[go: up one dir, main page]

 
 

Abiotic Stress Responses in Ornamental Crops: The State of the Art 2024

A special issue of Horticulturae (ISSN 2311-7524). This special issue belongs to the section "Floriculture, Nursery and Landscape, and Turf".

Deadline for manuscript submissions: closed (26 July 2024) | Viewed by 9151

Special Issue Editors

Key Laboratory for Quality Regulation of Tropical Horticultural Crops of Hainan Province, School of Tropical Agriculture and Forestry (School of Agricultural and Rural Affairs, School of Rural Revitalization), Hainan University, Haikou 570228, China
Interests: color formation of flowers and leaves; transcription regulation; gene expression and protein interactions
Special Issues, Collections and Topics in MDPI journals

E-Mail
Guest Editor
Key Laboratory of Tree Breeding of Zhejiang Province, Research Institute of Subtropical Forestry, Chinese Academy of Forestry, Hangzhou 311400, China
Interests: flavonoids metabolism; flower color; transcription regulation; gene expression and protein interactions
Special Issues, Collections and Topics in MDPI journals

E-Mail
Guest Editor
College of Landscape Architecture and Art, Fujian Agriculture and Forestry University, Fuzhou 350002, China
Interests: genetic breeding; molecular biology; flowering regulation

Special Issue Information

Dear Colleagues,

Abiotic stresses, such as high temperatures, cold, drought and salt, are important factors affecting the yield and quality of ornamental crops. Improving the stress resistance of ornamental crops is an important goal of breeding, and it is necessary for scientific research to serve production. Therefore, the study of the resistance mechanisms of ornamental crops and the use of the latest molecular biology technology to uncover resistance genes is of great importance for improving the production quality of ornamental crops and breeding new resistant varieties.

The purpose of this Special Issue “Abiotic Stress Responses in Ornamental Crops: The State of the Art 2024” is to present the latest advances in the research of ornamental crops in response to abiotic stresses, including but not limited to physiological responses and molecular mechanisms. Any innovative articles on the abiotic stress responses of ornamental crops are welcome in this Special Issue.

Dr. Yang Zhou
Dr. Weixin Liu
Dr. Yunxiao Guan
Guest Editors

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. Horticulturae is an international peer-reviewed open access monthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2200 CHF (Swiss Francs). 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

  • abiotic stress
  • resistance breeding
  • physiological response
  • molecular mechanism
  • transcriptional regulation

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.

Published Papers (8 papers)

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

Research

11 pages, 1152 KiB  
Article
Effects of Steel Slag Used as Substrate on the Growth of Hydrangea macrophylla Cuttings
by Jundan Mao, Huijie Chen, Huimin Zhou, Xiangyu Qi, Shuangshuang Chen, Jing Feng, Yuyan Jin, Chang Li, Yanming Deng and Hao Zhang
Horticulturae 2024, 10(10), 1053; https://doi.org/10.3390/horticulturae10101053 - 2 Oct 2024
Viewed by 694
Abstract
Steel slag is an industrial solid waste produced during the steelmaking process. To explore the application of steel slag in the agricultural field, the present experiment was carried out to study the effect of substrates with different contents of steel slag on the [...] Read more.
Steel slag is an industrial solid waste produced during the steelmaking process. To explore the application of steel slag in the agricultural field, the present experiment was carried out to study the effect of substrates with different contents of steel slag on the growth of Hydrangea macrophylla cuttings. The conventional substrate (perlite: vermiculite: peat = 1:1:1) was used as the control (CK), and the treatments were designed as T1 (steel slag: perlite: vermiculite: peat = 1:3:3:3, v/v/v/v), T2 (steel slag: perlite: vermiculite: peat = 1:2:2:2, v/v/v/v), T3 (steel slag: perlite: vermiculite: peat = 1:1:1:1, v/v/v/v), and T4 (steel slag: perlite: vermiculite: peat = 1:0:0:0, v/v/v/v). The results showed that the addition of steel slag significantly increased the substrate’s bulk density, EC, and pH and improved its water retention capacity to a certain extent. There were significant differences among different treatments in morphological indicators, root growth and development, and physiological and biochemical characteristics of cutting seedlings. All traits, including plant height, fresh weight, dry weight, root length, root surface area, root volume, the number of root tips, root activity, and soluble protein content of seedlings grown in T3 were significantly higher than those in other substrates. The results indicated that the appropriate addition of steel slag is helpful to hydrangea cuttings’ growth, and the optimal mixing ratio is steel slag: perlite: vermiculite: peat = 1:1:1:1 (v/v/v/v). This is a significant innovation in applying steel slag in agricultural production. Full article
Show Figures

Figure 1

Figure 1
<p>Hydrangea seedlings grown in different substrates.</p>
Full article ">Figure 2
<p>The root systems of hydrangea seedlings grown in different substrates.</p>
Full article ">
14 pages, 2831 KiB  
Article
Genome-Wide Identification of DREB Transcription Factor Family and Functional Analysis of PaDREB1D Associated with Low-Temperature Stress in Phalaenopsis aphrodite
by Ziang Hu, Shuang Wang, Yaoling Wang, Jiaming Li, Ping Luo, Jingjing Xin and Yongyi Cui
Horticulturae 2024, 10(9), 933; https://doi.org/10.3390/horticulturae10090933 - 31 Aug 2024
Viewed by 435
Abstract
Low temperatures are the most significant abiotic stressor for the conservation and production of Phalaenopsis in non-tropical areas. CBF/DREB1 transcription factors play an important role in the plant abiotic stress response. In this study, 31 DREB family members were identified in the Phalaenopsis [...] Read more.
Low temperatures are the most significant abiotic stressor for the conservation and production of Phalaenopsis in non-tropical areas. CBF/DREB1 transcription factors play an important role in the plant abiotic stress response. In this study, 31 DREB family members were identified in the Phalaenopsis genome. Expression pattern analysis showed that the expression of different PaDREB members varied among tissue sites. PaDREB1D was isolated from Phalaenopsis aphrodite, and multiple sequence alignment showed that PaDREB1D belonged to the A1 subgroup of the DREB family and was localized in the nucleus. PaDREB1D overexpression in protocorm-like bodies of Phalaenopsis reduced cell damage during low-temperature stress, increased antioxidant enzyme activity, and enhanced the low-temperature tolerance of protocorm-like bodies. The results of this study provide a theoretical basis for breeding for cold resistance and investigating the molecular mechanisms related to low-temperature responses in Phalaenopsis. Full article
Show Figures

Figure 1

Figure 1
<p>Phylogenetic tree of the DREB family in <span class="html-italic">Arabidopsis thaliana</span> and <span class="html-italic">Phalaenopsis aphrodite</span>. The AtDREB sequences used are shown in <a href="#app1-horticulturae-10-00933" class="html-app">Table S2</a>.</p>
Full article ">Figure 2
<p><span class="html-italic">PaDREB</span> expression patterns in different tissues. F: flower; L: leaf; R: root; SS: short stalk; LS: long stalk; SB: small bud; LB: large bud. The transcript per kilobase per million mapped reads of <span class="html-italic">PaDREBs</span> were screened from the <span class="html-italic">Phalaenopsis aphrodite</span> transcriptome data.</p>
Full article ">Figure 3
<p>Expression pattern analysis of <span class="html-italic">PaDREB1D</span> under abiotic stress and abscisic acid (ABA) treatments. (<b>A</b>) Low-temperature treatment; (<b>B</b>) drought treatment; (<b>C</b>) salt treatment; (<b>D</b>) ABA treatment. Untreated <span class="html-italic">Phalaenopsis aphrodite</span> was used as a control (0 h). The data represent the mean ± standard deviation (n = 3). Different letters show significant differences (<span class="html-italic">p</span> &lt; 0.05) between samples.</p>
Full article ">Figure 4
<p>Subcellular localization of PaDREB1D. The 35S:<span class="html-italic">PaDREB1D</span>-GFP and 35S:GFP controls were transiently expressed in tobacco leaves. mCherry was chosen as a nuclear localization marker. Scale bar = 50 μm.</p>
Full article ">Figure 5
<p>Phenotypic changes in transgenic <span class="html-italic">PaDREB1D</span> protocorm-like bodies after low-temperature stress. BS: before stress, AS: after stress, R 1 d: 1 d of recovery cultivation, R 7 d: 7 d of recovery cultivation. PLBs that remained green at 7 d of recovery were considered alive; those that were whitened, yellowed, or browned were considered dead. Scale bar = 1 cm.</p>
Full article ">Figure 6
<p>Survival rate and cellular damage in protocorm-like bodies under low-temperature stress. (<b>A</b>) Survival rate, (<b>B</b>) chlorophyll content, (<b>C</b>) relative conductivity, and (<b>D</b>) MDA content of the control and transgenic <span class="html-italic">PaDREB1D</span> PLBs under low-temperature stress. BS: before stress, AS: after stress, R 1 d: 1 d of recovery cultivation, R 7 d: 7 d of recovery cultivation. The data represent the mean ± standard deviation (n = 3). Different letters show significant differences (<span class="html-italic">p</span> &lt; 0.05) between samples.</p>
Full article ">Figure 7
<p>Changes in the soluble protein content and antioxidant enzyme activity of control and transgenic <span class="html-italic">PaDREB1D</span> protocorm-like bodies before and after 4 d of low-temperature stress. (<b>A</b>) Soluble protein content; (<b>B</b>) SOD activity; (<b>C</b>) POD activity; (<b>D</b>) CAT activity. BS: before stress, AS: after stress, R 1 d: 1 d of recovery cultivation, R 7 d: 7 d of recovery cultivation. The data represent the mean ± standard deviation (n = 3). Different letters show significant differences (<span class="html-italic">p</span> &lt; 0.05) between samples.</p>
Full article ">
18 pages, 9936 KiB  
Article
Integrated Transcriptome and Metabolome to Elucidate the Mechanism of Aluminum-Induced Blue-Turning of Hydrangea Sepals
by Wenfang Li, Penghu Lei, Tingting Zhu, Huijun Zhang, Hui Jiang and Haixia Chen
Horticulturae 2024, 10(7), 745; https://doi.org/10.3390/horticulturae10070745 - 15 Jul 2024
Viewed by 751
Abstract
Hydrangea macrophylla is an ornamental plant with varied calyx colors. Interestingly, from red, to purple, to blue, the colors of all Hydrangea macrophylla are formed by unique delphinidin-3-O-glucoside and aluminum ions (Al3+) and 5-O-p-coumaroylquinic acid. The sepals of ‘Blue Mama’ changed [...] Read more.
Hydrangea macrophylla is an ornamental plant with varied calyx colors. Interestingly, from red, to purple, to blue, the colors of all Hydrangea macrophylla are formed by unique delphinidin-3-O-glucoside and aluminum ions (Al3+) and 5-O-p-coumaroylquinic acid. The sepals of ‘Blue Mama’ changed from pink to blue, and the contents of delphinidin-3-O-glucoside and aluminum ions increased under 3 g/L aluminum sulfate treatment. However, the mechanism of the effect of aluminum ions on the synthesis and metabolism of anthocyanins in Hydrangea macrophylla is still unclear. In this project, transcriptome sequencing and anthocyanin metabolome analysis were performed on the sepals of ‘Blue Mama’ during flower development at the bud stage (S1), discoloration stage (S2) and full-bloom stage (S3) under aluminum treatment. It was found that delphinidin, delphinidin-3-O-glucoside and delphinidin-3-O-galactoside were the main differential metabolites. The structural genes CHS, F3H, ANS, DFR and BZI in the anthocyanin synthesis pathway were up-regulated with the deepening in sepal color. There was no significant difference between the aluminum treatment and the non-aluminum treatment groups. However, seven transcription factors were up-regulated and expressed to regulate anthocyanin synthesis genes CHS, F3H, BZI and 4CL, promoting the sepals to turn blue. The KEGG enrichment pathway analysis of differentially expressed genes showed that the glutathione metabolism and the ABC transporter pathway were closely related to anthocyanin synthesis and aluminum-ion transport. GST (Hma1.2p1_0158F.1_g069560.gene) may be involved in the vacuolar transport of anthocyanins. The expression of anthocyanin transporter genes ABCC1 (Hma1.2p1_0021F.1_g014400.gene), ABCC2 (Hma1.2p1_0491F.1_g164450.gene) and aluminum transporter gene ALS3 (Hma1.2p1_0111F.1_g053440.gene) were significantly up-regulated in the aluminum treatment group, which may be an important reason for promoting the transport of anthocyanin and aluminum ions to vacuoles and making the sepals blue. These results preliminarily clarified the mechanism of aluminum ion in the synthesis and transport of anthocyanin in Hydrangea macrophylla, laying a foundation for the further study of the formation mechanism of ‘blue complex’ in Hydrangea macrophylla. Full article
Show Figures

Figure 1

Figure 1
<p>Changes in flower color and aluminum-ion content of ‘Blue Mama’. (<b>A</b>) Changes in flower color during the development of ‘Blue Mama’ under different concentrations of aluminum treatment. (<b>B</b>) Changes in the content of delphinidin-3-O-glucoside in sepals under different concentrations of aluminum treatment. (<b>C</b>) Aluminum-ion content in root, stem, leaf and sepals of plants at S3 stage. **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Statistical analysis of anthocyanin metabolites. (<b>A</b>) The type and quantity of metabolites identified. (<b>B</b>) The number of differential metabolites in different comparison groups. (<b>C</b>) Comparative analysis of the top 10 anthocyanin metabolites in terms of content in the S2 stage.</p>
Full article ">Figure 3
<p>Sequencing results statistics. (<b>A</b>) Statistical analysis of differential genes between groups. (<b>B</b>) The differential gene Venn diagram within and between groups.</p>
Full article ">Figure 4
<p>The expression profiles of significantly enriched differential genes and their KEGG pathway enrichment analysis. The above profiles significantly represent enriched expression patterns of differential genes, with the broken lines indicating the trend in gene expression across the three stages of flower development (<span class="html-italic">p</span> &lt; 0.05). Below each cluster, the top 10 most significantly enriched KEGG pathways are represented by a histogram based on the adjusted <span class="html-italic">p</span>-value, and the number of genes in each pathway is represented by a line chart. (<b>A</b>,<b>B</b>) are the CK group and the Al treatment group, respectively.</p>
Full article ">Figure 5
<p>Comparative analysis of gene expression and metabolites in anthocyanin biosynthesis pathway at different stages of flower development. (<b>A</b>) Comparative analysis of structural gene expression and metabolite content. (<b>B</b>) Correlation analysis between structural genes and metabolites. Only Pearson correlation coefficient |PCC| ≥ 0.90, <span class="html-italic">p</span> &lt; 0.05. Red dashed line, positive correlation between gene and metabolite; blue dashed line, negative correlation between gene and metabolite.</p>
Full article ">Figure 6
<p>Transcription factor analysis. (<b>A</b>) The number of top 15 transcription factor families identified. (<b>B</b>) Heatmap of NAC transcription factor. (<b>C</b>) Heatmap of bHLH transcription factor. (<b>D</b>) Heatmap of MYB transcription factor. (<b>E</b>) Co-expression map of NAC and anthocyanin synthesis structural genes. (<b>F</b>) Co-expression map of bHLH and anthocyanin synthesis structural genes. (<b>G</b>) Co-expression map of MYB and anthocyanin synthesis structural genes. Only Pearson correlation coefficient |PCC| ≥ 0.90, <span class="html-italic">p</span> &lt; 0.05. Red dashed line, positive correlation between genes; blue dashed line, negative correlation between genes.</p>
Full article ">Figure 7
<p>Expression analysis of anthocyanin transporter gene.</p>
Full article ">Figure 8
<p>Expression analysis of aluminum transport-related genes.</p>
Full article ">Figure 9
<p>Validation and expression analysis of selected genes using qRT-PCR.</p>
Full article ">
11 pages, 2923 KiB  
Article
Effects of Progressive Drought Stress on the Growth, Ornamental Values, and Physiological Properties of Begonia semperflorens
by Zhimin Zhao, Airong Liu, Yuanbing Zhang, Xiaodong Yang, Shuyue Yang and Kunkun Zhao
Horticulturae 2024, 10(4), 405; https://doi.org/10.3390/horticulturae10040405 - 16 Apr 2024
Viewed by 914
Abstract
Water is one of the most important elements affecting the growth of ornamental plants. To investigate the effects of drought stress on the growth, ornamental values, and physiological properties of Begonia semperflorens, watering treatments with 250 mL (control check, CK), 200 mL [...] Read more.
Water is one of the most important elements affecting the growth of ornamental plants. To investigate the effects of drought stress on the growth, ornamental values, and physiological properties of Begonia semperflorens, watering treatments with 250 mL (control check, CK), 200 mL (extremely light drought, ELD), 150 mL (light drought, LD), 100 mL (moderate drought, MD), 50 mL (severe drought, SD), and 25 mL (extremely severe drought, ESD) on the B. semperflorens variety “Chao Ao” were performed in this study. As a result, compared to the control (CK), the number of flowers, leaves, and branches, leaf size, plant height, crown diameter, as well as water content, transpiration rate, net photosynthetic rate, stomatal conductance, intercellular CO2 concentration, and chlorophyll content in leaves decreased, followed by an increased amount of drought stress. The contents of the osmotic adjustment substances, such as soluble sugar, soluble protein, proline, and betaine, were increased under drought stress. Indicators related to antioxidant activities, such as SOD activity, increased and then decreased. The POD activity, CAT activity, MDA content, and plasma membrane permeability of B. semperflorens were higher under increased drought stress than in the control condition. The APX activity decreased and then increased under drought stress. In conclusion, B. semperflorens responds to drought stress by increasing osmotic adjustment substances and antioxidant activities and reducing the water loss, growth potential, and photosynthetic rate. The correlation analysis showed that, except for APX, the drought resistance coefficients of 23 other indexes were correlated in different degrees. Therefore, this study suggests that B. semperflorens has a strong drought resistance ability, retaining high ornamental values in conditions of moderate drought stress, and can still survive under extremely high drought stress. Full article
Show Figures

Figure 1

Figure 1
<p>Phenotypes of <span class="html-italic">B. semperflorens</span> under drought stress; control check, extremely light drought, light drought, moderate drought, severe drought, and extremely severe drought conditions represent CK, ELD, LD, MD, SD, and ESD, respectively. Bar = 2 cm.</p>
Full article ">Figure 2
<p>Relative water content in the leaves and stems (<b>A</b>) and transpiration rates in the leaves (<b>B</b>) of <span class="html-italic">B. semperflorens</span> under drought stress. The values are the mean ± SE (<span class="html-italic">n</span> = 3). Significant differences were analyzed using Duncan’s multiple range test; the bars with different letters are significantly different from each other (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>The chlorophyll content (<b>A</b>), photosynthetic rate (<b>B</b>), intercellular CO<sub>2</sub> concentration (<b>C</b>), and stomatal conductance (<b>D</b>) in the <span class="html-italic">B. semperflorens</span> leaves under drought stress. Values are the mean ± SE (<span class="html-italic">n</span> = 3). Significant differences were analyzed using Duncan’s multiple range test; the bars with different letters are significantly different from each other (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>The osmotic adjustment substance protein content (<b>A</b>), betaine content (<b>B</b>), proline content (<b>C</b>), and soluble sugar (<b>D</b>) in <span class="html-italic">B. semperflorens</span> leaves under drought stress. Values are the mean ± SE (<span class="html-italic">n</span> = 3). Significant differences were analyzed using Duncan’s multiple range test; the bars with different letters are significantly different from each other (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of drought stress on the membrane permeability (<b>A</b>), MDA content (<b>B</b>), CAT activity (<b>C</b>), APX (<b>D</b>), SOD activity (<b>E</b>), and POD activity (<b>F</b>) in <span class="html-italic">B. semperflorens</span> leaves. Values are the mean ± SE (<span class="html-italic">n</span> = 3). Significant differences were analyzed using Duncan’s multiple range test; the bars with different letters are significantly different from each other (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Correlation analysis between 24 indexes of <span class="html-italic">B. semperflorens.</span> *, **, and *** indicate significant correlation at levels of <span class="html-italic">p</span> = 0.05, <span class="html-italic">p</span> = 0.01, and <span class="html-italic">p</span> = 0.001, respectively. Control check, extremely light drought, light drought, moderate drought, moderate drought, and extremely severe drought represent CK, ELD, LD, MD, SD, and ESD, respectively.</p>
Full article ">Figure 7
<p>Heat map analysis of 24 indexes of <span class="html-italic">B. semperflorens.</span> Control check, extremely light drought, light drought, moderate drought, moderate drought, and extremely severe drought represent CK, ELD, LD, MD, SD, and ESD, respectively.</p>
Full article ">
18 pages, 5923 KiB  
Article
Response of Dahlia Photosynthesis and Transpiration to High-Temperature Stress
by Jing-Jing Liu, Ying-Chan Zhang, Shan-Ce Niu, Li-Hong Hao, Wen-Bin Yu, Duan-Fen Chen and Di-Ying Xiang
Horticulturae 2023, 9(9), 1047; https://doi.org/10.3390/horticulturae9091047 - 18 Sep 2023
Cited by 3 | Viewed by 1699
Abstract
The high temperature may cause difficult growth or bloom in the summer, which is the key problem limiting the cultivation and application of dahlia. The photosynthetic physiological mechanisms of dahlia under high temperature stress were studied to provide a theoretical basis for expanding [...] Read more.
The high temperature may cause difficult growth or bloom in the summer, which is the key problem limiting the cultivation and application of dahlia. The photosynthetic physiological mechanisms of dahlia under high temperature stress were studied to provide a theoretical basis for expanding the application range of cultivation and annual production. Two dahlia varieties, ‘Tampico’ and ‘Hypnotica Tropical Breeze’, were used as test materials and were treated for 1 d or 2 d at temperatures of 35/30 °C or 40/35 °C (day/night: 14 h/10 h) and then recovered at 25/20 °C for 7 d. A 25/20 °C treatment was used as the control. The results are as follows: (1) High-temperature stress resulted in the chlorophyll (Chl) content, Fv/Fm, transpiration rate (Tr), net photosynthetic rate (Pn), and water potential decreasing significantly, and the Chl content, Tr, and stomatal density of ‘Tampico’ were higher than those of ‘Hypnotica Tropical Breeze’ during the same period. (2) After the two dahlia varieties were treated with high-temperature stress and recovered at 25/20 °C for 7 d, the plant morphology and various physiological indices under the 35/30 °C treatment gradually returned to normal, with ‘Tampico’ in better condition than ‘Hypnotica Tropical Breeze’. (3) Both dahlia varieties could not withstand the stress of 40/35 °C for 2 days. Full article
Show Figures

Figure 1

Figure 1
<p>Morphological change in ‘Tampico’ after high-temperature stress (<b>a</b>) 25/20 °C 1 d; (<b>b</b>) 25/20 °C 2 d; (<b>c</b>) 25/20 °C 7−d recovery; (<b>d</b>) 35/30 °C 1 d; (<b>e</b>) 35/30 °C 2 d; (<b>f</b>) 35/30 °C 7−d recovery; (<b>g</b>) 40/35 °C 1 d; (<b>h</b>) 40/35 °C 2 d; (<b>i</b>) 40/35 °C 7−d recovery.</p>
Full article ">Figure 2
<p>Morphological changes in ‘Hypnotica Tropical Breeze’ after high-temperature stress (<b>a</b>) 25/20 °C 1 d; (<b>b</b>) 25/20 °C 2 d; (<b>c</b>) 25/20 °C 7−d recovery; (<b>d</b>) 35/30 °C 1 d; (<b>e</b>) 35/30 °C 2 d; (<b>f</b>) 35/30 °C 7−d recovery; (<b>g</b>) 40/35 °C 1 d; (<b>h</b>) 40/35 °C 2 d; (<b>i</b>) 40/35 °C 7−d recovery.</p>
Full article ">Figure 3
<p>Changes in stomatal density per 1 mm<sup>−2</sup> leaf area in ‘Tampico’ (<b>a</b>) and ‘Hypnotica Tropical Breeze’ (<b>b</b>) after high temperature stress. Lowercase letters indicate the significance of different treatment times at the same temperature (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05), and capital letters indicate significant differences in different treatment temperatures for the same treatment days (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Change of stomatal aperture in ‘Tampico’ (<b>a</b>) and ‘Hypnotica Tropical Breeze’ (<b>b</b>) after high-temperature stress. Lowercase letters indicate the significance of different treatment times at the same temperature (Waller-Duncan test, <span class="html-italic">p</span> &lt; 0.05), and capital letters indicate significant differences in different treatment temperatures for the same treatment days (Waller–Duncan test, <span class="html-italic">p</span> &lt; 0.05). * is significantly different at the 0.05 probability levels (<span class="html-italic">t</span>-test).</p>
Full article ">Figure 5
<p>Change of the leaf water potential in ‘Tampico’ (<b>a</b>) and ‘Hypnotica Tropical Breeze’ (<b>b</b>) after high-temperature stress. Lowercase letters indicate the significance of different treatment times at the same temperature (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05), and capital letters indicate significant differences in different treatment temperatures for the same treatment days (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05). ** is significantly different at the 0.01 probability levels (<span class="html-italic">t</span>-test).</p>
Full article ">Figure 6
<p>Change of Chl a, Chl b and Car content in ‘Tampico’ (<b>a</b>,<b>c</b>,<b>e</b>) and ‘Hypnotica Tropical Breeze’ (<b>b</b>,<b>d</b>,<b>f</b>) after high-temperature stress. Lowercase letters indicate the significance of different treatment times at the same temperature (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05), and capital letters indicate significant differences in different treatment temperatures for the same treatment days (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Change of <span class="html-italic">F</span><sub>0</sub> in ‘Tampico’ (<b>a</b>) and ‘Hypnotica Tropical Breeze’ (<b>b</b>) after high-temperature stress. Lowercase letters indicate the significance of different treatment times at the same temperature (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05), and capital letters indicate significant differences in different treatment temperatures for the same treatment days (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Change of <span class="html-italic">Fm</span> in ‘Tampico’ (<b>a</b>) and ‘Hypnotica Tropical Breeze’ (<b>b</b>) after high-temperature stress. Lowercase letters indicate the significance of different treatment times at the same temperature (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05), and capital letters indicate significant differences in different treatment temperatures for the same treatment days (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05). ** is significantly different at the 0.01 probability levels (<span class="html-italic">t</span>-test).</p>
Full article ">Figure 9
<p>Change of <span class="html-italic">Fv</span>/<span class="html-italic">Fm</span> in ‘Tampico’ (<b>a</b>) and ‘Hypnotica Tropical Breeze’ (<b>b</b>) after high-temperature stress. Lowercase letters indicate the significance of different treatment times at the same temperature (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05), and capital letters indicate significant differences in different treatment temperatures for the same treatment days (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05). * and ** are significantly different at the 0.05 and 0.01 probability levels (<span class="html-italic">t</span>-test).</p>
Full article ">Figure 10
<p>Change of photosynthetic parameters <span class="html-italic">Pn</span> and <span class="html-italic">Tr</span> of ‘Tampico’ (<b>a</b>,<b>c</b>) and ‘Hypnotica Tropical Breeze’ (<b>b</b>,<b>d</b>) after high-temperature stress. Lowercase letters indicate the significance of different treatment times at the same temperature (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05), and capital letters indicate significant differences in different treatment temperatures for the same treatment days (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05). ** is significantly different at the 0.01 probability levels (<span class="html-italic">t</span>–test).</p>
Full article ">Figure 10 Cont.
<p>Change of photosynthetic parameters <span class="html-italic">Pn</span> and <span class="html-italic">Tr</span> of ‘Tampico’ (<b>a</b>,<b>c</b>) and ‘Hypnotica Tropical Breeze’ (<b>b</b>,<b>d</b>) after high-temperature stress. Lowercase letters indicate the significance of different treatment times at the same temperature (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05), and capital letters indicate significant differences in different treatment temperatures for the same treatment days (Waller−Duncan test, <span class="html-italic">p</span> &lt; 0.05). ** is significantly different at the 0.01 probability levels (<span class="html-italic">t</span>–test).</p>
Full article ">
14 pages, 2228 KiB  
Article
A B-Box Transcription Factor CoBBX24 from Camellia oleifera Delays Leaf Senescence and Enhances Drought Tolerance in Arabidopsis
by Yanan Liu, Zhiguo Zhu, Yang Wu, Yinxiang Gao, Lisha Zhang, Changshuai Yu, Sicheng Ye and Wenxin Liu
Horticulturae 2023, 9(9), 991; https://doi.org/10.3390/horticulturae9090991 - 1 Sep 2023
Cited by 1 | Viewed by 1137
Abstract
Plants face various biotic and abiotic stress factors during their growth and development, among which, drought is a serious adverse factor that affects yield and quality in agriculture and forestry. Several transcription factors are involved in regulating plant responses to drought stress. In [...] Read more.
Plants face various biotic and abiotic stress factors during their growth and development, among which, drought is a serious adverse factor that affects yield and quality in agriculture and forestry. Several transcription factors are involved in regulating plant responses to drought stress. In this study, the B-box (BBX) transcription factor CoBBX24 was cloned from Camellia oleifera. This gene encodes a 241-amino-acid polypeptide containing two B-box domains at the N-terminus. A phylogenetic analysis revealed that CoBBX24 and CsBBX24 from Camellia sinensis are in the same branch, with their amino acid sequences being identical by 96.96%. CoBBX24 was localized to the nucleus and acted as a transcriptional activator. The overexpression of CoBBX24 in Arabidopsis heightened its drought tolerance along with a relatively high survival rate, and the rate of water loss in the OX-CoBBX24 lines was observably lower than that of the wild-type. Compared to the wild-type, the root lengths of the OX-CoBBX24 lines were significantly inhibited with abscisic acid. Leaf senescence was delayed in the OX-CoBBX24 lines treated with abscisic acid. The expression of genes related to leaf senescence and chlorophyll breakdown (e.g., SAG12, SAG29, NYC1, NYE1, and NYE2) was downregulated in the OX-CoBBX24 lines. This study indicated that CoBBX24 positively regulates the drought tolerance in Arabidopsis through delayed leaf senescence. Full article
Show Figures

Figure 1

Figure 1
<p>Phylogenetic characteristics and structural domains of CoBBX24. (<b>A</b>) Phylogenetic evaluation of CoBBX24 and other plant BBX proteins. The phylogenetic trees were derived using the neighbor-joining (NJ) method with a bootstrap value of 1000 replicates. Bootstrap values indicate the divergence of each branch, with the scale representing the branch length. Red triangle indicate the protein sequence from <span class="html-italic">Camellia oleifera</span>. (<b>B</b>) Alignment of the deduced polypeptide sequences of CoBBX22 with those of other plant BBXs. Red lines indicate the conserved B1 and B2 B-box domains. The accession numbers of the proteins are: CsBBX24 (XP_028122857.1), VvBBX25 (XP_028122857.1), RcBBX24 (XP_024193584.1), ZjBBX24 (XP_015900039.1), PeBBX24 (XP_015900039.1), PmBBX24 (XP_008222603.1), CaBBX24 (XP_027061493.1), JrBBX24 (XP_018839979.2), and MdBBX24 (XP_028956746.1).</p>
Full article ">Figure 2
<p>Transcriptional profiling of <span class="html-italic">CoBBX24</span> under drought stress, transactivation analysis, and subcellular localization of CoBBX24. (<b>A</b>) Relative expression of <span class="html-italic">CoBBX24</span> under drought. <span class="html-italic">Camellia oleifera Tub3α</span> was used as the reference gene for normalization. Error bars indicate the standard deviation (SD); n = 3. Significant differences were determined by Duncan’s test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). (<b>B</b>) Transactivation activity analysis of the CoBBX24 protein in yeast cells. pCL1 served as a positive control, and pGBKT7 served as a negative control. (<b>C</b>) Subcellular localization of CoBBX24 in tobacco (<span class="html-italic">N. benthamiana</span>) cells. 35S::D53-RFP was used as a nuclear marker. Bars: 20 μm.</p>
Full article ">Figure 3
<p>Overexpression of <span class="html-italic">CoBBX24</span> enhanced the tolerance of transgenic <span class="html-italic">Arabidopsis</span> under drought stress. (<b>A</b>) Phenotypes of 3-week-old <span class="html-italic">CoBBX24</span> transgenic and WT plants withheld water for 15 d followed by recovery for 7 d with regular watering. Three independent assays were performed with similar findings. (<b>B</b>) Survival rates of <span class="html-italic">CoBBX24</span> transgenic and WT plants after 7 d of re-watering following a 15 d drought treatment. Three independent experiments were performed, a total of 118 plants were counted for each genotype. Significant differences were determined by Duncan’s test (<span class="html-italic">p</span> &lt; 0.01). (<b>C</b>) Water loss from detached leaves of <span class="html-italic">CoBBX24</span> transgenic lines and WT plants. The data are presented as means ± SD of three replicates. * represents a significant difference compared with WT; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 in the Student’s test.</p>
Full article ">Figure 4
<p>The constitutive expression of <span class="html-italic">CoBBX24</span> in <span class="html-italic">A. thaliana</span> enhanced root length sensitivity to ABA treatment. (<b>A</b>) Seedlings grown on 1/2 MS plates for 4 d followed by transfer to media containing 3 μM ABA; Bars: 1 cm. (<b>B</b>) The root length measured on the 6th d post-transfer. Significant differences were determined by Duncan’s test (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>The phenotypic effect of ABA treatment on <span class="html-italic">CoBBX24ox</span> and WT <span class="html-italic">Arabidopsis</span> plants. (<b>A</b>) Detached leaves of 4-week-old transgenic plants after maintaining them for 2 d under dark conditions. (<b>B</b>) The chlorophyll content of the leaves shown in (<b>A</b>); significant differences were determined by Duncan’s test (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>qRT-PCR assay revealed the expression of (<b>A</b>) <span class="html-italic">AtABF4</span>, (<b>B</b>) <span class="html-italic">AtSAG29</span>, (<b>C</b>) <span class="html-italic">AtSAG12</span>, (<b>D</b>) <span class="html-italic">AtNYC1</span>, (<b>E</b>) <span class="html-italic">AtNYE1</span>, and (<b>F</b>) <span class="html-italic">AtNYE2</span> in <span class="html-italic">CoBBX24ox</span> and WT plants under ABA treatment. The <span class="html-italic">Arabidopsis Actin2</span> gene was used as the reference gene for normalization. Error bars indicate the SD; n = 3. Significant differences were determined by Duncan’s test (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
14 pages, 8960 KiB  
Article
Genome-Wide Identification of Fatty Acyl-CoA Reductase (FAR) Genes in Dendrobium catenatum and Their Response to Drought Stress
by Yutong Ren, Peng Wang, Tingting Zhang, Wen Liu, Yujuan Wang, Jun Dai and Yang Zhou
Horticulturae 2023, 9(9), 982; https://doi.org/10.3390/horticulturae9090982 - 31 Aug 2023
Cited by 2 | Viewed by 1201
Abstract
Dendrobium catenatum is a high-value medicinal plant that is predominantly found in high mountain areas, thriving amidst cliffs and rock crevices. However, its wild resources face constant threats from adverse environmental conditions, especially drought stress. Fatty acyl-CoA reductase (FAR) is crucial in plant [...] Read more.
Dendrobium catenatum is a high-value medicinal plant that is predominantly found in high mountain areas, thriving amidst cliffs and rock crevices. However, its wild resources face constant threats from adverse environmental conditions, especially drought stress. Fatty acyl-CoA reductase (FAR) is crucial in plant drought resistance, but there is a lack of research on FAR genes in D. catenatum. In this study, the FAR family genes were identified from the D. catenatum genome. Their genomic characteristics were investigated using bioinformatics techniques, and their expression patterns in different tissues and under 20% PEG8000 conditions mimicking drought stress were analyzed using quantitative real-time RT-PCR (RT-qPCR). Seven DcFAR genes were identified from the D. catenatum genome. The encoded amino acids range between 377 and 587 aa, with molecular weights between 43.41 and 66.15 kD and isoelectric points between 5.55 and 9.02. Based on the phylogenetic relationships, the FAR family genes were categorized into three subgroups, each with similar conserved sequences and gene structures. The cis-acting elements of the promoter regions were assessed, and the results reveal that the DcFAR upstream promoter region contains multiple stress-related elements, suggesting its potential involvement in abiotic stress responses. The RT-qPCR results show distinct expression patterns of DcFAR genes in various plant tissues. It was observed that the expression of most DcFAR genes was upregulated under drought stress. Among them, the expression levels of DcFAR2, DcFAR3, DcFAR5, and DcFAR7 genes under drought stress were 544-, 193-, 183-, and 214-fold higher compared to the control, respectively. These results indicate that DcFAR2/3/5/7 might play significant roles in D. catenatum drought tolerance. This research offers insight into the function of DcFAR genes and provides theoretical support for breeding drought-resistant D. catenatum varieties. Full article
Show Figures

Figure 1

Figure 1
<p>Phylogenetic analysis and protein sequence comparison of FAR proteins. (<b>A</b>) Phylogenetic analysis of FAR proteins from <span class="html-italic">D. catenatum</span>, <span class="html-italic">Arabidopsis</span>, rice, and sorghum. The phylogenetic tree was constructed using the Neighbor-Joining (NJ) method in MEGA-X, with default parameters. Red star represents <span class="html-italic">D. catenatum</span> FAR (DcFAR); black triangle represents <span class="html-italic">Arabidopsis</span> FAR (AtFAR); green square represents sorghum FAR; and blue circle represents rice FAR (OsFAR). (<b>B</b>) Protein sequence comparison of FAR proteins using DNAMAN. The three red boxes represent the GXXGXX(G/A) and YXXXK conserved domains.</p>
Full article ">Figure 2
<p>Conserved motif and gene structure analyses of DcFAR family members. (<b>A</b>) Conserved motifs of DcFAR proteins. Rectangular boxes of different colors represent different conserved motifs. (<b>B</b>) Exon/intron structure of <span class="html-italic">DcFAR</span>. UTR(s), exon(s), and intron(s) are represented by green boxes, yellow boxes, and black lines, respectively.</p>
Full article ">Figure 3
<p><span class="html-italic">Cis</span>-acting element analysis in the promoter of <span class="html-italic">DcFAR</span> genes. (<b>A</b>) Locations of <span class="html-italic">cis</span>-acting elements in the promoter of <span class="html-italic">DcFAR</span>. Ellipses of different colors represent different types of <span class="html-italic">cis</span>-acting elements and their positions in each <span class="html-italic">DcFAR</span> gene promoter. (<b>B</b>) Statistics of the number of <span class="html-italic">cis</span>-acting elements in <span class="html-italic">DcFAR</span> promoters. Different colors and numbers represent the number of different <span class="html-italic">cis</span>-acting elements in each <span class="html-italic">DcFAR</span> promoter.</p>
Full article ">Figure 4
<p>Expression analysis of <span class="html-italic">DcFAR</span> genes under drought stress using RT-qPCR. The data are expressed as mean ± standard deviation (<span class="html-italic">n</span> = 3). Vertical bars represent the means of fold change in expression and standard deviations calculated from the replicates. Values of 0, 3, 6, 9, 12, 24, and 48 indicate hours after treatment. Asterisks (* or **) indicate a significant difference at <span class="html-italic">p</span> &lt; 0.05 or 0.01, respectively.</p>
Full article ">Figure 5
<p>Expression analysis of <span class="html-italic">DcFAR</span> genes in different tissues using RT-qPCR. Mean expression values were calculated from three independent biological replicates relative to the value in roots and visualized using TBtools. Green and red indicate low and high levels of expression, respectively. RO: root; ST: stem; LE: leaf; CA: capsule; E: petal; SE: sepal; FS: flower stalk; GY: gynostemia; LI: lip.</p>
Full article ">
29 pages, 10634 KiB  
Article
Transcriptomic and Metabolomic Analyses of Seedlings of Two Grape Cultivars with Distinct Tolerance Responses to Flooding and Post-Flooding Stress Conditions
by Yanjie Peng, Jinli Chen, Wenjie Long, Pan He, Qi Zhou, Xia Hu, Yong Zhou and Ying Zheng
Horticulturae 2023, 9(9), 980; https://doi.org/10.3390/horticulturae9090980 - 30 Aug 2023
Cited by 1 | Viewed by 1292
Abstract
Grapes, an important and widespread fruit crop providing multiple products, face increasing flooding risks due to intense and frequent extreme rainfall. It is thus imperative to fully understand the flood-tolerance mechanisms of grapevines. Here, RNA-seq and LC-MS/MS technologies were used to analyze the [...] Read more.
Grapes, an important and widespread fruit crop providing multiple products, face increasing flooding risks due to intense and frequent extreme rainfall. It is thus imperative to fully understand the flood-tolerance mechanisms of grapevines. Here, RNA-seq and LC-MS/MS technologies were used to analyze the transcriptome and metabolome changes in the roots of SO4 (tolerant to flooding) and Kyoho (sensitive to flooding) grapes under flooding and post-flooding conditions. The results showed that the abundance of many metabolites in the phenylpropanoids and polyketides, organic acids and their derivatives, and organic oxygen compounds superclasses changed in different patterns between the Kyoho and SO4 grapes under flooding and post-flooding conditions. Jasmonic acid and the ascorbic acid–glutathione cycle played a pivotal role in coping with both hypoxia stress and reoxygenation stress incurred during flooding and post-flooding treatments in the SO4 cultivar. Under flooding stress, the regulatory mechanistic shift from aerobic respiration to anaerobic fermentation under hypoxia is partly missing in the Kyoho cultivar. In the post-flooding stage, many genes related to ethylene, gibberellins, cytokinins, and brassinosteroids biosynthesis and brassinosteroids-responsive genes were significantly downregulated in the Kyoho cultivar, adversely affecting growth recovery; however, their expression was not reduced in the SO4 cultivar. These findings enhance our understanding of the flooding-tolerance mechanisms in grapes. Full article
Show Figures

Figure 1

Figure 1
<p>Principal component analysis (PCA) and statistical analysis of differentially expressed genes (DEGs). (<b>A</b>) PCA of the 18 RNA-seq libraries. (<b>B</b>) Numbers of DEGs. (<b>C</b>) Venn diagram of DEGs.</p>
Full article ">Figure 2
<p>Significantly enriched KEGG pathways under flooding and post-flooding conditions in the Kyoho and SO4 grapevine cultivars.</p>
Full article ">Figure 3
<p>Analysis of differentially expressed genes (DEGs) encoding transcription factors (TFs). (<b>A</b>) Statistical analysis of differentially expressed TF gene numbers at the gene family level. (<b>B</b>) The log2 fold-change values of candidate TF genes related to flood tolerance. All values are from comparisons of treatment vs. control groups. NS, not significantly changed in comparison with control.</p>
Full article ">Figure 3 Cont.
<p>Analysis of differentially expressed genes (DEGs) encoding transcription factors (TFs). (<b>A</b>) Statistical analysis of differentially expressed TF gene numbers at the gene family level. (<b>B</b>) The log2 fold-change values of candidate TF genes related to flood tolerance. All values are from comparisons of treatment vs. control groups. NS, not significantly changed in comparison with control.</p>
Full article ">Figure 4
<p>PLS-DA, quantitative analysis, and classification of DEMs (differentially expressed metabolites). (<b>A</b>) PLS-DA map of the 18 metabolomic libraries. (<b>B</b>) Numbers of upregulated and downregulated DEMs during flooding and post-flooding conditions. (<b>C</b>) Classification of DEMs according to the HMDB database.</p>
Full article ">Figure 4 Cont.
<p>PLS-DA, quantitative analysis, and classification of DEMs (differentially expressed metabolites). (<b>A</b>) PLS-DA map of the 18 metabolomic libraries. (<b>B</b>) Numbers of upregulated and downregulated DEMs during flooding and post-flooding conditions. (<b>C</b>) Classification of DEMs according to the HMDB database.</p>
Full article ">Figure 5
<p>Fold-changes for the log10 values of DEMs in the (<b>A</b>) phenylpropanoids and polyketides superclass, (<b>B</b>) organic acids and derivatives superclass, and (<b>C</b>) organic oxygen compounds superclass. Values for metabolites are the fold-change of their log10 values. All values are from comparisons of treatment vs. control groups. NS, not significantly changed in comparison with control.</p>
Full article ">Figure 5 Cont.
<p>Fold-changes for the log10 values of DEMs in the (<b>A</b>) phenylpropanoids and polyketides superclass, (<b>B</b>) organic acids and derivatives superclass, and (<b>C</b>) organic oxygen compounds superclass. Values for metabolites are the fold-change of their log10 values. All values are from comparisons of treatment vs. control groups. NS, not significantly changed in comparison with control.</p>
Full article ">Figure 6
<p>Integrative analysis of transcriptome and metabolome data for the two grapevine cultivars. (<b>A</b>) Procrustes analysis of the six groups. (<b>B</b>) KEGG enrichment analysis and corresponding <span class="html-italic">p</span>-value.</p>
Full article ">Figure 6 Cont.
<p>Integrative analysis of transcriptome and metabolome data for the two grapevine cultivars. (<b>A</b>) Procrustes analysis of the six groups. (<b>B</b>) KEGG enrichment analysis and corresponding <span class="html-italic">p</span>-value.</p>
Full article ">Figure 7
<p>Metabolite changes and expression change of genotype-specific DEGs (differentially expressed genes) related to JA response (<b>A</b>) and ETH response (<b>B</b>) under flooding (at day 7) in the Kyoho and SO4 grapevine cultivars. Values for metabolites are the fold-change of their log10 values; those for DEGs are their log2 fold-change values. All values are from the comparisons of the treatment vs. control groups. NS, not significantly changed in comparison with control. The left blocks belong to KF7, while the right blocks belong to SF7.</p>
Full article ">Figure 7 Cont.
<p>Metabolite changes and expression change of genotype-specific DEGs (differentially expressed genes) related to JA response (<b>A</b>) and ETH response (<b>B</b>) under flooding (at day 7) in the Kyoho and SO4 grapevine cultivars. Values for metabolites are the fold-change of their log10 values; those for DEGs are their log2 fold-change values. All values are from the comparisons of the treatment vs. control groups. NS, not significantly changed in comparison with control. The left blocks belong to KF7, while the right blocks belong to SF7.</p>
Full article ">Figure 8
<p>Metabolite changes and expression change of genotype-specific DEGs (differentially expressed genes) related to carbohydrates metabolism under flooding (day 7) in the Kyoho and SO4 grapevine cultivars. Values of metabolites are the fold-change of their log10 values; those for DEGs are their log2 fold-change values. All values are from the comparisons of the treatment vs. control groups. NS, not significantly changed. The left blocks belong to KF7, while the right blocks belong to SF7.</p>
Full article ">Figure 9
<p>Metabolite changes and expression change of genotype-specific DEGs (differentially expressed genes) related to phytohormone regulation under post-flooding (day 10) in the Kyoho and SO4 grapevine cultivars. (<b>A</b>) JA; (<b>B</b>) BR and CK; (<b>C</b>) ETH, GA, and ABA. Values for metabolites are the fold-change of their log10 values; those for DEGs are their log2 fold-change values. All values are from the comparisons of the treatment vs. control groups. NS, not significantly changed in comparison with control. The left blocks belong to KF7R, while the right blocks belong to SF7R.</p>
Full article ">Figure 9 Cont.
<p>Metabolite changes and expression change of genotype-specific DEGs (differentially expressed genes) related to phytohormone regulation under post-flooding (day 10) in the Kyoho and SO4 grapevine cultivars. (<b>A</b>) JA; (<b>B</b>) BR and CK; (<b>C</b>) ETH, GA, and ABA. Values for metabolites are the fold-change of their log10 values; those for DEGs are their log2 fold-change values. All values are from the comparisons of the treatment vs. control groups. NS, not significantly changed in comparison with control. The left blocks belong to KF7R, while the right blocks belong to SF7R.</p>
Full article ">Figure 10
<p>Metabolite changes and expression change of genotype-specific DEGs (differentially expressed genes) related to (<b>A</b>) flavonoid and monolignol biosynthesis and (<b>B</b>) endoplasmic reticulum-associated degradation (ERAD) under post-flooding (day 10) in the Kyoho and SO4 grapevine cultivars. Values for metabolites are the fold-change values of their log10 values; those for DEGs are their log2 fold-change values. All values are from the comparison of the treatment vs. control groups. NS, not significantly changed in comparison with control. The left blocks belong to KF7R, while the right blocks belong to SF7R.</p>
Full article ">Figure 11
<p>A brief schematic introducing the research steps used in this study.</p>
Full article ">Figure 12
<p>A briefly summary of the mechanism of the flooding response and post-flooding response in flood-tolerant SO4.</p>
Full article ">
Back to TopTop