> ≤ 0.0001 represents significance.</p>
Full article ">Figure 4
<p><span class="html-italic">SG</span>-transgenic <span class="html-italic">Arabidopsis</span> tolerant to salinity stress. (<b>A</b>) Plant growth in response to various concentrations of NaCl (0, 100, 120, and 200 mM). (<b>B</b>) Plant growth in terms of fresh weight under 100 mM NaCl for four weeks. (<b>C</b>) Cell damage in terms of MDA release under 100 mM NaCl for four weeks. (<b>D</b>) Representative image showing the transgenic plants growing in earth-pot for one month. (<b>E</b>) Comparison of plant height for the transgenic plants growing in earth-pot for one month. Data are represented as means ± SE from three replicates. *** <span class="html-italic">p</span> ≤ 0.001 and **** <span class="html-italic">p</span> ≤ 0.0001 are of significance compared to <span class="html-italic">EV</span> controls. Scale bar represents 1 cm. The center line in (<b>E</b>) represents median.</p>
Full article ">Figure 5
<p><span class="html-italic">SG</span>-transgenic <span class="html-italic">Arabidopsis</span> was tolerant to drought stress. (<b>A</b>) Plant growth in response to different concentrations of PEG6000 (0, 250, 400, and 550 g/L). (<b>B</b>) Plant growth in terms of fresh weight under varied concentrations of PEG6000 for four weeks. (<b>C</b>) Cell damage in terms of MDA release under varied concentrations of PEG6000 for four weeks. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01; **** <span class="html-italic">p</span> ≤ 0.0001 represent significance. Scale bar represents 1 cm.</p>
Full article ">Figure 6
<p>Dendrobine synthesis-related gene screening by Venn diagram and KEGG analysis. (<b>A</b>) Venn distribution of DEGs for transcriptomes of different tissues (stem vs. root; leaf vs. root) from three <span class="html-italic">Dendrobium</span> species (<span class="html-italic">D. houshanense</span>; <span class="html-italic">D. catenatum</span>, and <span class="html-italic">D. moniliforme</span>). (<b>B</b>) KEGG pathway enrichment of 648 DEGs from (<b>A</b>). The x-axis represents the enrichment ratio and the y-axis represents the pathway name. (<b>C</b>) Venn diagram representation of the number of DEGs in samples from protocorm-like bodies (PLBs), samples under <span class="html-italic">MF23</span> treatment, and the 648 DEGs in (<b>A</b>). (<b>D</b>) KEGG pathway enrichment of DEGs from (<b>C</b>).</p>
Full article ">Figure 7
<p>Dendrobine synthesis-related hub gene screening by WGCNA. (<b>A</b>) Hierarchical cluster dendrogram showing six expression modules of co-expressed genes. Each leaf in the tree represents an individual gene, with the branch representing a module of highly connected genes. The designated color rows below correspond to module membership. (<b>B</b>) Scale-free fit index at different threshold values (<span class="html-italic">β</span>). Asterisk indicates the selected soft-thresholding power. (<b>C</b>) Heatmap of connectivity of eigengenes. (<b>D</b>) Module-trait correlations and corresponding <span class="html-italic">p</span>-values (in parenthesis). The color in the box indicates −log(<span class="html-italic">P</span>) and the color scale indicates the <span class="html-italic">p</span>-value from the Fisher exact test. Treatment means <span class="html-italic">MF23</span> infection.</p>
Full article ">Figure 8
<p>Functional verification of downstream genes in dendrobine synthesis. (<b>A</b>) <span class="html-italic">CMEAO</span> overexpression was verified by qRT-PCR. Samples were collected at 24 h post-infiltration. (<b>B</b>) Transiently infiltrated <span class="html-italic">D. catenatum</span> leaves (5 dpi) were harvested and subjected to dendrobine measurement. Empty vectors are used as controls. (<b>C</b>) qRT-PCR verification of <span class="html-italic">MYB61</span> overexpression in transiently infiltrated one-year-old <span class="html-italic">D. catenatum</span> leaves. (<b>D</b>) Dendrobine content in <span class="html-italic">D. catenatum</span> leaves transiently overexpressing <span class="html-italic">MYB61</span>. Statistical significance was demonstrated as following ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> ≤ 0.001, **** <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">
13 pages, 4335 KiB  
Article
Technology Invention and Mechanism Analysis of Rapid Rooting of Taxus × media Rehder Branches Induced by Agrobacterium rhizogenes
by Ying Wang, Xiumei Luo, Haotian Su, Ge Guan, Shuang Liu and Maozhi Ren
Int. J. Mol. Sci. 2024, 25(1), 375; https://doi.org/10.3390/ijms25010375 - 27 Dec 2023
Viewed by 1013
Abstract
Taxus, a vital source of the anticancer drug paclitaxel, grapples with a pronounced supply–demand gap. Current efforts to alleviate the paclitaxel shortage involve expanding Taxus cultivation through cutting propagation. However, traditional cutting propagation of Taxus is difficult to root and time-consuming. Obtaining [...] Read more.
Taxus, a vital source of the anticancer drug paclitaxel, grapples with a pronounced supply–demand gap. Current efforts to alleviate the paclitaxel shortage involve expanding Taxus cultivation through cutting propagation. However, traditional cutting propagation of Taxus is difficult to root and time-consuming. Obtaining the roots with high paclitaxel content will cause tree death and resource destruction, which is not conducive to the development of the Taxus industry. To address this, establishing rapid and efficient stem rooting systems emerges as a key solution for Taxus propagation, facilitating direct and continuous root utilization. In this study, Agrobacterium rhizogenes were induced in the 1–3-year-old branches of Taxus × media Rehder, which has the highest paclitaxel content. The research delves into the rooting efficiency induced by different A. rhizogenes strains, with MSU440 and C58 exhibiting superior effects. Transcriptome and metabolome analyses revealed A. rhizogenes’ impact on hormone signal transduction, amino acid metabolism, zeatin synthesis, and secondary metabolite synthesis pathways in roots. LC-MS-targeted quantitative detection showed no significant difference in paclitaxel and baccatin III content between naturally formed and induced roots. These findings underpin the theoretical framework for T. media rapid propagation, contributing to the sustainable advancement of the Taxus industry. Full article
Show Figures

Figure 1

Figure 1
<p>The technical process of branch rooting of <span class="html-italic">T. media</span>. (<b>a</b>) Girdling treatment of 1–3-year-old branches. (<b>b</b>) Plant high-pressure propagation box fixing at the girdling site. (<b>c</b>) Swelling and formation of root apical meristem at the upper girdling site. (<b>d</b>) Formation of young roots at the upper girdling site. (<b>e</b>) Elongation and growth of branch adventitious roots.</p>
Full article ">Figure 2
<p>Demonstration of branch rooting. (<b>a</b>) Morphology of roots after opening the high-pressure propagation box, with new shoots emerging from the branch. (<b>b</b>) The morphology of roots formed under the treatment of different <span class="html-italic">A. rhizogenes</span> strains, showing their growth from the girdling site. CK means the normally formed roots of <span class="html-italic">T. media</span> branches without the treatment of <span class="html-italic">A. rhizogenes</span>. (<b>c</b>) The rooting rate of <span class="html-italic">T. media</span> branches under the treatment of different <span class="html-italic">A. rhizogenes</span> strains that, without the treatment of <span class="html-italic">A. rhizogenes,</span> was as a control (CK). (<b>d</b>) The number of adventitious roots in <span class="html-italic">T. media</span> branches under treatments of different <span class="html-italic">A. rhizogenes</span> strains that, without the treatment of <span class="html-italic">A. rhizogenes,</span> was as a control (CK). (<b>e</b>) The growth status of the rooting branches induced by <span class="html-italic">A. rhizogenes</span> strains MSU440 and C58 after being transplanted into pots. Each experiment was repeated three times. Each replicate contains at least 35 branch rooting treatments. Asterisks indicate significant differences (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Analysis of differentially expressed genes (DEGs) in TR vs. WR; gene ontology (GO); and the Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment analysis of DEGs in TR vs. WR. (<b>a</b>) Differentially expressed genes in TR vs. WR. (<b>b</b>) GO enrichment analysis showed that biological processes, cellular anatomical entity, and oxidoreductase activity were involved in the biological process, cellular component, and molecular function, respectively. (<b>c</b>) KEGG enrichment analysis showed that metabolic pathways, biosynthesis of secondary metabolites, and phenylpropanoid biosynthesis were the top three significant enrichment pathways.</p>
Full article ">Figure 4
<p>Analysis of differentially expressed metabolites (DEMs) in TR vs. WR. (<b>a</b>) Differentially expressed metabolites in TR vs. WR. (<b>b</b>) Principal component analysis revealed clear differences between the metabolites in TR and WR. (<b>c</b>) Cluster analysis of DEMs. (<b>d</b>) KEGG enrichment analysis showed that DEMs were the main components for the biosynthesis of amino acids and their derivatives.</p>
Full article ">Figure 5
<p>Combined analysis of transcriptome and metabolome (DEGs/DEMs) in TR vs. WR. (<b>a</b>) Top ten KEGG pathways correlated with DEGs/DEMs in TR vs. WR. (<b>b</b>) Plant hormone signal transduction pathway in the correlated DEGs/DEMs in TR vs. WR. (<b>c</b>) Zeatin biosynthesis pathway in the correlated DEGs/DEMs in TR vs. WR. In (<b>b</b>,<b>c</b>), the red rectangles indicate the DEGs encoded in the protein were all upregulated, the blue rectangles indicate the DEGs encoded in the protein were all downregulated and the yellow rectangle indicates the genes encoded in the protein contain both upregulated and downregulated DEGs.</p>
Full article ">Figure 6
<p>The quantitative detection of paclitaxel and baccatin III. (<b>a</b>) The LC-MS/MS analysis of paclitaxel and baccatin III in 20-year-old roots, WR, and TR. The retention time (RT) of paclitaxel is 3.08 min and that of baccatin III is 2.86 min. (<b>b</b>) The content of paclitaxel in 20-year-old roots, WR, and TR. (<b>c</b>) The content of baccatin III in 20-year-old roots, WR, and TR. Asterisks indicate significant differences (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
15 pages, 4163 KiB  
Article
Comparison of the Metabolomics of Different Dendrobium Species by UPLC-QTOF-MS
by Tingting Zhang, Xinxin Yang, Fengzhong Wang, Pengfei Liu, Mengzhou Xie, Cong Lu, Jiameng Liu, Jing Sun and Bei Fan
Int. J. Mol. Sci. 2023, 24(24), 17148; https://doi.org/10.3390/ijms242417148 - 5 Dec 2023
Viewed by 1315
Abstract
Dendrobium Sw. (family Orchidaceae) is a renowned edible and medicinal plant in China. Although widely cultivated and used, less research has been conducted on differential Dendrobium species. In this study, stems from seven distinct Dendrobium species were subjected to UPLC-QTOF-MS/MS analysis. A total [...] Read more.
Dendrobium Sw. (family Orchidaceae) is a renowned edible and medicinal plant in China. Although widely cultivated and used, less research has been conducted on differential Dendrobium species. In this study, stems from seven distinct Dendrobium species were subjected to UPLC-QTOF-MS/MS analysis. A total of 242 metabolites were annotated, and multivariate statistical analysis was employed to explore the variance in the extracted metabolites across the various groups. The analysis demonstrated that D. nobile displays conspicuous differences from other species of Dendrobium. Specifically, D. nobile stands out from the remaining six taxa of Dendrobium based on 170 distinct metabolites, mainly terpene and flavonoid components, associated with cysteine and methionine metabolism, flavonoid biosynthesis, and galactose metabolism. It is believed that the variations between D. nobile and other Dendrobium species are mainly attributed to three metabolite synthesis pathways. By comparing the chemical composition of seven species of Dendrobium, this study identified the qualitative components of each species. D. nobile was found to differ significantly from other species, with higher levels of terpenoids, flavonoids, and other compounds that are for the cardiovascular field. By comparing the chemical composition of seven species of Dendrobium, these qualitative components have relevance for establishing quality standards for Dendrobium. Full article
Show Figures

Figure 1

Figure 1
<p>Representative total ion current (TIC) chromatograms of different species of DD (<span class="html-italic">Dendrobium denneanum</span> Kerr.), DC (<span class="html-italic">D. chrysotoxum</span> Lindl.), DN (<span class="html-italic">D. nobile</span> Lindl.) DF (<span class="html-italic">D. fimbriatum</span> Hook.), DT (<span class="html-italic">D. thyrsiflorum</span> Rchb. f.), DO (<span class="html-italic">D. officinale</span> Kimura. et Migo.), DV (<span class="html-italic">D. devonianum</span> Paxton.) (<b>A</b>) and of DN alone (<b>B</b>) under positive ionization mode.</p>
Full article ">Figure 2
<p>Proposed schematic of the mass fragmentation pathways of (<b>A</b>) (compound <b>195</b>, caryoptosidic acid); (<b>B</b>) (compound <b>10</b>, naringenin chalcone); (<b>C</b>) (compound <b>33</b>, mangiferin); (<b>D</b>) (compound <b>63</b>, <span class="html-italic">N</span>-Oxysophocarpine); (<b>E</b>) (compound <b>197</b>, 6,7-dihydroxy-4-methyl coumarin); and (<b>F</b>) (compound <b>134</b>, lusianthridin).</p>
Full article ">Figure 3
<p>Principal component analysis of different species of DD (<span class="html-italic">Dendrobium denneanum</span> Kerr.), DC (<span class="html-italic">D. chrysotoxum</span> Lindl.), DN (<span class="html-italic">D. nobile</span> Lindl.), DF (<span class="html-italic">D. fimbriatum</span> Hook.), DT (<span class="html-italic">D. thyrsiflorum</span> Rchb. f.), DO (<span class="html-italic">D. officinale</span> Kimura. et Migo.), DV (<span class="html-italic">D. devonianum</span> Paxton.) (<b>A</b>); a comparison of verified DN and other species of <span class="html-italic">Dendrobium</span> (<b>B</b>).</p>
Full article ">Figure 4
<p>A Venn diagram of variance components (<b>A</b>); a heat map of the differential compositions (<b>B</b>).</p>
Full article ">Figure 5
<p>The outcomes of the metabolic pathway analysis. The main differentiating metabolites were annotated and analyzed about metabolic pathways. Each bubble on the bubble diagram represents a metabolic pathway. Technical term abbreviations are explained upon their first use. The horizontal axis and the size of the bubble indicate the size of the pathway influencing factor in the topology analysis. The greater the scale, the bigger the influencing factor. The vertical position of the bubble shows the <span class="html-italic">p</span>-value of the enrichment analysis.</p>
Full article ">Figure 6
<p>The key differential metabolites of DN and different species of <span class="html-italic">Dendrobium</span> were mapped to metabolic pathways, and the results of metabolic pathway analysis are shown as KEGG pathway results: (<b>A</b>,<b>D</b>) galactose metabolism; (<b>B</b>,<b>F</b>) flavonoid biosynthesis; (<b>C</b>) cysteine and methionine metabolism; (<b>E</b>) flavone and flavonol biosynthesis.</p>
Full article ">
17 pages, 2190 KiB  
Article
Evaluation of Reference Genes for Normalizing RT-qPCR and Analysis of the Expression Patterns of WRKY1 Transcription Factor and Rhynchophylline Biosynthesis-Related Genes in Uncaria rhynchophylla
by Detian Mu, Yingying Shao, Jialong He, Lina Zhu, Deyou Qiu, Iain W. Wilson, Yao Zhang, Limei Pan, Yu Zhou, Ying Lu and Qi Tang
Int. J. Mol. Sci. 2023, 24(22), 16330; https://doi.org/10.3390/ijms242216330 - 15 Nov 2023
Cited by 4 | Viewed by 1147
Abstract
Uncaria rhynchophylla (Miq.) Miq. ex Havil, a traditional medicinal herb, is enriched with several pharmacologically active terpenoid indole alkaloids (TIAs). At present, no method has been reported that can comprehensively select and evaluate the appropriate reference genes for gene expression analysis, especially the [...] Read more.
Uncaria rhynchophylla (Miq.) Miq. ex Havil, a traditional medicinal herb, is enriched with several pharmacologically active terpenoid indole alkaloids (TIAs). At present, no method has been reported that can comprehensively select and evaluate the appropriate reference genes for gene expression analysis, especially the transcription factors and key enzyme genes involved in the biosynthesis pathway of TIAs in U. rhynchophylla. Reverse transcription quantitative PCR (RT-qPCR) is currently the most common method for detecting gene expression levels due to its high sensitivity, specificity, reproducibility, and ease of use. However, this methodology is dependent on selecting an optimal reference gene to accurately normalize the RT-qPCR results. Ten candidate reference genes, which are homologues of genes used in other plant species and are common reference genes, were used to evaluate the expression stability under three stress-related experimental treatments (methyl jasmonate, ethylene, and low temperature) using multiple stability analysis methodologies. The results showed that, among the candidate reference genes, S-adenosylmethionine decarboxylase (SAM) exhibited a higher expression stability under the experimental conditions tested. Using SAM as a reference gene, the expression profiles of 14 genes for key TIA enzymes and a WRKY1 transcription factor were examined under three experimental stress treatments that affect the accumulation of TIAs in U. rhynchophylla. The expression pattern of WRKY1 was similar to that of tryptophan decarboxylase (TDC) under ETH treatment. This research is the first to report the stability of reference genes in U. rhynchophylla and provides an important foundation for future gene expression analyses in U. rhynchophylla. The RT-qPCR results indicate that the expression of WRKY1 is similar to that of TDC under ETH treatment. It may coordinate the expression of TDC, providing a possible method to enhance alkaloid production in the future through synthetic biology. Full article
Show Figures

Figure 1

Figure 1
<p>The biosynthetic pathway of rhynchophylline and isorhynchophylline in <span class="html-italic">U. rhynchophylla.</span> Solid arrows represent genes identified in the pathway and dashed arrows represent presumed genes.</p>
Full article ">Figure 2
<p>The Ct values of 10 candidate reference genes in all samples. The expression data are displayed as the Ct value of each reference gene in the sample of <span class="html-italic">U. rhynchophylla.</span> The boxplot indicates the 25th/75th percentiles and it contains the mean, maximum, and minimum values.</p>
Full article ">Figure 3
<p>Average expression stability values (M) of the ten candidate reference genes using GeNorm. The expression stability was evaluated in samples from leaves of <span class="html-italic">U. rhynchophylla</span> subjected to low temperatures, MeJA (100 μmol), ETH (100 μmol), control (not treated), and total (all treated samples). The least stable genes are on the left with higher M values and the most stable genes are on the right with lower M values.</p>
Full article ">Figure 4
<p>The pairwise variation (Vn/n + 1) scores of ten candidates measured by GeNorm. Different treatments are identified by different colors. The value used to determine the appropriate number of reference genes for RT-qPCR normalization is 0.15.</p>
Full article ">Figure 5
<p>The relative expression of the <span class="html-italic">WRKY1</span> transcription factor and key enzyme genes in response to MeJA, ETH, and low temperatures. Different colors represent different times. (<b>a</b>) MeJA treatment; (<b>b</b>) ethylene treatment; (<b>c</b>) low-temperature treatment. The error bars represent the mean ± SD from three biological replicates, and asterisks indicate statistically significant differences compared with the controls (0 h).* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
22 pages, 10575 KiB  
Article
Reference Genes Screening and Gene Expression Patterns Analysis Involved in Gelsenicine Biosynthesis under Different Hormone Treatments in Gelsemium elegans
by Yao Zhang, Detian Mu, Liya Wang, Xujun Wang, Iain W. Wilson, Wenqiang Chen, Jinghan Wang, Zhaoying Liu, Deyou Qiu and Qi Tang
Int. J. Mol. Sci. 2023, 24(21), 15973; https://doi.org/10.3390/ijms242115973 - 4 Nov 2023
Cited by 3 | Viewed by 1137
Abstract
Reverse transcription quantitative polymerase chain reaction (RT-qPCR) is an accurate method for quantifying gene expression levels. Choosing appropriate reference genes to normalize the data is essential for reducing errors. Gelsemium elegans is a highly poisonous but important medicinal plant used for analgesic and [...] Read more.
Reverse transcription quantitative polymerase chain reaction (RT-qPCR) is an accurate method for quantifying gene expression levels. Choosing appropriate reference genes to normalize the data is essential for reducing errors. Gelsemium elegans is a highly poisonous but important medicinal plant used for analgesic and anti-swelling purposes. Gelsenicine is one of the vital active ingredients, and its biosynthesis pathway remains to be determined. In this study, G. elegans leaf tissue with and without the application of one of four hormones (SA, MeJA, ETH, and ABA) known to affect gelsenicine synthesis, was analyzed using ten candidate reference genes. The gene stability was evaluated using GeNorm, NormFinder, BestKeeper, ∆CT, and RefFinder. The results showed that the optimal stable reference genes varied among the different treatments and that at least two reference genes were required for accurate quantification. The expression patterns of 15 genes related to the gelsenicine upstream biosynthesis pathway was determined by RT-qPCR using the relevant reference genes identified. Three genes 8-HGO, LAMT, and STR, were found to have a strong correlation with the amount of gelsenicine measured in the different samples. This research is the first study to examine the reference genes of G. elegans under different hormone treatments and will be useful for future molecular analyses of this medically important plant species. Full article
Show Figures

Figure 1

Figure 1
<p>The biosynthesis pathway of MIAs in <span class="html-italic">G. elegans</span> (solid arrows represent known synthetic pathways and the dashed arrows represent speculative pathways). GES: Geraniol synthase; G8H: Geraniol 8-hydroxylase; 8-HGO: 8-hydroxygeraniol dehydrogenase; ISY: (S)-8-oxocitronellyl enol synthase; NEPSI: (+)-cis,trans-nepetalactol synthase; 7-DLS: 7-deoxyloganetic acid synthase; 7-DLGT: 7-deoxyloganetic acid glucosyltransferase; 7-DLH: 7-deoxyloganate 7-hydroxylase; LAMT: Loganate methyltransferase; SLS: Secologanin synthase; AS: Anthranilate synthase; AnPRT: Anthranilate phosphoribosyltransferase; PRAI: Phosphoribosylanthranilate isomerase; IGPS: Indole-3-glycerol phosphate synthase; TSA: Tryptophan synthase alpha chain; TSB: Tryptophan synthase beta chain; TDC: L-tryptophan decarboxylase; STR: Strictosidine synthase; SGD: Strictosidine-β-D- glucosidase; CYP450: Cytochrome P450 enzymes; NMT: N-methyltransferase.</p>
Full article ">Figure 2
<p>The content of gelsenicine under four hormone treatments at 0–48 h. The (*) indicates <span class="html-italic">p</span> &lt; 0.05, and the (**) indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Distribution of Ct values among the 10 candidate reference genes. Lines shown in the box-plot graph of Ct value display the median values. Lower and upper boxes represent the 25th percentile to the 75th percentile. Whiskers indicate the maximum and minimum values.</p>
Full article ">Figure 4
<p>Average expression stability values (M) of 10 candidate reference genes using GeNorm.</p>
Full article ">Figure 5
<p>Pairwise variation (V) values obtained by GeNorm analysis on 10 candidate reference genes in the six groups.</p>
Full article ">Figure 6
<p>Validation of the reference gene by the relative expression of the target gene <span class="html-italic">STR</span> in different hormone treatments. Graphs (<b>a</b>–<b>d</b>) are the results of SA, MeJA, ETH and ABA. The (*) indicates <span class="html-italic">p</span> &lt; 0.05, and the (**) indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>(<b>a</b>) The results for the expression patterns of 15 candidate genes related to the upstream pathway of MIA biosynthesis at 0–48 h of SA treatment; (<b>b</b>) the results for the expression patterns of 15 candidate genes related to the upstream pathway of MIA biosynthesis at 0–48 h of MeJA treatment; (<b>c</b>) the results for the expression patterns of 15 candidate genes related to the upstream pathway of MIA biosynthesis at 0–48 h of ETH treatment; (<b>d</b>) the results for the expression patterns of 15 candidate genes related to the upstream pathway of MIA biosynthesis at 0–48 h of ABA treatment. The (*) indicates <span class="html-italic">p</span> &lt; 0.05, and the (**) indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>(<b>a</b>) The results for the heatmaps of gelsenicine and the 15 upstream biosynthetic pathway genes associated with it under SA treatment; (<b>b</b>) the results for the heatmaps of gelsenicine and the 15 upstream biosynthetic pathway genes associated with it under MeJA treatment; (<b>c</b>) the results for the heatmaps of gelsenicine and the 15 upstream biosynthetic pathway genes associated with it under ETH treatment; (<b>d</b>) the results for the heatmaps of gelsenicine and the 15 upstream biosynthetic pathway genes associated with it under ABA treatment. The darker the colors, the stronger is the correlation.</p>
Full article ">Figure 9
<p>(<b>a</b>) The results for the correlation analyses between gelsenicine and 15 upstream biosynthetic pathway genes under SA treatment; (<b>b</b>) the results for the correlation analyses between gelsenicine and 15 upstream biosynthetic pathway genes under MeJA treatment; (<b>c</b>) the results for the correlation analyses between gelsenicine and 15 upstream biosynthetic pathway genes under ETH treatment; (<b>d</b>) the results for the correlation analyses between gelsenicine and 15 upstream biosynthetic pathway genes under ABA treatment. Thick lines (weak correlation), thin lines (strong correlation), red and rightward slash (positive correlation), blue and leftward slash (negative correlation), only the more relevance cases are shown with asterisks [<a href="#B43-ijms-24-15973" class="html-bibr">43</a>]. The darker the colors, the more asterisks, the stronger is the correlation. The (*) indicates <span class="html-italic">p</span> &lt; 0.05, and the (**) indicates <span class="html-italic">p</span> &lt; 0.01. The (***) indicates <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>) The results for the correlation analyses between gelsenicine and 15 upstream biosynthetic pathway genes under SA treatment; (<b>b</b>) the results for the correlation analyses between gelsenicine and 15 upstream biosynthetic pathway genes under MeJA treatment; (<b>c</b>) the results for the correlation analyses between gelsenicine and 15 upstream biosynthetic pathway genes under ETH treatment; (<b>d</b>) the results for the correlation analyses between gelsenicine and 15 upstream biosynthetic pathway genes under ABA treatment. Thick lines (weak correlation), thin lines (strong correlation), red and rightward slash (positive correlation), blue and leftward slash (negative correlation), only the more relevance cases are shown with asterisks [<a href="#B43-ijms-24-15973" class="html-bibr">43</a>]. The darker the colors, the more asterisks, the stronger is the correlation. The (*) indicates <span class="html-italic">p</span> &lt; 0.05, and the (**) indicates <span class="html-italic">p</span> &lt; 0.01. The (***) indicates <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 10
<p>Screening of genes related to the biosynthetic pathway of MIAs based on co-expression correlation analysis, shown here with <span class="html-italic">8-HGO</span> as an example.</p>
Full article ">
14 pages, 3086 KiB  
Article
Synthesis of Crocin I and Crocin II by Multigene Stacking in Nicotiana benthamiana
by Lei Xie, Zuliang Luo, Xunli Jia, Changming Mo, Xiyang Huang, Yaran Suo, Shengrong Cui, Yimei Zang, Jingjing Liao and Xiaojun Ma
Int. J. Mol. Sci. 2023, 24(18), 14139; https://doi.org/10.3390/ijms241814139 - 15 Sep 2023
Cited by 3 | Viewed by 2108
Abstract
Crocins are a group of highly valuable water-soluble carotenoids that are reported to have many pharmacological activities, such as anticancer properties, and the potential for treating neurodegenerative diseases including Alzheimer’s disease. Crocins are mainly biosynthesized in the stigmas of food–medicine herbs Crocus sativus [...] Read more.
Crocins are a group of highly valuable water-soluble carotenoids that are reported to have many pharmacological activities, such as anticancer properties, and the potential for treating neurodegenerative diseases including Alzheimer’s disease. Crocins are mainly biosynthesized in the stigmas of food–medicine herbs Crocus sativus L. and Gardenia jasminoides fruits. The distribution is narrow in nature and deficient in resources, which are scarce and expensive. Recently, the synthesis of metabolites in the heterologous host has opened up the potential for large-scale and sustainable production of crocins, especially for the main active compounds crocin I and crocin II. In this study, GjCCD4a, GjALDH2C3, GjUGT74F8, and GjUGT94E13 from G. jasminoides fruits were expressed in Nicotiana benthamiana. The highest total content of crocins in T1 generation tobacco can reach 78,362 ng/g FW (fresh weight) and the dry weight is expected to reach 1,058,945 ng/g DW (dry weight). Surprisingly, the primary effective constituents crocin I and crocin II can account for 99% of the total crocins in transgenic plants. The strategy mentioned here provides an alternative platform for the scale-up production of crocin I and crocin II in tobacco. Full article
Show Figures

Figure 1

Figure 1
<p>The crocin biosynthetic pathway in <span class="html-italic">Crocus sativus</span> L. (pink) and <span class="html-italic">Gardenia jasminoides</span> (blue). PSY, phytoene synthase; PDS, phytoene desaturase; Z-ISO, ζ-carotene isomerase; ZDS, ζ-carotene desaturase; CRTISO, carotenoid isomerase; LCYB, lycopene β-cyclase; BHY, β-carotene hydrolase; GjCCD4a and CsCCD2, carotenoid cleavage dioxygenase; GjALDH and CsALDH3l1, aldehyde dehydrogenase; GjUGT74F8, GjUGT94E13, CsUGT74AD1, and UGT709G1, UDP-glucosyltransferase.</p>
Full article ">Figure 2
<p>Genetic transformation of AU-CU vector plants of T0 generation. (<b>A</b>). T-DNA region of the multigene vector AU-CU that was transformed into the tobacco plants. (<b>B</b>). Compare the leaves, roots, and stems of AU-CU transgenic plants with those of the wild-type.</p>
Full article ">Figure 3
<p>Genetic Transformation of <span class="html-italic">Nicotiana benthamiana</span> with multigene vector synthesizing crocins and molecular identification. (<b>A</b>) The relative gene expression level analysis of crocins synthesis genes in transgenic tobacco lines. The Nbactin was used as an internal control and the wild-type was set to 1. The data are presented as the mean values ± SDs, <span class="html-italic">n</span> = 3 biologically independent samples. (<b>B</b>) PCR detection of crocins biosynthetic genes. The first lane is the DNA maker (4500 bp), the second is the wild type, and then from left to right are transgenic <span class="html-italic">Nicotiana benthamiana</span> lines N16, N18, N20, N22, and N34. (<b>C</b>) PCR to identify whether the T1 generation carries <span class="html-italic">SgCCD4a</span> and <span class="html-italic">SgUGT74F8</span>. The first lane of each row is the DNA maker (2000 bp), the second is the wild type, and the rest are the T1 generations.</p>
Full article ">Figure 4
<p>Comparison of the production of crocins crocetin between <span class="html-italic">Crocus sativus</span> L. stigmas, <span class="html-italic">Gardenia jasminoides</span> fruits, and transgenic tobacco. The red curve represents HPLC-MS/MS analysis of crocins in <span class="html-italic">Crocus sativus</span> L. stigmas and <span class="html-italic">Gardenia jasminoides</span> fruits. The blue curve represents HPLC-MS/MS analysis of crocins and crocetin standards. The orange curve represents HPLC-MS/MS analysis of crocins in N18, N20, and N34 lines; t-I, c-I, t-II, c-II, t-III, c-III, t-IV, and t-V represent trans-crocin I, cis-crocin I, trans-crocin II, cis-crocin II, trans-crocin III, cis-crocin III, trans-crocin IV and trans-crocin V, respectively.</p>
Full article ">Figure 5
<p>Probability density distribution curve of T1 generation crocin I and crocin II. (<b>A</b>). Probability density distribution curve of crocin I. (<b>B</b>). Probability density distribution curve of crocin II. The red curve represents the T1 generations of the N16 line; the green represents the T1 generations of the N18 lines; and the blue represents the T1 generations of the N20 line. The black represents the T1 generations of the N22 line and the pink represents the T1 generations of the N34 line. The curve plotting data are the mean values ± SDs, <span class="html-italic">n</span> = 3 biologically independent samples.</p>
Full article ">Figure 6
<p>Accumulation of crocins in transgenic tobacco lines of T1 generation. The data are presented as the mean values ± SDs, <span class="html-italic">n</span> = 3 biologically independent samples.</p>
Full article ">
15 pages, 3372 KiB  
Article
Characterization of Volatile Organic Compounds in Five Celery (Apium graveolens L.) Cultivars with Different Petiole Colors by HS-SPME-GC-MS
by Yue Sun, Mengyao Li, Xiaoyan Li, Jiageng Du, Weilong Li, Yuanxiu Lin, Yunting Zhang, Yan Wang, Wen He, Qing Chen, Yong Zhang, Xiaorong Wang, Ya Luo, Aisheng Xiong and Haoru Tang
Int. J. Mol. Sci. 2023, 24(17), 13343; https://doi.org/10.3390/ijms241713343 - 28 Aug 2023
Cited by 4 | Viewed by 1639
Abstract
Celery (Apium graveolens L.) is an important vegetable crop cultivated worldwide for its medicinal properties and distinctive flavor. Volatile organic compound (VOC) analysis is a valuable tool for the identification and classification of species. Currently, less research has been conducted on aroma [...] Read more.
Celery (Apium graveolens L.) is an important vegetable crop cultivated worldwide for its medicinal properties and distinctive flavor. Volatile organic compound (VOC) analysis is a valuable tool for the identification and classification of species. Currently, less research has been conducted on aroma compounds in different celery varieties and colors. In this study, five different colored celery were quantitatively analyzed for VOCs using HS-SPME, GC-MS determination, and stoichiometry methods. The result revealed that γ-terpinene, d-limonene, 2-hexenal,-(E)-, and β-myrcene contributed primarily to the celery aroma. The composition of compounds in celery exhibited a correlation not only with the color of the variety, with green celery displaying a higher concentration compared with other varieties, but also with the specific organ, whereby the content and distribution of volatile compounds were primarily influenced by the leaf rather than the petiole. Seven key genes influencing terpenoid synthesis were screened to detect expression levels. Most of the genes exhibited higher expression in leaves than petioles. In addition, some genes, particularly AgDXS and AgIDI, have higher expression levels in celery than other genes, thereby influencing the regulation of terpenoid synthesis through the MEP and MVA pathways, such as hydrocarbon monoterpenes. This study identified the characteristics of flavor compounds and key aroma components in different colored celery varieties and explored key genes involved in the regulation of terpenoid synthesis, laying a theoretical foundation for understanding flavor chemistry and improving its quality. Full article
Show Figures

Figure 1

Figure 1
<p>Total ion flow diagram of volatile compounds in leaves (<b>A</b>) and petioles (<b>B</b>) of five celery varieties. Different colors of writing screened colors of celery, including green for ‘Hanyu Nencuixiqin’ and ‘Siji Lvxiangqin’, white for ‘Jingpin Saixue’, purple for ‘Ziyu Xiangqin’, and yellow for ‘Shanghai Huangxinqin’.</p>
Full article ">Figure 2
<p>Analysis of compound species of five celery varieties. (<b>A</b>) A Venn diagram showed the differences and common compounds of five celery varieties. (<b>B</b>) The percentage stacking chart of eight kinds of compounds in five celery varieties.</p>
Full article ">Figure 3
<p>Characteristic compound analysis of high-content compounds (Benzene,-1-methyl-3-(1-methylethyl)-abbreviated as Isopropyl M-Tolyl Sulfide). (<b>A</b>) The nine principle compounds in celery. Each error line represents the mean ±SD (standard deviation), <span class="html-italic">t</span>-test with significance level of 0.05 (<span class="html-italic">p</span> &lt; 0.05). The different letters with lowcase (a–e) indicate significance differences, the same letter indicating no significance difference. (<b>B</b>) The correlation between tissues, cultivars, and VOCs. The horizontal axis stands for five cultivars. On the x-axis, from A to E, there were five types of celery. A for ‘Hanyu Nencuixiqin’, B for ‘Jingpin Saixue’, C for ‘Ziyu Xiangqin’, D for ‘Shanghai Huangxinqin’, and E for ‘Siji Lvxiangqin’; 1 for leaves, 2 for petioles.</p>
Full article ">Figure 4
<p>PCA analysis of two tissues, the leaf and petiole, of five celery varieties. (<b>A</b>)The summary of PC1 and PC2 and significant variability between different celery varieties and different celery parts. (<b>B</b>) Principle analysis of the influence of compounds in different samples.</p>
Full article ">Figure 5
<p>OPLS-DA chart of 18 compounds in five celery varieties. (<b>A</b>) Orthogonal partial least squares discriminant analysis of 18 compounds. The score t1 (first component) explains the largest variation of the X space, followed by t2. The numbers besides t[1] and t[2] shown the weight of eh regrression coefficient. (<b>B</b>) Variable importance for the projection numbers. (<b>C</b>) All compounds differ between two tissues in five celery varieties. The letters (A–J) shows the ten examples of celery, for leaf and petiole of ‘Hanyu Nencuixiqin’, ‘Jingpin Saixue’, ‘Ziyu Xiangqin’, ‘Shanghai Huangxinqin’, ‘Siji Lvxiangqin’, respectively.</p>
Full article ">Figure 6
<p>Analysis of key regulatory genes in the terpenoid synthesis pathway. (<b>A</b>) Gene expression in distinctively colored celery. (<b>B</b>) Key genes and compounds in MVA and MEP pathways. (<b>C</b>) Gene expression in distinctive tissues. (<b>D</b>) RT-qPCR analysis of gene expression levels in pathway of terpenoid synthesis. Column chart in the same color with the expression of the leaf on the left side and the petiole on the right side. Each error line represents the mean ± SD (standard deviation), <span class="html-italic">t</span>-test with significance level of 0.05 (<span class="html-italic">p</span> &lt; 0.05). The different letters with lowcase (a–e) indicate significance differences, the same letter indicating no significance difference.</p>
Full article ">
19 pages, 4059 KiB  
Article
Biosynthesis of α-Bisabolol by Farnesyl Diphosphate Synthase and α-Bisabolol Synthase and Their Related Transcription Factors in Matricaria recutita L.
by Yuling Tai, Honggang Wang, Ping Yao, Jiameng Sun, Chunxiao Guo, Yifan Jin, Lu Yang, Youhui Chen, Feng Shi, Luyao Yu, Shuangshuang Li and Yi Yuan
Int. J. Mol. Sci. 2023, 24(2), 1730; https://doi.org/10.3390/ijms24021730 - 15 Jan 2023
Cited by 3 | Viewed by 2996
Abstract
The essential oil of German chamomile (Matricaria recutita L.) is widely used in food, cosmetics, and the pharmaceutical industry. α-Bisabolol is the main active substance in German chamomile. Farnesyl diphosphate synthase (FPS) and α-bisabolol synthase (BBS) are key enzymes related to the [...] Read more.
The essential oil of German chamomile (Matricaria recutita L.) is widely used in food, cosmetics, and the pharmaceutical industry. α-Bisabolol is the main active substance in German chamomile. Farnesyl diphosphate synthase (FPS) and α-bisabolol synthase (BBS) are key enzymes related to the α-bisabolol biosynthesis pathway. However, little is known about the α-bisabolol biosynthesis pathway in German chamomile, especially the transcription factors (TFs) related to the regulation of α-bisabolol synthesis. In this study, we identified MrFPS and MrBBS and investigated their functions by prokaryotic expression and expression in hairy root cells of German chamomile. The results suggest that MrFPS is the key enzyme in the production of sesquiterpenoids, and MrBBS catalyzes the reaction that produces α-bisabolol. Subcellular localization analysis showed that both MrFPS and MrBBS proteins were located in the cytosol. The expression levels of both MrFPS and MrBBS were highest in the extension period of ray florets. Furthermore, we cloned and analyzed the promoters of MrFPS and MrBBS. A large number of cis-acting elements related to light responsiveness, hormone response elements, and cis-regulatory elements that serve as putative binding sites for specific TFs in response to various biotic and abiotic stresses were identified. We identified and studied TFs related to MrFPS and MrBBS, including WRKY, AP2, and MYB. Our findings reveal the biosynthesis and regulation of α-bisabolol in German chamomile and provide novel insights for the production of α-bisabolol using synthetic biology methods. Full article
Show Figures

Figure 1

Figure 1
<p>Gene cloning and SDS-PAGE analysis of MrFPS and MrBBS expressed in <span class="html-italic">E. coli</span>, and enzyme activity measurement. Note: (<b>a</b>): Electrophoreretogram of <span class="html-italic">MrFPS</span> amplification; (<b>b</b>): SDS-PAGE analysis of MrFPS; (<b>c</b>): GC–MS analysis of the enzyme activity of MrFPS; (<b>d</b>): Mass spectral analysis of farnesol; (<b>e</b>): Electrophoreretogram of <span class="html-italic">MrBBS</span> amplification; (<b>f</b>): SDS-PAGE analysis of MrBBS; (<b>g</b>): GC–MS analysis of the enzyme activity of MrBBS; (<b>h</b>): Mass spectral analysis of α-bisabolol; (<b>i</b>): Standard mass spectra of farnesol; (<b>j</b>): Standard mass spectra of α-bisabolol; b: Lane 1, before IPTG induction of empty vector pet32; lane 2, after IPTG induction of empty vector pet32; lane 3, soluble pet32 protein, lane 4, insoluble pet32 protein, lane 5, before IPTG induction of recombinant protein MrFPS, lane 6, after IPTG induction of recombinant protein MrFPS, lane 7, soluble MrFPS protein, lane 8, insoluble MrFPS protein; f:Lane 1, before IPTG induction of empty vector pet32; lane 2, after IPTG induction of empty vector pCold TF;; lane 3, soluble pCold TF; protein, lane 4, insoluble pCold TF; protein, lane 5, before IPTG induction of recombinant protein MrBBS, lane 6, after IPTG induction of recombinant protein MrBBS, lane 7, soluble MrBBS protein, lane 8, insoluble MrBBS protein.</p>
Full article ">Figure 2
<p>PCR analysis of DNA from <span class="html-italic">MrFPS</span> (<b>a</b>) and <span class="html-italic">MrBBS</span> (<b>b</b>) transgenic hairy roots using primers specific to the <span class="html-italic">rol B</span> and <span class="html-italic">rolC</span> genes. (<b>c</b>) PCR analysis of DNA from transgenic hairy roots using primers specific to <span class="html-italic">MrFPS</span> and <span class="html-italic">MrBBS.</span> Note: Lane M, markers; c: lanes 1–3, ATCC15834 DNA; lanes 4–6, pK7GWF2.0 DNA; lanes 7–9, Pcambia1302; lanes 10–12, PCR products of <span class="html-italic">MrBBS</span> from putative transformant; lanes 11–15, PCR products of <span class="html-italic">MrFPS</span> from putative transformants; qPCR analysis on DNA from wild-type roots and overexpression of <span class="html-italic">MrFPS</span> in hairy roots (<b>d</b>); qPCR analysis on DNA from wild-type roots and overexpression of <span class="html-italic">MrBBS</span> in hairy roots (<b>e</b>).</p>
Full article ">Figure 3
<p>GC–MS analysis of overexpression of <span class="html-italic">MrFPS</span> (<b>a</b>–<b>c</b>) and <span class="html-italic">MrBBS</span> (<b>d</b>) in German chamomile hairy roots. Relative content analysis of α-guaiene, <span class="html-italic">cis</span>-α-bisabolene, α-farnesene, and α-bisabolol in overexpressing German chamomile hairy roots (<b>e</b>–<b>h</b>). (<b>i</b>) Mass spectrometry analysis of α-guaiene. (<b>j</b>) Mass spectrometry analysis of <span class="html-italic">cis</span>-α-bisabolene. (<b>k</b>) Mass spectrometry analysis of α-farnesene f. (<b>l</b>) Mass spectrometry analysis of α-bisabolol. Note: pCAMBIA1302 indicates hairy roots transformed with empty vector. Error bars are shown with three biological replicates (<span class="html-italic">t</span>-test). One asterisk (*) indicates a significant difference (0.01 &lt; <span class="html-italic">p</span> &lt; 0.05) and ** indicate a very significant difference (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Subcellular localization of MrFPS and MrBBS in tobacco. All samples were visualized using confocal microscopy. GFP means GFP fluorescence; Bright means Light field; Merged means superposition of fluorescence. Control vector (35S: GFP) and recombinant vectors (35S: <span class="html-italic">MrFPS</span>-GFP and 35S: <span class="html-italic">MrBBS</span>-GFP) were expressed in protoplasts of tobacco.</p>
Full article ">Figure 5
<p>Subcellular localization of MrFPS and MrBBS in protoplasts of <span class="html-italic">Arabidopsis thaliana.</span> All samples were visualized using confocal microscopy. GFP means GFP fluorescence; Chlorophyll means Chlorophyll fluorescence; Bright means Light field; Merged means superposition of fluorescence. Control vector (35S: GFP) and recombinant vectors (35S: <span class="html-italic">MrFPS</span>-GFP and 35S: <span class="html-italic">MrBBS</span>-GFP) were expressed in protoplasts of <span class="html-italic">Arabidopsis thaliana</span>.</p>
Full article ">Figure 6
<p>Expression patterns of <span class="html-italic">MrFPS</span> (<b>a</b>) and <span class="html-italic">MrBBS</span> (<b>b</b>) in German chamomile, and phylogenetic tree analysis of <span class="html-italic">MrBBS</span> in German chamomile (<b>c</b>). Each point represents one independent measurement.</p>
Full article ">Figure 7
<p>PCR amplification of TFs related to <span class="html-italic">MrFPS</span> (<b>a</b>–<b>c</b>) and <span class="html-italic">MrBBS</span> (<b>d</b>). Lane M, markers; lane 1, <span class="html-italic">MrMYB1</span>; lane 2, <span class="html-italic">MrAP22</span>; lane 3, <span class="html-italic">MrWRKY1</span>; lane 4, <span class="html-italic">MrAP21</span>; lane 5, <span class="html-italic">MrMYB2</span>; lane 6, <span class="html-italic">MrMYB3</span>; lane 7, <span class="html-italic">MrMYB4</span>; lane 8, <span class="html-italic">MrMYB5</span>; lane 9, <span class="html-italic">MrMYB6</span>; lane 10, <span class="html-italic">MrAP23</span>; lane 11, <span class="html-italic">MrAP24</span>; lane 12, <span class="html-italic">MrAP25</span>; lane 13, <span class="html-italic">MrAP26</span>; lane 14, <span class="html-italic">MrAP27</span>; lane 15, <span class="html-italic">MrAP28</span>; lane 16, <span class="html-italic">MrWRKY2</span>; lane 17, <span class="html-italic">MrWRKY3</span>; lane 18, <span class="html-italic">MrWRKY4</span>.</p>
Full article ">Figure 8
<p>Dual-luciferase reporter analysis of <span class="html-italic">MrFPS</span> and <span class="html-italic">MrBBS</span> promoters and transcription factors. (<b>a</b>–<b>c</b>) Double luciferase activity of the <span class="html-italic">MrFPS</span> promoter with <span class="html-italic">MrWRKY1</span>, <span class="html-italic">MrAP21</span>, and <span class="html-italic">MrAP22</span>. (<b>d</b>–<b>l</b>) Double luciferase activity of the <span class="html-italic">MrBBS</span> promoter with <span class="html-italic">MrMYB3</span>, <span class="html-italic">MrMYB4</span>, <span class="html-italic">MrAP24</span>, <span class="html-italic">MrAP25</span>, <span class="html-italic">MrAP26</span>, <span class="html-italic">MrAP28</span>, <span class="html-italic">MrWRKY2</span>, <span class="html-italic">MrWRKY3</span>, and <span class="html-italic">MrWRKY4</span>. (<b>m</b>) Analysis of the double luciferase activity of the <span class="html-italic">MrFPS</span> promoter with <span class="html-italic">MrWRKY1</span>, <span class="html-italic">MrAP21</span> and <span class="html-italic">MrMYB1</span>, and <span class="html-italic">MrAP22</span>. (<b>n</b>) Analysis of the double luciferase activity of the <span class="html-italic">MrBBS</span> promoter with <span class="html-italic">MrMYB3</span>, <span class="html-italic">MrMYB4</span>, <span class="html-italic">MrAP24</span>, <span class="html-italic">MrAP25</span>, <span class="html-italic">MrAP26</span>, <span class="html-italic">MrAP28</span>, <span class="html-italic">MrWRKY2</span>, <span class="html-italic">MrWRKY3</span>, and <span class="html-italic">MrWRKY4</span>. Note: Bacterial suspension containing pGreen-0800 and transcription factors was mixed for controls. Error bars reflect three biological replicates. The color difference represents the intensity of the fluorescence value.</p>
Full article ">Figure 9
<p>The biosynthesis pathway of α-bisabolol in German chamomile.</p>
Full article ">
14 pages, 5021 KiB  
Article
A Novel R2R3-MYB Transcription Factor SbMYB12 Positively Regulates Baicalin Biosynthesis in Scutellaria baicalensis Georgi
by Wentao Wang, Suying Hu, Jing Yang, Caijuan Zhang, Tong Zhang, Donghao Wang, Xiaoyan Cao and Zhezhi Wang
Int. J. Mol. Sci. 2022, 23(24), 15452; https://doi.org/10.3390/ijms232415452 - 7 Dec 2022
Cited by 13 | Viewed by 1680
Abstract
Scutellaria baicalensis Georgi is an annual herb from the Scutellaria genus that has been extensively used as a traditional medicine for over 2000 years in China. Baicalin and other flavonoids have been identified as the principal bioactive ingredients. The biosynthetic pathway of baicalin [...] Read more.
Scutellaria baicalensis Georgi is an annual herb from the Scutellaria genus that has been extensively used as a traditional medicine for over 2000 years in China. Baicalin and other flavonoids have been identified as the principal bioactive ingredients. The biosynthetic pathway of baicalin in S. baicalensis has been elucidated; however, the specific functions of R2R3-MYB TF, which regulates baicalin synthesis, has not been well characterized in S. baicalensis to date. Here, a S20 R2R3-MYB TF (SbMYB12), which encodes 263 amino acids with a length of 792 bp, was expressed in all tested tissues (mainly in leaves) and responded to exogenous hormone methyl jasmonate (MeJA) treatment. The overexpression of SbMYB12 significantly promoted the accumulation of flavonoids such as baicalin and wogonoside in S. baicalensis hairy roots. Furthermore, biochemical experiments revealed that SbMYB12 is a nuclear-localized transcription activator that binds to the SbCCL7-4, SbCHI-2, and SbF6H-1 promoters to activate their expression. These results illustrate that SbMYB12 positively regulates the generation of baicalin and wogonoside. In summary, this work revealed a novel S20 R2R3-MYB regulator and enhances our understanding of the transcriptional and regulatory mechanisms of baicalin biosynthesis, as well as sheds new light on metabolic engineering in S. baicalensis. Full article
Show Figures

Figure 1

Figure 1
<p>Expression profiles of <span class="html-italic">SbMYB12</span> in different tissues of <span class="html-italic">S. baicalensis</span>. The data represent the means of three biological replicates, and the error bars indicate the standard deviation (SD). Significant differences between means were identified (depicted by different letters; <span class="html-italic">p</span> &lt; 0.05) using one-way ANOVA (followed by Tukey’s comparisons).</p>
Full article ">Figure 2
<p>Prediction of <span class="html-italic">Sb</span>MYB12 protein structure and domains. (<b>A</b>) Predicted protein secondary structure. (<b>B</b>) Predicted protein domains. (<b>C</b>) Predicted tertiary structure.</p>
Full article ">Figure 3
<p>Analysis of <span class="html-italic">Sb</span>MYB12 compared with related sequences. (<b>A</b>) <span class="html-italic">Sb</span>MYB12 gene phylogenetic analysis along with 17 representatives of the R2R3-MYB S20 subgroup. (<b>B</b>) Analysis of sequence alignments between <span class="html-italic">Sb</span>MYB12 and other R2R3-MYB S20 subgroup proteins from different plants. Red frames indicate conserved R2 and R3 domains, and blue frames indicate the WxPRL core sequence.</p>
Full article ">Figure 4
<p>Subcellular localization and transactivation activities of the <span class="html-italic">Sb</span>MYB12 protein. (<b>A</b>) Subcellular location of <span class="html-italic">Sb</span>MYB12 in <span class="html-italic">Arabidopsis</span> mesophyll protoplasts. (<b>B</b>) Transcriptional activation activity analysis of BD-<span class="html-italic">Sb</span>MYB12 in yeast (Scale bar: 8mm).</p>
Full article ">Figure 5
<p>Identification of transgenic hairy roots of <span class="html-italic">S. baicalensis</span>. (<b>A</b>) Transgenic hairy root lines and observations with red fluorescent protein. A4: ArA4 lines (negative control); EV: ArA4 strain that harbors the pK7WG2R plasmids (positive control); <span class="html-italic">MYB12</span>-1~3: <span class="html-italic">SbMYB12</span> overexpression hairy root lines. (<b>B</b>) PCR identification of transgenic hairy roots of <span class="html-italic">S. baicalensis</span>. PCR screening of <span class="html-italic">SbMYB12</span> overexpressing lines. M: DNA marker (DL2000); P: ArA4 strain containing recombinant plasmid (positive control); N: <span class="html-italic">S. baicalensis</span> sterile plantlet (negative control); W: wild seedling of <span class="html-italic">S. baicalensis</span>. (<b>C</b>) Real-time quantitative PCR analysis of <span class="html-italic">SbMYB12</span> in transgenic lines. CK1-3: ArA4 lines; OE1–3: <span class="html-italic">SbMYB12</span> overexpression transgenic line; ** represents significant difference (<span class="html-italic">p</span> &lt; 0.01) via Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 6
<p><span class="html-italic">SbMYB12</span> promoted the accumulation of flavonoids (baicalin and wogonoside) in <span class="html-italic">S. baicalensis</span> transgenic hairy roots. (<b>A</b>) Phenotypes of the two-month-old <span class="html-italic">SbMYB12</span>-overexpression (OE) lines and the control (CK). (<b>B</b>) Concentrations of flavonoids in the transgenic hairy roots and the control (CK). Significant differences between the OE lines and the CK were identified (depicted by * <span class="html-italic">p</span> &lt; 0.05) via Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 7
<p>Expression analyses of enzyme genes for the flavonoid biosynthesis pathway in <span class="html-italic">SbMYB12</span> overexpressed transgenic hairy roots. (<b>A</b>) Proposed biosynthetic pathway for flavonoids in <span class="html-italic">S. baicalensis</span>. (<b>B</b>) Flavonoid biosynthesis pathway gene expression analysis in S<span class="html-italic">bMYB12</span> overexpressed strain. Significant differences between the OE lines and the CK were identified (depicted by * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01) via Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 8
<p><span class="html-italic">SbMYB12</span> binds to the <span class="html-italic">SbCCL7-4</span>, <span class="html-italic">SbCHI-2</span>, and <span class="html-italic">SbF6H-1</span> promoters. (<b>A</b>) Distribution of the MYB-binding sites in the <span class="html-italic">SbCCL7-4</span>, <span class="html-italic">SbCHI-2</span>, and <span class="html-italic">SbF6H-1</span> promoters. (<b>B</b>) Y1H assays indicated interactions between <span class="html-italic">SbMYB12</span> and the promoter regions of <span class="html-italic">SbCCL7-4</span>, <span class="html-italic">SbCHI-2</span>, and <span class="html-italic">SbF6H-1</span>. MRE: AACCTAA; MBS1: CAACTG; MBS3: TAACTG. (Scale bar: 6 mm).</p>
Full article ">
15 pages, 3769 KiB  
Article
Transcription Factor SmSPL2 Inhibits the Accumulation of Salvianolic Acid B and Influences Root Architecture
by Xiangzeng Wang, Yao Cao, Jiaxin Yang, Tong Zhang, Qianqian Yang, Yanhua Zhang, Donghao Wang and Xiaoyan Cao
Int. J. Mol. Sci. 2022, 23(21), 13549; https://doi.org/10.3390/ijms232113549 - 4 Nov 2022
Cited by 5 | Viewed by 1590
Abstract
The SQUAMOSA PROMOTER-BINDING PROTEIN-LIKE (SPL) transcription factor play vital roles in plant growth and development. Although 15 SPL family genes have been recognized in the model medical plant Salvia miltiorrhiza Bunge, most of them have not been functionally characterized to date. Here, we [...] Read more.
The SQUAMOSA PROMOTER-BINDING PROTEIN-LIKE (SPL) transcription factor play vital roles in plant growth and development. Although 15 SPL family genes have been recognized in the model medical plant Salvia miltiorrhiza Bunge, most of them have not been functionally characterized to date. Here, we performed a careful characterization of SmSPL2, which was expressed in almost all tissues of S. miltiorrhiza and had the highest transcriptional level in the calyx. Meanwhile, SmSPL2 has strong transcriptional activation activity and resides in the nucleus. We obtained overexpression lines of SmSPL2 and rSmSPL2 (miR156-resistant SmSPL2). Morphological changes in roots, including longer length, fewer adventitious roots, decreased lateral root density, and increased fresh weight, were observed in all of these transgenic lines. Two rSmSPL2-overexpressed lines were subjected to transcriptome analysis. Overexpression of rSmSPL2 changed root architectures by inhibiting biosynthesis and signal transduction of auxin, while triggering that of cytokinin. The salvianolic acid B (SalB) concentration was significantly decreased in rSmSPL2-overexpressed lines. Further analysis revealed that SmSPL2 binds directly to the promoters of Sm4CL9, SmTAT1, and SmPAL1 and inhibits their expression. In conclusion, SmSPL2 is a potential gene that efficiently manipulate both root architecture and SalB concentration in S. miltiorrhiza. Full article
Show Figures

Figure 1

Figure 1
<p>Expression profiles of <span class="html-italic">SmSPL2</span>. (<b>A</b>) Expression level of <span class="html-italic">SmSPL2</span> in different tissues. Different letters indicate significant differences between means at the <span class="html-italic">p</span> &lt; 0.05 level by one-way ANOVA test (followed by Tukey’s comparisons). (<b>B</b>–<b>J</b>) GUS staining signals of transgenic Arabidopsis expressing SmSPL2pro::GUS. (<b>B</b>–<b>D</b>) Seedlings after germination at two (<b>B</b>), five (<b>C</b>), and nine (<b>D</b>) days. (<b>E</b>–<b>I</b>) Stem leaf (<b>E</b>), rosette leaf (<b>F</b>), stem (<b>G</b>), flower (<b>H</b>), and root (<b>I</b>) at the flowering stage. (<b>J</b>) Silique.</p>
Full article ">Figure 2
<p>Subcellular location and transactivation activity analysis of SmSPL2 protein. (<b>A</b>) Subcellular location of SmSPL2 in onion epidermal cells. The fluorescence signal was observed via a laser scanning confocal microscope. DAPI, 4′,6-diamidino-2-phenylindole is a blue-fluorescent DNA stain used to indicate the nucleus region. (<b>B</b>) Transcriptional activation analysis of SmSPL2 protein. Yeast colonies with three different dilutions were cultured on SD/-Trp and SD/-Trp/-Ade/-His/X-α-gal media. X-α-gal: 5-Bromo-4-chloro-3-indolyl-α-D-galactoside medium, the color reaction substrate of α-galactosidase.</p>
Full article ">Figure 3
<p>Phenotype of transgenic <span class="html-italic">S. miltiorrhiza</span> plantlets overexpressing SmSPL2 (SmSPL2-OE) and miR156-resistant SmSPL2 (rSmSPL2-OE). (<b>A</b>–<b>C</b>) Phenotype of 15-day-old (<b>A</b>), one-month-old (<b>B</b>), and two-month-old (<b>C</b>) transgenic <span class="html-italic">S. miltiorrhiza</span> plants. (<b>D</b>) Expression levels of SmSPL2 in the roots of different transgenic lines. (<b>E</b>–<b>G</b>) Adventitious root (AR) length (<b>E</b>), AR numbers (<b>F</b>), fresh weight of root (<b>G</b>) of two-month-old transgenic lines. All data are mean values of three biological replicates, with the error bar representing SD. * and ** indicate significant differences from the control at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 level, respectively, by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>Changes in transcriptional levels of genes for cytokinin and auxin biosynthesis and signal transduction. (<b>A</b>) Proposed biosynthetic and signal transduction pathway for cytokinin. DMAPP, dimethylallyl pyrophosphate; IPT, isopentenyl transferases; CYP, cytochrome P450 enzyme; LOG, LONELY GUY; AHPs, histidine phosphotransfer proteins; CRF, cytokinin response factors; B-AAR, type-B response regulators; A-AAR, type-A response regulators. (<b>B</b>) DEGs involved in cytokinin biosynthesis and the signal transduction pathway. For each gene, the relative expression (<span class="html-italic">rSmSPL2</span>-OE2 vs. control and <span class="html-italic">rSmSPL2</span>-OE7 vs. control) was expressed in Log<sub>2</sub>FC. Blue boxes (gene expression is down-regulated); Red boxes (gene expression is up-regulated). * indicates differentially expressed genes. (<b>C</b>) Concentrations of cytokinin in root extracts from control and transgenic line <span class="html-italic">rSmSPL2</span>-OE2 and <span class="html-italic">rSmSPL2</span>-OE7. *, values are significantly different from control at <span class="html-italic">p</span> &lt; 0.05 by Student’s <span class="html-italic">t</span>-test. (<b>D</b>) Proposed biosynthetic and signal transduction pathways for auxin. TAA, Tryptophan aminotransferase of Arabidopsis; TAR, Tryptophan aminotransferase related; YUC, YUCCA; AIM, Indole-3-acetamide hydrolase; AAO, Aldehyde oxidase; TIR1, Transport Inhibitor Response 1; AUX/IAA, auxin/indole acetic acid; GH3, Gretchen Hagen 3; SAUR, small auxin-up RNA. (<b>E</b>) DEGs involved in auxin biosynthesis and signal transduction pathway. For each gene, the relative expression (<span class="html-italic">rSmSPL2</span>-OE2 vs. control and <span class="html-italic">rSmSPL2</span>-OE7 vs. control) was expressed in Log<sub>2</sub>FC. Blue boxes (gene expression is down-regulated); Red boxes (gene expression is up-regulated). * indicates differentially expressed genes. Gene ID were listed in <a href="#app1-ijms-23-13549" class="html-app">Table S2</a>. (<b>F</b>) Concentrations of auxin in root extracts from control and transgenic line <span class="html-italic">rSmSPL2</span>-OE2 and <span class="html-italic">rSmSPL2</span>-OE7. **, values are significantly different from control at <span class="html-italic">p</span> &lt; 0.01 by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 5
<p>Overexpression of miR156-resistant <span class="html-italic">SmSPL2</span> (<span class="html-italic">rSmSPL2</span>) decreased the accumulation of rosmarinic acid (RA) and salvianolic acid B (SalB). (<b>A</b>) Transcriptome analysis of enzyme genes for salvianolic acid biosynthesis pathway in transgenic lines (<span class="html-italic">rSmSPL2</span>-OE2 and <span class="html-italic">rSmSPL2</span>-OE7). For each gene, the relative expression (<span class="html-italic">rSmSPL2</span>-OE2 vs. control and <span class="html-italic">rSmSPL2</span>-OE7 vs. control) was expressed as Log<sub>2</sub>FC. Blue boxes (gene expression is down-regulated); Red boxes (gene expression is up-regulated). TAT, tyrosine aminotransferase; HPPR, hydroxyl phenylpyruvate reductase; PAL, phenylalanine ammonia lyase; C4H, cinnamate 4-hydroxylase; 4CL, hydroxycinnamate-CoA ligase; RAS, rosmarinic acid synthase; and CYP, cytochrome P450 enzymes. (<b>B</b>) Validation by qRT-PCR of six enzyme genes involved in salvianolic acid biosynthesis in transgenic lines and control. (<b>C</b>) Chromatograms of RA and SalB in transgenic lines and the control. (<b>D</b>) RA and SalB concentrations in the roots of two-month-old transgenic lines and the control. All data are mean values of three biological replicates, with the error bar representing SD. * and ** indicate significant difference from the control at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 level, respectively, by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 6
<p><span class="html-italic">Sm</span>SPL2 binds to the promoter regions of <span class="html-italic">Sm4CL9</span>, <span class="html-italic">SmTAT1</span>, and <span class="html-italic">SmPAL1</span> and inhibits their expression. (<b>A</b>) GTAC motifs in the promoter regions of <span class="html-italic">Sm4CL9</span>, <span class="html-italic">SmCYP98A14</span>, <span class="html-italic">SmTAT1</span>, and <span class="html-italic">SmPAL1</span>. Black rectangles represent the GTAC motif. (<b>B</b>) Yeast one-hybrid detected interactions between the <span class="html-italic">Sm</span>SPL6 and the promoters of <span class="html-italic">Sm4CL9</span>, <span class="html-italic">SmCYP98A14</span>, <span class="html-italic">SmTAT1</span>, and <span class="html-italic">SmPAL1</span>. The p53HIS2/pGADT7-p53 and p53HIS2/pGADT7 were the positive and negative controls, respectively. (<b>C</b>) Schematic diagram of constructs used in assays of transient transcriptional activity. (<b>D</b>) <span class="html-italic">Sm</span>SPL2 inhibited the expression of <span class="html-italic">Sm4CL9</span>, <span class="html-italic">SmTAT1</span>, and <span class="html-italic">SmPAL1</span>. Effector <span class="html-italic">Sm</span>SPL2 was co-transformed with reporter <span class="html-italic">p4CL9</span>-LUC, <span class="html-italic">pTAT1</span>-LUC, or <span class="html-italic">pPAL1</span>-LUC. All data are mean values of three biological replicates, with the error bar representing SD. * and ** indicate significant differences from the control at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 levels, respectively, by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">
18 pages, 4844 KiB  
Article
Integrating Metabolomics and Transcriptomics to Unveil Atisine Biosynthesis in Aconitum gymnandrum Maxim
by Lingli Chen, Mei Tian, Baolong Jin, Biwei Yin, Tong Chen, Juan Guo, Jinfu Tang, Guanghong Cui and Luqi Huang
Int. J. Mol. Sci. 2022, 23(21), 13463; https://doi.org/10.3390/ijms232113463 - 3 Nov 2022
Cited by 4 | Viewed by 2142
Abstract
Diterpene alkaloids (DAs) are characteristic compounds in Aconitum, which are classified into four skeletal types: C18, C19, C20, and bisditerpenoid alkaloids. C20-DAs are thought to be the precursor of the other types. Their biosynthetic [...] Read more.
Diterpene alkaloids (DAs) are characteristic compounds in Aconitum, which are classified into four skeletal types: C18, C19, C20, and bisditerpenoid alkaloids. C20-DAs are thought to be the precursor of the other types. Their biosynthetic pathway, however, is largely unclear. Herein, we combine metabolomics and transcriptomics to unveil the methyl jasmonate (MJ) inducible biosynthesis of DAs in the sterile seedling of A. gymnandrum, the only species in the Subgenus Gymnaconitum (Stapf) Rapaics. Target metabolomics based on root and aerial portions identified 51 C19-DAs and 15 C20-DAs, with 40 inducible compounds. The highest content of C20-DA atisine was selected for further network analysis. PacBio Isoform sequencing integrated with RNA sequencing not only provided the full-length transcriptome but also their response to induction, revealing 1994 genes that exhibited up-regulated expression. Further, 38 genes involved in terpenoid biosynthesis were identified, including 7 diterpene synthases. In addition to the expected function of the four diterpene synthases, AgCPS5 was identified to be a new ent-8,13-CPP synthase in Aconitum and could also combine with AgKSL1 to form the C20-DAs precursor ent-atiserene. Combined with multiple network analyses, six CYP450 and seven 2-ODD genes predicted to be involved in the biosynthesis of atisine were also identified. This study not only sheds light on diterpene synthase evolution in Aconitum but also provides a rich dataset of full-length transcriptomes, systemic metabolomes, and gene expression profiles, setting the groundwork for further investigation of the C20-DAs biosynthesis pathway. Full article
Show Figures

Figure 1

Figure 1
<p>The production of Das is a major metabolic response of <span class="html-italic">A. gymnandrum</span> sterile seedlings to induction. MJ and MJL refer to the sample of root and aerial portions. (<b>a</b>) Principal component analysis of <span class="html-italic">A. gymnandrum of different plant portions</span> shows compounds with peak accumulation occurring at the indicated time, demonstrating that MJ has a greater impact on the root than aerial portions. (<b>b</b>) Principal component analysis of 77 metabolites in the root; the five compounds most representative of the first principal component are atisine, compound <b>45</b>, compound <b>56</b>, compound <b>63</b>, and compound <b>66</b>. (<b>c</b>) Cluster analysis of 40 metabolites in the root from <span class="html-italic">A. gymnandrum</span> at different induction times. The value of each compound was taken from relative content Log<sub>2</sub> standardization. (<b>d</b>) Plots to demonstrate the increasing accumulation of the atisine in the root (error bars represent the standard error). Untreated and treated refer to the control group and experimental group, respectively.</p>
Full article ">Figure 2
<p>Venn diagram of up-regulated DEGs by different induction groups in the <span class="html-italic">A. gymnandrum</span> root.</p>
Full article ">Figure 3
<p>The expression profile for the genes involved in terpenoid biosynthesis in the root of <span class="html-italic">A. gymnandrum</span>. The transcriptome expression of each identified gene is Log2 normalization of fpkm values. Abbreviations: AACT, aceto-acetyl-CoA thiolase; CMK, 4-(cytidine 50-diphospho)-2-C-methyl-D-erythritol kinase; DXS, 1-deoxy-D-xylulose 5-phosphate synthase; DXR, 1-deoxy-D-xylulose-5-phosphate reductoisomerase; FPS, farnesyl pyrophosphate synthase; HDR, (E)-4-hydroxy-3-methylbut-2-enyl diphosphate reductase; HDS, (E)-4-hydroxy-3-methylbut-2-enyl diphosphate synthase; HMGS, 3-hydroxy-3-methylglutaryl-CoA synthase; HMGR, 3-hydroxy-3-methylglutaryl-CoA reductase; GGPPS, geranylgeranyl pyrophosphate synthase; MCT, 2-C-methyl-D-erythritol 4-phosphate cytidylyltransferase; IDI, isopentenyl diphosphate isomerase; MDS, 2-C-methyl-D-erythritol 2,4-cyclodiphosphate synthase; MDD, mevalonate diphosphate decarboxylase; MDS, 2-C-methyl-D-erythritol 2,4-cyclodiphosphate synthase; MVK, mevalonate kinase; PMK, phosphomevalonate kinase; TPS, terpene synthases (including monoterpene synthases and sesquiterpene synthases); CPS, copalyl diphosphate synthase; KSL, kaurene synthase.</p>
Full article ">Figure 4
<p>GC-MS analysis of AgCPSs and AgKSL1 reaction products obtained from in vitro assays. (<b>a</b>) 1~5: The product of AgCPS1/2/4/5 and ZmCPS2 (specific to <span class="html-italic">ent</span>-CPP) enzymatic reaction with GGPP, it is proved that the products of AgCPS1/2/4 and AgCPS 5 were CPP and 8,13-CPP; 6~8: The product of AgCPS1/2/4/5 and ZmCPS2 combined with AtKS (specific to <span class="html-italic">ent</span>-kaurene), it is proved that the products of AgCPS1/2/4 were the <span class="html-italic">ent</span>-configuration; 9~12: the product of AgCPS1/2/4/5 and ZmCPS2 combined with AgKSL1/IrKSL4 (specific to <span class="html-italic">ent</span>-atiserene), it is proved that the products of AgKSL1 was <span class="html-italic">ent</span>-atiserene. The enzymes ZmCPS2, PcTPS1, IrKSL4, and AtKS were used as controls. (<b>b</b>) Extracted ion chromatograms (EIC) of the product in different combinations above. Peak1: <span class="html-italic">ent</span>-CPP; Peak2: <span class="html-italic">ent</span>-8,13-CPP; Peak3: <span class="html-italic">ent</span>-kaurene; Peak4: <span class="html-italic">ent</span>-atiserene.</p>
Full article ">Figure 5
<p>Identification of co-expression network modules in <span class="html-italic">A. gymnandrum</span>. (<b>a</b>) Gene dendrogram was obtained by clustering the dissimilarity based on consensus topological overlap with the corresponding module colors indicated by the color row. (<b>b</b>) Modular eigenvector clustering heatmap. The heatmap shows the relatedness of the 6 co-expression modules identified in WGCNA, with red indicating highly related and blue indicating not related. (<b>c</b>) Module-atisine correlation heatmap. The right transverse panel with red-blue indicates a color scale for module–trait correlation, from –0.5 to 0.5.</p>
Full article ">Figure 6
<p>Co-expression analysis in module green of <span class="html-italic">A. gymnandrum.</span> Transcript 45862, transcript 49735, transcript 49882, transcript 19239, transcript 65043, transcript 42527, transcript 29295, transcript 33670, and transcript 49129 were filtered as hub genes by degree. The circle depicts genes co-expressed with 10 hub genes (r &gt; 0.8; 77 genes in total). The size and color (red to green) of the circle were arranged according to the degree, and the lines indicate the <span class="html-italic">p</span>-value between the genes (purple is the largest and yellow is the smallest).</p>
Full article ">Figure 7
<p>The predicted biosynthetic pathway of atisine and the involved enzyme genes in <span class="html-italic">A. gymnandrum</span>, red colors represent functional qualification, blue colors were screened from transcripts, and orange colors were filtered in the module green. Abbreviations: KO: <span class="html-italic">ent</span>-kaurene oxidase, SDC: serine decarboxylase, TA: transaminase.</p>
Full article ">
18 pages, 2698 KiB  
Article
Plant Metabolic Engineering by Multigene Stacking: Synthesis of Diverse Mogrosides
by Jingjing Liao, Tingyao Liu, Lei Xie, Changming Mo, Xiyang Huang, Shengrong Cui, Xunli Jia, Fusheng Lan, Zuliang Luo and Xiaojun Ma
Int. J. Mol. Sci. 2022, 23(18), 10422; https://doi.org/10.3390/ijms231810422 - 9 Sep 2022
Cited by 6 | Viewed by 2940
Abstract
Mogrosides are a group of health-promoting natural products that extracted from Siraitia grosvenorii fruit (Luo-han-guo or monk fruit), which exhibited a promising practical application in natural sweeteners and pharmaceutical development. However, the production of mogrosides is inadequate to meet the need worldwide, and [...] Read more.
Mogrosides are a group of health-promoting natural products that extracted from Siraitia grosvenorii fruit (Luo-han-guo or monk fruit), which exhibited a promising practical application in natural sweeteners and pharmaceutical development. However, the production of mogrosides is inadequate to meet the need worldwide, and uneconomical synthetic chemistry methods are not generally recommended for structural complexity. To address this issue, an in-fusion based gene stacking strategy (IGS) for multigene stacking has been developed to assemble 6 mogrosides synthase genes in pCAMBIA1300. Metabolic engineering of Nicotiana benthamiana and Arabidopsis thaliana to produce mogrosides from 2,3-oxidosqualene was carried out. Moreover, a validated HPLC-MS/MS method was used for the quantitative analysis of mogrosides in transgenic plants. Herein, engineered Arabidopsis thaliana produced siamenoside I ranging from 29.65 to 1036.96 ng/g FW, and the content of mogroside III at 202.75 ng/g FW, respectively. The production of mogroside III was from 148.30 to 252.73 ng/g FW, and mogroside II-E with concentration between 339.27 and 5663.55 ng/g FW in the engineered tobacco, respectively. This study provides information potentially applicable to develop a powerful and green toolkit for the production of mogrosides. Full article
Show Figures

Figure 1

Figure 1
<p>The Flow Chart of mogrosides biosynthetic pathway in <span class="html-italic">Siraitia grosvenorii</span> fruits. Mogrosides synthase genes which were transformed in this study are marked in blue, including <span class="html-italic">SgSQE1</span>, squalene epoxidases; <span class="html-italic">SgCS</span>, curbitadienol synthase; <span class="html-italic">SgEPH2</span>, epoxide hydrolases; <span class="html-italic">SgP450</span>, cytochrome P450 mono-oxygenase; <span class="html-italic">SgUGT269-1</span> and <span class="html-italic">SgUGT289-3</span>, UDP-glucosyltransferases; The substrate, 2,3-oxidosqualene is shown with a yellow background. MI-A1, mogroside I-A1; MII-E, mogroside II-E; MIII, mogroside III; SI, siamenoside I; MV, mogroside V (in the dotted bordered rectangle).</p>
Full article ">Figure 2
<p>Molecular analysis of transgenic tobacco lines. (<b>A</b>) PCR-based analysis of the transgenic tobacco lines. The lanes from left to right represent the WT, N16, N22, N30, N31, N32, N45, N46, N47, and N48. An image of the DNA marker (4.5 kb) is shown in the upper-left corner of the figure. (<b>B</b>) Relative expression level analysis of 6 mogrosides biosynthesis genes in transgenic tobacco lines (N16, N22, N30, N32, N45, N47). The <span class="html-italic">Nbactin</span> is used as an internal control. Expression of tobacco WT plants was set to 1. The data are presented as the mean values ± SDs, <span class="html-italic">n</span> = 3 biologically independent samples, ** represents significant difference at <span class="html-italic">p</span> &lt; 0.01 (LSD test).</p>
Full article ">Figure 3
<p>Molecular analysis of transgenic <span class="html-italic">Arabidopsis</span> lines. (<b>A</b>) <span class="html-italic">Arabidopsis</span> WT plants and transgenic lines. (<b>B</b>) PCR-based analysis of transgenic <span class="html-italic">Arabidopsis</span> lines. The lanes from left to right represent the WT, AA3, AA5, AA6, AA7, AU6, AU7, AU8, AU10, AU11, AU12, and AU13. DNA marker (4500 bp) is used in the figure. (<b>C</b>) RT-PCR detection of 6 mogrosides biosynthesis genes in the WT and transgenic <span class="html-italic">Arabidopsis</span> lines AA3, AA5, AA6, AU7, AU10, AU11, A12 (from left to right). The <span class="html-italic">Atactin</span> is used as an internal control.</p>
Full article ">Figure 4
<p>Production of mogrosides in transgenic tobacco lines. (<b>A</b>) HPLC-MS/MS analysis of MIII and MII-E in transgenic tobacco lines. (<b>B</b>) Accumulation of MIII in transgenic tobacco lines. (<b>C</b>) Accumulation of MII-E in transgenic tobacco lines. n.d., not detected. The data are presented as the mean values ± SDs, <span class="html-italic">n</span> = 3 biologically independent samples. The black arrows indicate the peak of MIII and MIIE. (<b>D</b>) Full-scan product ion of MIII and MII-E.5.0 × 10<sup>4</sup>.</p>
Full article ">Figure 4 Cont.
<p>Production of mogrosides in transgenic tobacco lines. (<b>A</b>) HPLC-MS/MS analysis of MIII and MII-E in transgenic tobacco lines. (<b>B</b>) Accumulation of MIII in transgenic tobacco lines. (<b>C</b>) Accumulation of MII-E in transgenic tobacco lines. n.d., not detected. The data are presented as the mean values ± SDs, <span class="html-italic">n</span> = 3 biologically independent samples. The black arrows indicate the peak of MIII and MIIE. (<b>D</b>) Full-scan product ion of MIII and MII-E.5.0 × 10<sup>4</sup>.</p>
Full article ">Figure 5
<p>Production of mogrosides in transgenic <span class="html-italic">Arabidopsis</span> lines. (<b>A</b>) HPLC-MS/MS analysis of MIII in transgenic <span class="html-italic">Arabidopsis</span> lines. (<b>B</b>) HPLC-MS/MS analysis of SI in transgenic <span class="html-italic">Arabidopsis</span> lines. The black arrows indicate the peak of MIII and SI. (<b>C</b>) Full-scan product ion of SI.</p>
Full article ">Figure 6
<p>Accumulation of mogrosides in transgenic <span class="html-italic">Arabidopsis</span> lines AA3, AA6, and AU7. n.d., not detected. The data are presented as the mean values ± SDs, <span class="html-italic">n</span> = 3 biologically independent samples.</p>
Full article ">
Back to TopTop