Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (551)

Search Parameters:
Keywords = leaf senescence

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 6469 KiB  
Article
Genome-Wide Identification and Analysis of the WRKY Transcription Factor Family Associated with Leaf Senescence in Alfalfa
by Xiaojing Peng, Jinning Hu, Xiangxue Duan, Maofeng Chai, Jiangqi Wen, Zengyu Wang and Hongli Xie
Plants 2024, 13(19), 2725; https://doi.org/10.3390/plants13192725 (registering DOI) - 29 Sep 2024
Abstract
Leaves are the most significant parts of forage crops such as alfalfa. Senescence is the terminal stage of leaf development and is controlled by an integrated myriad of endogenous signals and environmental stimuli. WRKY transcription factors (TFs) play essential roles in regulating leaf [...] Read more.
Leaves are the most significant parts of forage crops such as alfalfa. Senescence is the terminal stage of leaf development and is controlled by an integrated myriad of endogenous signals and environmental stimuli. WRKY transcription factors (TFs) play essential roles in regulating leaf senescence; however, only a few studies on the analysis and identification of the WRKY TF family in Medicago Sativa have been reported. In this study, we identified 198 WRKY family members from the alfalfa (M. sativa L.) cultivar ’XinjiangDaye’ using phylogenetic analysis and categorized them into three subfamilies, Groups I, II, and III, based on their structural characteristics. Group II members were further divided into five subclasses. In addition, several hormone- and stress-related cis-acting elements were identified in the promoter regions of MsWRKYs. Furthermore, 14 aging-related MsWRKYs genes from a previous transcriptome in our laboratory were selected for RT-qPCR validation of their expression patterns, and subsequently cloned for overexpression examination. Finally, MsWRKY5, MsWRKY66, MsWRKY92, and MsWRKY141 were confirmed to cause leaf yellowing in Nicotiana benthaminana using a transient expression system. Our findings lay a groundwork for further studies on the mechanism of M. sativa leaf aging and for the creation of new germplasm resources. Full article
(This article belongs to the Special Issue Abiotic and Biotic Stress of the Crops and Horticultural Plants)
Show Figures

Figure 1

Figure 1
<p>A neighbor-joining phylogenetic tree was constructed using MEGA11.0 software with 1000 boot-strap replications, by comparing WRKY TFs from <span class="html-italic">M. sativa</span> L. (Ms) and <span class="html-italic">Arabidopsis thaliana</span> (At). Various highlighted colors correspond to the different subgroups.</p>
Full article ">Figure 2
<p>Analysis of phylogenetic relationships, motifs, and gene structure of WRKY TFs from <span class="html-italic">M. sativa</span>. (<b>a</b>) Phylogenetic tree of 198 MsWRKYs in <span class="html-italic">M. sativa</span>. The colors highlighted the different subgroups are same as that in <a href="#plants-13-02725-f001" class="html-fig">Figure 1</a>. (<b>b</b>) Conserved motif arrangements of MsWRKYs. The motifs are highlighted in various colored boxes. Motif 1 represents the WRKY domain. (<b>c</b>) Exon-intron organizations of <span class="html-italic">MsWRKYs</span>. Green boxes indicate exons; black lines indicate introns.</p>
Full article ">Figure 3
<p>The 2 kb promoter sequences of the <span class="html-italic">MsWRKY</span> gene contain various <span class="html-italic">cis</span>-acting elements. Different colored rectangles indicate different <span class="html-italic">cis</span>-elements, positioned according to their locations within the promoters.</p>
Full article ">Figure 4
<p>A heatmap displaying the RNA−Seq data for 198 <span class="html-italic">MsWRKYs</span>, with expression levels normalized by row using the Z−Scores algorithm. The color scale on the right of the heatmap shows relative expression, with the color gradient from blue to red indicating increased expression levels. X0 (top not fully unfolded leaf), X1 (top fully unfolded first leaf), X2 (top fully unfolded second leaf), X3 (top fully unfolded third leaf), X4 (bottom leaf with senescent symptom); D0, D1, D2, D4, D6 (leaves treated in the dark for 0, 1, 2, 4 and 6 days); S1, S2, S4, S6 (leaves treated with salt for 1, 2, 4 and 6 days). Fourteen genes selected for RT-qPCR were labeled with an asterisk.</p>
Full article ">Figure 5
<p>RT-qPCR results of 14 <span class="html-italic">MsWRKYs</span> in the process of leaf senescence under (<b>a</b>) natural condition (X0, X1, X2, X3, X4 represent different stages of leaf development), (<b>b</b>) dark stress (D0, D1, D2, D4, D6 represent 0, 1, 2, 4, and 6 days of dark treatment), (<b>c</b>) salt stress (S1, S2, S4, S6 represent 1, 2, 4, and 6 days of the 150 mM NaCl treatment). The error bars indicate the standard deviation of three biological replicates. Relative expression was calculated using the 2<sup>–ΔCT</sup> method. The data for gene expression are presented as the mean ± SD and were analyzed to detect significant differences by ANOVA using GraphPad Prism 8 (NS: not significant; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01) against D0 or X0.</p>
Full article ">Figure 6
<p>Functional validation of selected <span class="html-italic">MsWRKYs</span> was conducted using an <span class="html-italic">Agrobacterium</span>-mediated transient expression assay. Symptoms of leaf senescence in representative <span class="html-italic">Nicotiana benthamiana</span> leaves appeared after infiltration with various constructs encoding <span class="html-italic">MsWRKY5</span>, <span class="html-italic">MsWRKY66</span>, <span class="html-italic">MsWRKY92</span>, <span class="html-italic">MsWRKY141</span>, <span class="html-italic">MsSGR</span> and an empty vector with YFP. Positive control: SGR. Negative control: empty vector with YFP. Bar = 1 cm.</p>
Full article ">
15 pages, 16488 KiB  
Article
YELLOW LEAF AND DWARF 7, Encoding a Novel Ankyrin Domain-Containing Protein, Affects Chloroplast Development in Rice
by Yongtao Cui, Jian Song, Liqun Tang and Jianjun Wang
Genes 2024, 15(10), 1267; https://doi.org/10.3390/genes15101267 - 27 Sep 2024
Abstract
Background: The proper development of grana and stroma within chloroplasts is critical for plant vitality and crop yield in rice and other cereals. While the molecular mechanisms underpinning these processes are known, the genetic networks governing them require further exploration. Methods and Results: [...] Read more.
Background: The proper development of grana and stroma within chloroplasts is critical for plant vitality and crop yield in rice and other cereals. While the molecular mechanisms underpinning these processes are known, the genetic networks governing them require further exploration. Methods and Results: In this study, we characterize a novel rice mutant termed yellow leaf and dwarf 7 (yld7), which presents with yellow, lesion-like leaves and a dwarf growth habit. The yld7 mutant shows reduced photosynthetic activity, lower chlorophyll content, and abnormal chloroplast structure. Transmission electron microscopy (TEM) analysis revealed defective grana stacking in yld7 chloroplasts. Additionally, yld7 plants accumulate high levels of hydrogen peroxide (H2O2) and exhibit an up-regulation of senescence-associated genes, leading to accelerated cell death. Map-based cloning identified a C-to-T mutation in the LOC_Os07g33660 gene, encoding the YLD7 protein, which is a novel ankyrin domain-containing protein localized to the chloroplast. Immunoblot analysis of four LHCI proteins indicated that the YLD7 protein plays an important role in the normal biogenesis of chloroplast stroma and grana, directly affecting leaf senescence and overall plant stature. Conclusions: This study emphasizes the significance of YLD7 in the intricate molecular mechanisms that regulate the structural integrity of chloroplasts and the senescence of leaves, thus providing valuable implications for the enhancement of rice breeding strategies and cultivation. Full article
(This article belongs to the Special Issue Genetics and Breeding of Rice)
Show Figures

Figure 1

Figure 1
<p>Phenotypic comparison of wild-type (WT) and <span class="html-italic">yld7</span> plants. (<b>a</b>) Morphological differences in WT and <span class="html-italic">yld7</span> at seedling stages. (<b>b</b>), Leaf phenotype of WT and <span class="html-italic">yld7</span> plants at seedling stages (the youngest fully expanded leaf). (<b>c</b>) Cross-sections of leaf of WT and <span class="html-italic">yld7</span> at seedling stage. (<b>d</b>) Stomatal density comparison with red arrows marking stomata. Bars = 100 μm. (<b>e</b>) Statistical data on stomatal length (<b>d</b>) and width (<b>e</b>), with three and twenty independent replicates, respectively (** <span class="html-italic">p</span> &lt; 0.01). (<b>f</b>) Morphological differences in WT and <span class="html-italic">yld7</span> at tillering stages. (<b>g</b>) Chlorophyll content analysis. Error bars represent SD, n = 10 (** <span class="html-italic">p</span> &lt; 0.01). (<b>h</b>) Comparison of photosynthetic rates. Error bars denote SD, n = 15 (** <span class="html-italic">p</span> &lt; 0.01). (<b>i</b>,<b>j</b>) Gene expression related to chlorophyll synthesis (<b>i</b>) and chloroplast development (<b>j</b>) at the tillering stage. (<b>k</b>) Statistical data on tillering number, plant height, grains per panicle, setting rate, primary branch, and secondary branch in both plant types. Values are means ± SD of three biological replicates (** <span class="html-italic">p</span> &lt; 0.01, Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 2
<p>ROS accumulation in WT and <span class="html-italic">yld7</span> plants. (<b>a</b>,<b>b</b>) DAB and NBT staining at the seedling stage. (<b>c</b>–<b>f</b>) Quantitative analysis of H<sub>2</sub>O<sub>2</sub>, ORF content, and CAT, POD enzyme activities. Mean ± SD from five replicates (** <span class="html-italic">p</span> &lt; 0.01). (<b>g</b>) Relative expression of ROS detoxification genes. Error bars represent SD, n = 3 (** <span class="html-italic">p</span> &lt; 0.01, Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 3
<p>Cell death detection in WT and <span class="html-italic">yld7</span> plants. (<b>a</b>) Cell death detection in <span class="html-italic">yld7</span> plants. Bars = 100 μm (<b>b</b>,<b>c</b>) Expression analysis of CDGs (<b>b</b>) and other SAGs (<b>c</b>). Error bars represent SD, n = 3 (** <span class="html-italic">p</span> &lt; 0.01, Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 4
<p><span class="html-italic">YLD7</span> candidate gene analysis. (<b>a</b>) <span class="html-italic">YLD7</span>’s chromosomal location with marker data. Numbers indicate recombinants. (<b>b</b>) Structure and sequence variations in LOC_Os07g33660 between WT and <span class="html-italic">yld7</span>. (<b>c</b>) Genome and protein-level mutations with red arrows marking the site. (<b>d</b>) <span class="html-italic">YLD7</span> transcript levels at the 3-leaf stage, normalized to rice ubiquitin. Error bars represent SD, n = 3 (** <span class="html-italic">p</span> &lt; 0.01, Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 5
<p><span class="html-italic">YLD7</span> targeted deletion via CRISPR/Cas9. (<b>a</b>) Confirmation of <span class="html-italic">YLD7</span> deletion in WT with Cas9/sgRNA constructs. Target sequence highlighted, PAM site in yellow. Altered sequences in red. (<b>b</b>) Phenotypes of mutants with <span class="html-italic">YLD7</span> Cas9/sgRNA constructs (bar = 1 cm). (<b>c</b>) <span class="html-italic">YLD7</span> transcript levels in WT and transgenic plants. Error bars represent SD, n = 3 (** <span class="html-italic">p</span> &lt; 0.01, Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 6
<p>Phylogenetic analysis of YLD7 homologues. (<b>a</b>) Amino acid sequence alignment with conserved residues shaded. Amino acids that were fully or partially conserved are dark blue and pink, respectively. (<b>b</b>) Phylogenetic tree constructed using MEGA 7.0, showing evolutionary relationships. The scale bar represents the percentage of substitutions per site.</p>
Full article ">Figure 7
<p>Subcellular localization of YLD7-GFP. i. Control GFP image. ii. YLD7-GFP fusion localized to the chloroplast (bar = 10 μm).</p>
Full article ">Figure 8
<p>Chloroplast ultrastructure in WT and <span class="html-italic">yld7</span>. (<b>a</b>) WT chloroplasts. Bars = 1 μm. (<b>b</b>) Close-up of WT chloroplast, highlighting stromal thylakoids (ST). Bars = 0.5 μm. (<b>c</b>,<b>d</b>) Abnormal chloroplasts in <span class="html-italic">yld7</span>. Bars = 1 μm and 0.5 μm, respectively. GT, Grana thylakoid stacks; ST, stromal thylakoids. (<b>e</b>) Immunodetection of thylakoid proteins with specific antibodies.</p>
Full article ">Figure 9
<p>RNA-seq analysis in WT and <span class="html-italic">yld7</span>. (<b>a</b>) Transcript expression comparison. (<b>b</b>) Volcano plot of gene expression changes. Red represents highly expressed genes. Blue represents low expressed genes. (<b>c</b>) Cluster analysis of differentially expressed genes. Red represents highly expressed genes. Green represents low expressed genes.</p>
Full article ">Figure 10
<p><span class="html-italic">YLD7</span> genetic diversity analysis. (<b>a</b>,<b>b</b>) SNP variation types in <span class="html-italic">YLD7</span> CDs across 3127 accessions. Top panel shows the SNP variants (<b>a</b>), and the bottom panel displays the distribution of these variants across different rice subspecies (<b>b</b>).</p>
Full article ">Figure 11
<p>Simplified schematic model in YLD7. Arrow: activate; Bar: repress. Down red arrow: down-regulation; Up red arrow: up-regulation.</p>
Full article ">
19 pages, 5399 KiB  
Article
Identification of a Novel Gene MtbZIP60 as a Negative Regulator of Leaf Senescence through Transcriptome Analysis in Medicago truncatula
by Jiayu Xing, Jialan Wang, Jianuo Cao, Ke Li, Xiao Meng, Jiangqi Wen, Kirankumar S. Mysore, Geng Wang, Chunjiang Zhou and Pengcheng Yin
Int. J. Mol. Sci. 2024, 25(19), 10410; https://doi.org/10.3390/ijms251910410 - 27 Sep 2024
Abstract
Leaves are the primary harvest portion in forage crops such as alfalfa (Medicago sativa). Delaying leaf senescence is an effective strategy to improve forage biomass production and quality. In this study, we employed transcriptome sequencing to analyze the transcriptional changes and [...] Read more.
Leaves are the primary harvest portion in forage crops such as alfalfa (Medicago sativa). Delaying leaf senescence is an effective strategy to improve forage biomass production and quality. In this study, we employed transcriptome sequencing to analyze the transcriptional changes and identify key senescence-associated genes under age-dependent leaf senescence in Medicago truncatula, a legume forage model plant. Through comparing the obtained expression data at different time points, we obtained 1057 differentially expressed genes, with 108 consistently up-regulated genes across leaf growth and senescence. Gene Ontology and Kyoto Encyclopedia of Genes and Genomes pathway enrichment analyses showed that the 108 SAGs mainly related to protein processing, nitrogen metabolism, amino acid metabolism, RNA degradation and plant hormone signal transduction. Among the 108 SAGs, seven transcription factors were identified in which a novel bZIP transcription factor MtbZIP60 was proved to inhibit leaf senescence. MtbZIP60 encodes a nuclear-localized protein and possesses transactivation activity. Further study demonstrated MtbZIP60 could associate with MtWRKY40, both of which exhibited an up-regulated expression pattern during leaf senescence, indicating their crucial roles in the regulation of leaf senescence. Our findings help elucidate the molecular mechanisms of leaf senescence in M. truncatula and provide candidates for the genetic improvement of forage crops, with a focus on regulating leaf senescence. Full article
(This article belongs to the Special Issue Advances in Plant Genomics and Genetics)
Show Figures

Figure 1

Figure 1
<p>Physiological and biochemical analysis of leaf growth and senescence in <span class="html-italic">Medicago truncatula</span>. (<b>A</b>) Phenotype of the fifth compound leaf at different growth and senescence stages. The day when the fifth compound leaf emerged is considered as day 1. (<b>B</b>) Chlorophyll content in the fifth compound leaf at 3 d, 5 d, 15 d, 30 d, 45 d and 60 d. The chlorophyll content was obtained by measuring the SPAD value using a chlorophyll meter. Date is shown as mean ± SD (<span class="html-italic">n</span> = 4). Significant differences revealed by Tukey’s multiple comparison test are indicated by letters above bars (<span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) MDA content was measured in the fifth compound leaf at indicated time points. Date is shown as mean ± SD (<span class="html-italic">n</span> = 4). Significant differences revealed by Tukey’s multiple comparison test are indicated by letters above bars (<span class="html-italic">p</span> &lt; 0.05). (<b>D</b>,<b>E</b>) Expression of senescence up-regulated and down-regulated marker genes <span class="html-italic">MtORE1</span> (<b>D</b>) and <span class="html-italic">MtCAB1</span> (<b>E</b>) at 3 d, 5 d, 15 d, 30 d, 45 d and 60 d. Date is shown as mean ± SD (<span class="html-italic">n</span> = 3). Significant differences revealed by Tukey’s multiple comparison test are indicated by letters above bars (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Transcriptomic overview of age-dependent leaf senescence in <span class="html-italic">Medicago truncatula</span>. (<b>A</b>) Principal component analysis (PCA) plot of transcriptome data from different leaf growth and senescence stages; (<b>B</b>) Pearson correlation coefficient of transcriptome profiles from different leaf growth and senescence stages; (<b>C</b>) The number of differentially expressed genes (DEGs) identified from various comparison combination, as determined based on FPKM values using DESeq2 with adjusted <span class="html-italic">p</span>-value &lt; 0.01 and |log<sub>2</sub> (fold change)|&gt;1 or &lt;−1; (<b>D</b>) Venn diagram of overlap DEGs among different comparisons.</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>D</b>) Validation of selected transcripts by qRT-PCR. The expression of the specific gene in RNA-seq was represented by the lines on the right y-axis, whereas the relative expression of the same gene detected by qRT-PCR from independent replicates was depicted by the columns on the left y-axis. Date is shown as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Short Time-series Expression Miner (STEM) clustering of 1057 DEGs shared by the three time points. (<b>A</b>–<b>D</b>) indicate different types of model expression profile, with the model profile ID number on the top left-hand corner and <span class="html-italic">p</span>-value on the bottom left. Different colors represent different significance. Only the model temporal expression profiles that has a significant number of assigned genes compared to the number of expected genes, with a <span class="html-italic">p</span>-value &lt; 0.05, are displayed.</p>
Full article ">Figure 5
<p>Comparison of <span class="html-italic">SAGs</span> among different comparison groups and Gene Ontology (GO) enrichment and Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment analysis. (<b>A</b>) Venn diagram of <span class="html-italic">SAGs</span> in different comparison groups. (<b>B</b>) The GO enrichment analysis of the 108 <span class="html-italic">SAGs</span> shared by all the comparison group. The X-axis indicates different GO terms. The left Y-axis represents the percentage of genes, and the right Y-axis represents the number of genes enriched for the relevant GO terms. (<b>C</b>) KEGG pathway enrichment analysis of the 108 <span class="html-italic">SAGs</span> shared by all the comparison group. The X-axis indicates enrichment factor, which represents the ratio of the proportion of genes annotated to a specific pathway among DEGs to the proportion of genes annotated to the same pathway among all genes. Dot color indicates <span class="html-italic">q</span>-value, which is the <span class="html-italic">p</span>-value corrected by multiple hypotheses testing.</p>
Full article ">Figure 6
<p>MtbZIP60 functions as negative regulator of leaf senescence. (<b>A</b>) Prediction of TFs within the 108 <span class="html-italic">SAGs</span>. (<b>B</b>) Expression trend of <span class="html-italic">MtbZIP60</span> during leaf growth and senescence. (<b>C</b>) Schematic representation of the structure of <span class="html-italic">MtbZIP60</span> and the position of <span class="html-italic">Tnt1</span> in <span class="html-italic">mtbzip60</span>. (<b>D</b>) RT-PCR analysis of <span class="html-italic">MtbZIP60</span> in wild type (R108) and <span class="html-italic">mtbzip60</span>. (<b>E</b>) Detached leaves of WT and <span class="html-italic">mtbzip60</span> in the dark for 5 days. Bars = 1 cm. (<b>F</b>) Measurement of chlorophyll content (SPAD) in (<b>E</b>). (<b>G</b>,<b>H</b>) qRT-PCR detection of <span class="html-italic">MtCAB1</span> (<b>G</b>) and <span class="html-italic">MtORE1</span> (<b>H</b>) transcriptional level in (<b>E</b>). DAT, days after treatment. Date is shown as mean ± SD. Asterisks indicate significant difference from the WT (Student’s <span class="html-italic">t</span> test, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns: no significant). At least three biological replicates were performed.</p>
Full article ">Figure 7
<p>Expression pattern, subcellular localization and transcriptional activity analysis of MtbZIP60. (<b>A</b>) Expression pattern of <span class="html-italic">MtbZIP60</span> in different tissues detected using qRT-PCR. <span class="html-italic">MtActin</span> was used as an internal control. (<b>B</b>) Subcellular localization of GFP and MtbZIP60-GFP in epidermal cells of tobacco leaf. The mRFP-AHL22 fusion was used as nuclear localization marker. Green represents the green fluorescent protein, red represents the red fluorescent protein, and yellow is obtained by combining the two fluorescent proteins. Bars = 20 μm. (<b>C</b>) Transactivation activity analysis of MtbZIP60 in yeast. Full-length CDS of <span class="html-italic">MtbZIP60</span> and <span class="html-italic">TaNAC6</span> were cloned into the pGBKT7 vector and transformed into yeast strain AH109. Transformants harboring different plasmids were inoculated onto selective medium, with BD-TaNAC6 serving as the positive control.</p>
Full article ">Figure 8
<p>MtbZIP60 physically interacts with MtWRKY40 in the regulation of leaf senescence. (<b>A</b>) Interaction of MtbZIP60 and MtWRKY40 in an Y2H assay. Yeast cells were co-transformed with AD-MtbZIP60 and BD-MtWRKY40 and cultured on selective medium. Auxotrophic growth on the SD/-Trp-Leu-His-Ade medium indicates the interaction. (<b>B</b>) Interaction of MtbZIP60 and MtWRKY40 in <span class="html-italic">N.benthamiana</span> leaf using an LCI assay. MtbZIP60 was fused with the n-terminal of LUC to generate MtbZIP60-nLUC. MtWRKY40 was fused with the c-terminal of LUC to generate cLUC-MtWRKY40. Different pairs of constructs were used for infiltrating tobacco leaves. (<b>C</b>) Expression trend of <span class="html-italic">MtWRKY40</span> during leaf growth and senescence.</p>
Full article ">
13 pages, 3951 KiB  
Article
Functional Characterization of the Ciliate Stylonychia lemnae Serotonin N-Acetyltransferase, a Pivotal Enzyme in Melatonin Biosynthesis and Its Overexpression Leads to Peroxidizing Herbicide Tolerance in Rice
by Kyungjin Lee and Kyoungwhan Back
Antioxidants 2024, 13(10), 1177; https://doi.org/10.3390/antiox13101177 - 27 Sep 2024
Abstract
Serotonin N-acetyltransferase (SNAT) is a pivotal enzyme for melatonin biosynthesis in all living organisms. It catalyzes the conversion of serotonin to N-acetylserotonin (NAS) or 5-methoxytrypytamine (5-MT) to melatonin. In contrast to animal- and plant-specific SNAT genes, a novel clade of archaeal [...] Read more.
Serotonin N-acetyltransferase (SNAT) is a pivotal enzyme for melatonin biosynthesis in all living organisms. It catalyzes the conversion of serotonin to N-acetylserotonin (NAS) or 5-methoxytrypytamine (5-MT) to melatonin. In contrast to animal- and plant-specific SNAT genes, a novel clade of archaeal SNAT genes has recently been reported. In this study, we identified homologues of archaeal SNAT genes in ciliates and dinoflagellates, but no animal- or plant-specific SNAT homologues. Archaeal SNAT homologue from the ciliate Stylonychia lemnae was annotated as a putative N-acetyltransferase. To determine whether the putative S. lemnae SNAT (SlSNAT) exhibits SNAT enzyme activity, we chemically synthesized and expressed the full-length SlSNAT coding sequence (CDS) in Escherichia coli, from which the recombinant SlSNAT protein was purified by Ni2+ affinity column chromatography. The recombinant SlSNAT exhibited SNAT enzyme activity toward serotonin (Km = 776 µM) and 5-MT (Km = 246 µM) as substrates. Furthermore, SlSNAT-overexpressing (SlSNAT-OE) transgenic rice plants showed higher levels of melatonin synthesis than wild-type controls. The SlSNAT-OE rice plants exhibited delayed leaf senescence and tolerance against treatment with the reactive oxygen species (ROS)-inducing herbicide butafenacil by decreasing hydrogen peroxide (H2O2) and malondialdehyde (MDA) levels, suggesting that melatonin alleviates ROS production in vivo. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Phylogenetic tree of <span class="html-italic">Stylonichia lemnae SNAT</span> and archaeal ortholog genes. The scale bar represents 0.4 substitutions per site. <span class="html-italic">S. lemnae SNAT</span> is written in bold for emphasis. (<b>B</b>) Amino acid sequence identity and similarity between <span class="html-italic">S. lemnae</span> SNAT and human Naa50 (SNAT). The conserved acetyl-coenzyme-A-binding sites are underlined. Dashes denote gaps. GenBank accession numbers are archaea SNAT (NC_002689), <span class="html-italic">E. coli</span> RimI (WP_137442509), human Naa50 (BAB14397), rice SNAT3 (AK241100), and <span class="html-italic">S. lemnae</span> SNAT (CDW73552).</p>
Full article ">Figure 2
<p>(<b>A</b>) Nucleotide alignment between native (red; CDW73552) and synthetic (blue) <span class="html-italic">S. lemnae SNAT</span>. Identity is denoted by stars. Black letters, amino acids. (<b>B</b>) Modification of <span class="html-italic">S. lemnae SNAT</span> codons. The nucleotide sequence of synthetic <span class="html-italic">S. lemnae SNAT</span> was manually codon optimized with reference to the rice <span class="html-italic">SNAT2</span> codon.</p>
Full article ">Figure 3
<p><span class="html-italic">Escherichia coli</span> expression, affinity purification of SlSNAT recombinant protein, and its enzymatic characteristics. (<b>A</b>) Expression of SlSNAT as a thioredoxin (Trx) fusion protein using a pET32b vector and expression of SlSNAT as an N-terminal His × 6-tagged SlSNAT protein using a pET300 vector. (<b>B</b>) Serotonin <span class="html-italic">N</span>-acetyltransferase enzyme activity (SNAT) as a function of various substrates. The expression of recombinant SlSNAT protein is marked by arrows.</p>
Full article ">Figure 4
<p>SNAT enzyme kinetic analysis. Serotonin <span class="html-italic">N</span>-acetyltransferase enzyme activity as a function of (<b>A</b>) various temperature, (<b>B</b>) various pH, (<b>C</b>) <span class="html-italic">K</span><sub>m</sub> and <span class="html-italic">V</span><sub>max</sub> values for serotonin substrate, (<b>D</b>) <span class="html-italic">K</span><sub>m</sub> and <span class="html-italic">V</span><sub>max</sub> values for 5-methoxytryptamine (5-MT) substrate. Values are means ± SD (n = 3). nd, not detectable.</p>
Full article ">Figure 5
<p>Generation of <span class="html-italic">SlSNAT</span> overexpression transgenic rice and the melatonin content of rice seedlings. (<b>A</b>) RT-PCR analyses of transgenic and wild-type 7-day-old rice seedlings. (<b>B</b>) Melatonin contents of 7-day-old rice seedlings. (<b>C</b>) Photograph of seed length. (<b>D</b>) Photograph of lamina angle in 3-week-old rice seedling. (<b>E</b>) Measurement of lamina angle. WT, wild type; <span class="html-italic">UBQ5</span>, rice ubiquitin 5 gene. GenBank accession number of <span class="html-italic">UBQ5</span> is AK061988. Different letters indicate significant differences among lines (Tukey’s HSD; <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Enhanced senescence tolerance in <span class="html-italic">SlSNAT</span>-overexpressing transgenic rice plants. (<b>A)</b> Photograph of senescence-treated 5-week-old rice leaves. (<b>B</b>) Chlorophyll contents in senescence-treated rice leaves. (<b>C</b>) Malondialdehyde (MDA) contents. (<b>D</b>) Gene expression profiles of senescence marker genes by RT-PCR. (<b>E</b>) Gene expression profiles of senescence marker genes by quantitative RT-PCR. Fourth and fifth leaves from 5-week-old rice plants grown in soil pots were detached and this was followed by senescence treatment for 12 days. WT, wild type; <span class="html-italic">UBQ5</span>, rice ubiquitin 5 gene. GenBank accession numbers are <span class="html-italic">Osl2</span> (AF251073), <span class="html-italic">Osl20</span> (AF251067), <span class="html-italic">Osl185</span> (AF251075), and <span class="html-italic">UBQ5</span> (AK061988). Different letters indicate significant differences among the lines (Tukey’s HSD; <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Enhanced tolerance of SlSNAT-overexpressing transgenic rice plants against peroxidizing herbicide butafenacil. (<b>A</b>) Photograph of rice seedlings after butafenacil treatment. (<b>B</b>) Effect of butafenacil treatment on cellular leakage. (<b>C</b>) MDA production from butafenacil-treated rice seedlings. (<b>D</b>) H<sub>2</sub>O<sub>2</sub> content from butafenacil-treated rice seedlings. Seven-day-old rice seedlings were challenged with 0.1 µM butafenacil for 48 h. WT, wild type; FW, fresh weight. Different letters indicate significant differences among the lines (Tukey’s HSD; <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Sequence comparison and phylogenetic tree of SNAT in the Ciliophora and dinoflagellates. (<b>A</b>) Consensus amino acid sequences among three SNAT proteins including the human Naa50, the ciliate <span class="html-italic">Stylonichia lemnae</span> SNAT, and the dinoflagellate <span class="html-italic">Polarella glacialis</span> SNAT. Bold red letters indicate consensus amino acids. Dashes denote gaps for maximizing alignment of conserved residues. A coenzyme-A-binding pocket is underlined. (<b>B</b>) Phylogenetic tree analysis of SNAT proteins from the ciliates and dinoflagellates. GenBank accession numbers of various <span class="html-italic">SNAT</span> genes are as follows: human <span class="html-italic">Naa50</span> (BAB14397), <span class="html-italic">Cladocopium goreaui</span> (CAI3999280); <span class="html-italic">Effrenium voratum</span> (CAJ1361560); <span class="html-italic">Polarella glacialis</span> (CAK0876941); <span class="html-italic">Stylonichia lemna</span> (CCKQ01002460); <span class="html-italic">Paramecium sonneborni</span> (CAD8056267); <span class="html-italic">Pseudocohnilembus persalius</span> (KRX00195); <span class="html-italic">Tetrahymena thermophila</span> SB210 (XP_001025216); <span class="html-italic">Ichthyophthirius multifiliis</span> (XP-004035125). The scale bar represents 0.3 substitutions per site.</p>
Full article ">
13 pages, 1953 KiB  
Article
A Maize Mutant Impaired in SL Biosynthesis (zmccd8) Shows a Lower Growth, an Altered Response to Nitrogen Starvation, and a Potential Secondary Effect on Drought Tolerance
by Laura Ravazzolo, Andrea Chichi, Franco Meggio, Leonardo Buzzicotti, Benedetto Ruperti, Serena Varotto, Mario Malagoli and Silvia Quaggiotti
Stresses 2024, 4(4), 614-626; https://doi.org/10.3390/stresses4040039 - 25 Sep 2024
Abstract
Strigolactones (SLs) are essential phytohormones involved in plant development and interaction with the rhizosphere, regulating shoot branching, root architecture, and leaf senescence for nutrient reallocation. The Zea mays L. zmccd8 mutant, defective in SL biosynthesis, shows various architectural changes and reduced growth. This [...] Read more.
Strigolactones (SLs) are essential phytohormones involved in plant development and interaction with the rhizosphere, regulating shoot branching, root architecture, and leaf senescence for nutrient reallocation. The Zea mays L. zmccd8 mutant, defective in SL biosynthesis, shows various architectural changes and reduced growth. This study investigates zmccd8 and wild-type (WT) maize plants under two nutritional treatments (N-shortage vs. N-provision as urea). Morphometric analysis, chlorophyll and anthocyanin indexes, drought-related parameters, and gene expression were measured at specific time points. The zmccd8 mutant displayed reduced growth, such as shorter stems, fewer leaves, and lower kernel yield, regardless of the nutritional regime, confirming the crucial role of SLs. Additionally, zmccd8 plants exhibited lower chlorophyll content, particularly under N-deprivation, indicating SL necessity for proper senescence and nutrient mobilization. Increased anthocyanin accumulation in zmccd8 under N-shortage suggested a stress mitigation attempt, unlike WT plants. Furthermore, zmccd8 plants showed signs of increased water stress, likely due to impaired stomatal regulation, highlighting SLs role in drought tolerance. Molecular analysis confirmed higher expression of SL biosynthesis genes in WT under N-shortage, while zmccd8 lacked this response. These findings underscore SL importance in maize growth, stress responses, and nutrient allocation, suggesting potential agricultural applications for enhancing crop resilience. Full article
(This article belongs to the Topic Plant Responses to Environmental Stress)
Show Figures

Figure 1

Figure 1
<p>Phenotypic analysis of stem height (<b>A</b>), leaf number (<b>B</b>), internode length (<b>C</b>), stem circumference (<b>D</b>), kernel weight (<b>E</b>), and ears length (<b>F</b>) for wild-type (WT) and <span class="html-italic">zmccd8</span> mutant plants at different days after sowing (DAS) under two N treatments. Error bars represent the mean ± SE (<span class="html-italic">n</span> = 24). At 58 DAS, urea was provided as the N source (dashed red line). Different letters indicate significant differences (at <span class="html-italic">p</span> &lt; 0.05 according to LSD test) at each DAS. Based on ANOVA, the significance of F values was reported as follows: ‘***’ <span class="html-italic">p</span> &lt; 0.001; ‘**’ <span class="html-italic">p</span> &lt; 0.01; ‘*’ <span class="html-italic">p</span> &lt; 0.05; no asterisks <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 2
<p>Profiles in chlorophyll content (<b>A</b>) and anthocyanin levels (<b>B</b>) in four different groups of maize leaves (L1-2; L3-4-5; L6-7-8; L9-10-11). Error bars represent the mean of six biological replicates ± SE. At 58 DAS, urea was provided as the N source (dashed red line). Different letters indicate significant differences (at <span class="html-italic">p</span> &lt; 0.05 according to LSD test) at each DAS. Based on ANOVA, the significance of F values was reported as follows: ‘***’ <span class="html-italic">p</span> &lt; 0.001; ‘**’ <span class="html-italic">p</span> &lt; 0.01; ‘*’ <span class="html-italic">p</span> &lt; 0.05; no asterisks <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 3
<p>Profiles of stomatal conductance (gsw, mol H<sub>2</sub>O m<sup>−2</sup>s<sup>−1</sup>) (<b>A</b>), leaf transpiration (E-app, mol H<sub>2</sub>O m<sup>−2</sup>s<sup>−1</sup>) (<b>B</b>), and photosystem II efficiency (PhiPS II) (<b>C</b>) in the group of leaves 9-10-11. Error bars represent the mean of six biological replicates ± SE. At 58 DAS, urea was provided as the N source (dashed red line). Different letters indicate significant differences (at <span class="html-italic">p</span> &lt; 0.05 according to LSD test) at each DAS. Based on ANOVA, the significance of F values was reported as follows: ‘***’ <span class="html-italic">p</span> &lt; 0.001; ‘*’ <span class="html-italic">p</span> &lt; 0.05; no asterisks <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 4
<p>Relative gene expression of three genes involved in SL biosynthesis (<span class="html-italic">CCD7</span>, <span class="html-italic">CCD8</span>), signalling (<span class="html-italic">MAX2</span>), and drought stress (<span class="html-italic">SULTR6</span>) in leaf samples at three different days after sowing (DAS). Data are means ± SE for three biological replicates. Different letters indicate significant differences (at <span class="html-italic">p</span> &lt; 0.05 according to LSD test) at each DAS. Based on ANOVA, the significance of F values was reported as follows: ‘***’ <span class="html-italic">p</span> &lt; 0.001; ‘**’ <span class="html-italic">p</span> &lt; 0.01; ‘*’ <span class="html-italic">p</span> &lt; 0.05; no asterisks <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">
19 pages, 1221 KiB  
Article
Growth, Photosynthesis and Yield Responses of Common Wheat to Foliar Application of Methylobacterium symbioticum under Decreasing Chemical Nitrogen Fertilization
by Francesco Valente, Anna Panozzo, Francesco Bozzolin, Giuseppe Barion, Pranay Kumar Bolla, Vittorio Bertin, Silvia Potestio, Giovanna Visioli, Yu Wang and Teofilo Vamerali
Agriculture 2024, 14(10), 1670; https://doi.org/10.3390/agriculture14101670 - 24 Sep 2024
Abstract
Current agriculture intensifies crop cultivation to meet food demand, leading to unsustainable use of chemical fertilizers. This study investigates a few physiological and agronomic responses of common wheat following the inoculation with plant growth-promoting bacteria to reduce nitrogen inputs. A field trial was [...] Read more.
Current agriculture intensifies crop cultivation to meet food demand, leading to unsustainable use of chemical fertilizers. This study investigates a few physiological and agronomic responses of common wheat following the inoculation with plant growth-promoting bacteria to reduce nitrogen inputs. A field trial was conducted in 2022–2023, in Legnago (Verona, Italy) on Triticum aestivum var. LG-Auriga comparing full (180 kg ha−1) and reduced (130 kg ha−1) N doses, both with and without foliar application at end tillering of the N-fixing bacterium Methylobacterium symbioticum. Biofertilization did not improve shoot growth, while it seldom increased the root length density in the arable layer. It delayed leaf senescence, prolonged photosynthetic activity, and amplified stomatal conductance and PSII efficiency under the reduced N dose. Appreciable ACC-deaminase activity of such bacterium disclosed augmented nitrogen retrieval and reduced ethylene production, explaining the ameliorated stay-green. Yield and test weight were unaffected by biofertilization, while both glutenin-to-gliadin and HMW-to-LMW ratios increased together with dough tenacity. It is concluded that Methylobacterium symbioticum can amplify nitrogen metabolism at a reduced nitrogen dose, offering a viable approach to reduce chemical fertilization under suboptimal growing conditions for achieving a more sustainable agriculture. Further research over multiple growing seasons and soil types is necessary to corroborate these preliminary observations. Full article
Show Figures

Figure 1

Figure 1
<p>Dynamics of leaf chlorophyll content (as SPAD values) and canopy greenness (NDVI) (mean ± S.E.; n = 3) in wheat under “100%N”, “100%N + bact”, “75%N”, and “75%N + bact” treatments during 2023. Percentages: variation vs. reference treatment 100%N (Ref.). Letters: statistical comparison among treatments (Tukey test, <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 2
<p>Photosynthetic parameters: CO<sub>2</sub> assimilation rate (A), stomatal conductance (gsw), transpiration rate (Emm), and PSII efficiency (<span class="html-italic">Fv</span>′/<span class="html-italic">Fm</span>′) (mean ± S.E.; n = 3) in wheat plants under “100%N”, “100%N + bact”, “75%N”, and “75%N + bact” treatments. Percentages: variation vs. reference treatment 100%N (Ref.). Letters: statistical comparison among treatments (Tukey test, <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 3
<p>Root length density (RLD), root area density (RAD), and root diameter (mean; n = 3) in wheat plants under “100%N”, “100%N + bact”, “75%N”, and “75%N + bact” treatments at various soil depths. Asterisk (*): statistically significant differences among treatments (Tukey test, <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 4
<p>Qualitative characteristics of wheat grains: grain protein content (GPC), gliadins, glutenins, gliadin-to-glutenin ratio (GLI/GLU), and high-to-low molecular weight glutenin subunits ratio (HMW/LMW) (mean ± S.E.; n = 3) in wheat plants under “100%N”, “100%N + bact”, “75%N”, and “75%N + bact” treatments. Percentages: variation vs. reference treatment 100%N (Ref.). Letters: statistical comparison among treatments (Tukey test, <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">
12 pages, 3281 KiB  
Article
Responses of Leaf Senescence for Stipa krylovii to Interactive Environmental Factors
by Xingyang Song and Guangsheng Zhou
Agronomy 2024, 14(9), 2145; https://doi.org/10.3390/agronomy14092145 - 20 Sep 2024
Abstract
The effects of temperature, and photoperiod on autumn phenology are well established for many species. However, the impact of multiple environmental factors and their interactions on regulating autumn phenology remains insufficiently explored. A large-scale controlled experiment in an artificial climate chamber was conducted [...] Read more.
The effects of temperature, and photoperiod on autumn phenology are well established for many species. However, the impact of multiple environmental factors and their interactions on regulating autumn phenology remains insufficiently explored. A large-scale controlled experiment in an artificial climate chamber was conducted from April to October 2021 at the Hebei Gucheng Agricultural Meteorology National Observation and Research Station, Hebei Province. This study aimed to investigate the interactive effects of temperature [T1.5, (1.5 °C above the control), T2, (2 °C above the control)], photoperiod [LP, long photoperiod (4 h photoperiod above the control), SP, short photoperiod (4 h photoperiod below the control)], and nitrogen addition [LN, low nitrogen, (nitrogen at 5 g N·m−2·a−1), MN, medium nitrogen, (nitrogen at 10 g N·m−2·a−1), HN, high nitrogen, (nitrogen at 20 g N·m−2·a−1), control for temperature and photoperiod was the mean monthly temperature and average photoperiod (14 h) from 1989–2020 for Stipa krylovii, while the control for nitrogen treatment was without nitrogen addition] on leaf senescence in Stipa krylovii. A three-way analysis of variance (ANOVA) revealed significant effects of temperature, photoperiod, and nitrogen addition on leaf senescence (p < 0.01), with effects varying across different levels of each factor. Increased temperature notably delayed leaf senescence, with delays averaging of 4.0 and 6.3 days for T1.5 and T2, respectively. The LP treatment advanced leaf senescence by an average of 4.0 days, while the SP treatment delayed it by an average of 6.2 days; nitrogen addition advanced leaf senescence, with the effect intensifying as nitrogen levels increased, resulting in average advancements of 1.5, 1.9, and 4.3 days for LN, MN, and HN, respectively. Additionally, we observed that temperature altered the sensitivity of leaf senescence to the photoperiod, diminishing the advancement caused by LP at 2 °C and amplifying the delay caused by SP. These findings underscore the differential impacts of these three factors on the leaf senescence of Stipa krylovii and provide critical insights into plant phenology in response to varying environmental conditions. Full article
(This article belongs to the Section Agroecology Innovation: Achieving System Resilience)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The three large artificial climate chambers; (<b>b</b>) six cubicles in each chamber separated by wooden boards and opaque curtains; (<b>c</b>) sodium lamp.</p>
Full article ">Figure 2
<p>Variation days of the leaf senescence of <span class="html-italic">Stipa krylovii</span> in temperature (<b>a</b>), photoperiod (<b>b</b>) and N addition (<b>c</b>). * means significant difference compared with control at 0.05 level (<span class="html-italic">p</span> &lt; 0.05). Variation days in figures were calculated by subtracting the corresponding control values from the treatment values. Negative values indicate an advancement in the leaf senescence phase, while positive values indicate a delay.</p>
Full article ">Figure 3
<p>Variation days of the leaf senescence of <span class="html-italic">Stipa krylovii</span> affected by interactive temperature and photoperiod. LCO: Leaf coloration onset stage, LCG: leaf coloration in general stage, LCF: leaf fully coloring stage. T1.5, 1.5 °C above the control, T2, 2 °C above the control, LP: long photoperiods, SP: short photoperiods. * means significant difference compared with control at 0.05 level (<span class="html-italic">p</span> &lt; 0.05). Variation days in the figure were calculated by subtracting the corresponding control values from the treatment values. Negative values indicate an advancement in the leaf senescence phase, while positive values indicate a delay.</p>
Full article ">Figure 4
<p>Variation days of the leaf senescence of <span class="html-italic">Stipa krylovii</span> affected by interactive temperature and N addition. LCO: Leaf coloration onset stage, LCG: leaf coloration in general stage, LCF: leaf fully coloring stage. T1.5, 1.5 °C above the control, T2, 2 °C above the control, LN: low nitrogen addition, MN: medium nitrogen addition, HN: high nitrogen addition. * means significant difference compared with control at 0.05 level (<span class="html-italic">p</span> &lt; 0.05). Variation days in the figure were calculated by subtracting the corresponding control values from the treatment values. Negative values indicate an advancement in the leaf senescence phase, while positive values indicate a delay.</p>
Full article ">Figure 5
<p>Variation days of the leaf senescence of <span class="html-italic">Stipa krylovii</span> affected by the interactive photoperiod and N addition. LCO: Leaf coloration onset stage, LCG: leaf coloration in general stage, LCF: leaf fully coloring stage. LP: long photoperiods, SP: short photoperiods, LN: low nitrogen addition, MN: medium nitrogen addition, HN: high nitrogen addition. * means significant difference compared with control at 0.05 level (<span class="html-italic">p</span> &lt; 0.05). Variation days in the figure were calculated by subtracting the corresponding control values from the treatment values. Negative values indicate an advancement in the leaf senescence phase, while positive values indicate a delay.</p>
Full article ">Figure 6
<p>Variation days of the leaf senescence of <span class="html-italic">Stipa krylovii</span> from the interactive effects of temperature, photoperiod, and N addition. LCO: Leaf coloration onset stage, LCG: leaf coloration in general stage, LCF: leaf fully coloring stage.T1.5, 1.5 °C above the control, T2, 2 °C above the control, LP: long photoperiods, SP: short photoperiods, LN: low nitrogen addition, MN: medium nitrogen addition, HN: high nitrogen addition. * means significant difference compared with control at 0.05 level (<span class="html-italic">p</span> &lt; 0.05). Variation days in the figure were calculated by subtracting the corresponding control values from the treatment values. Negative values indicate an advancement in the leaf senescence phase, while positive values indicate a delay.</p>
Full article ">
24 pages, 3983 KiB  
Article
Comparative Analysis of Phytohormone Biosynthesis Genes Responses to Long-Term High Light in Tolerant and Sensitive Wheat Cultivars
by Zhi-Ang Li, Muhammad Fahad, Wan-Chang Li, Leeza Tariq, Miao-Miao Liu, Ya-Nan Liu and Tai-Xia Wang
Plants 2024, 13(18), 2628; https://doi.org/10.3390/plants13182628 - 20 Sep 2024
Abstract
Phytohormones are vital for developmental processes, from organ initiation to senescence, and are key regulators of growth, development, and photosynthesis. In natural environments, plants often experience high light (HL) intensities coupled with elevated temperatures, which pose significant threats to agricultural production. However, the [...] Read more.
Phytohormones are vital for developmental processes, from organ initiation to senescence, and are key regulators of growth, development, and photosynthesis. In natural environments, plants often experience high light (HL) intensities coupled with elevated temperatures, which pose significant threats to agricultural production. However, the response of phytohormone-related genes to long-term HL exposure remains unclear. Here, we examined the expression levels of genes involved in the biosynthesis of ten phytohormones, including gibberellins, cytokinins, salicylic acid, jasmonic acid, abscisic acid, brassinosteroids, indole-3-acetic acid, strigolactones, nitric oxide, and ethylene, in two winter wheat cultivars, Xiaoyan 54 (XY54, HL tolerant) and Jing 411 (J411, HL sensitive), when transferred from low light to HL for 2–8 days. Under HL, most genes were markedly inhibited, while a few, such as TaGA2ox, TaAAO3, TaLOG1, and TaPAL2, were induced in both varieties. Interestingly, TaGA2ox2 and TaAAO3 expression positively correlated with sugar content but negatively with chlorophyll content and TaAGP expression. In addition, we observed that both varieties experienced a sharp decline in chlorophyll content and photosynthesis performance after prolonged HL exposure, with J411 showing significantly more sensitivity than XY54. Hierarchical clustering analysis classified the phytohormone genes into the following three groups: Group 1 included six genes highly expressed in J411; Group 2 contained 25 genes drastically suppressed by HL in both varieties; and Group 3 contained three genes highly expressed in XY54. Notably, abscisic acid (ABA), and jasmonic acid (JA) biosynthesis genes and their content were significantly higher, while gibberellins (GA) content was lower in XY54 than J411. Together, these results suggest that the differential expression and content of GA, ABA, and JA play crucial roles in the contrasting responses of tolerant and sensitive wheat cultivars to leaf senescence induced by long-term HL. This study enhances our understanding of the mechanisms underlying HL tolerance in wheat and can guide the development of more resilient wheat varieties. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) total chlorophyll (Chl a + b); (<b>B</b>) chlorophyll a (Chl a); (<b>C</b>) chlorophyll b (Chl b); (<b>D</b>) carotenoid (Car) contents of wheat under sustained strong light stress; (<b>E</b>) changes in the ratio of chlorophyll a and chlorophyll b (Chl a/b); and (<b>F</b>) the ratio of total chlorophyll to carotenoid content (Chl a + b/Car). Data are expressed as mean ± standard deviation, and different letters indicate a significant level of <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>(<b>A</b>) maximum light quantum yield (F<sub>v</sub>/F<sub>m</sub>); (<b>B</b>) performance index (PI); (<b>C</b>) energy captured per unit reaction center (TR<sub>o</sub>/CS) of photosystem II under continuous strong light stress; (<b>D</b>) the energy transferred by electrons per unit reaction center (ET<sub>o</sub>/CS); (<b>E</b>) the energy dissipated per unit reflection center (DI<sub>o</sub>/CS); and (<b>F</b>) the number of open reaction centers per unit area (RC/CS<sub>m</sub>) response to bright light. Data are expressed as mean ± standard deviation, and different letters indicate a significant level of <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Expression responses of six GA biosynthesis-related genes encoding: (<b>A</b>) <span class="html-italic">ent-copalyl diphosphate synthase 1</span> (<span class="html-italic">TaCPS1</span>); (<b>B</b>) <span class="html-italic">ent-kaurene oxidase (TaKO)</span>; (<b>C</b>) <span class="html-italic">ent-kaurenoic acid oxidase 1 (TaKAO1)</span>; (<b>D</b>) <span class="html-italic">GA20-oxidase 2 (TaGA20ox2)</span>; (<b>E</b>) <span class="html-italic">GA3-oxidase 1 (TaGA3ox1)</span>; and (<b>F</b>) <span class="html-italic">GA2-oxidase (TaGA2ox)</span> in XY54 and J411 to leaf senescence induced by HL. The data are represented as mean ± SE. The different letters indicate a significant difference at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Eexpression responses of five JA biosynthesis-related genes encoding: (<b>A</b>) phospholipase A1 (<span class="html-italic">TaPLA1</span>); (<b>B</b>) lipoxygenase (<span class="html-italic">TaLOX</span>); (<b>C</b>) allene oxide synthase 2 (<span class="html-italic">TaAOS2</span>); (<b>D</b>) allene oxide cyclase 1 (<span class="html-italic">TaAOC1</span>); and (<b>E</b>) 12-oxophytodienoate 2 (<span class="html-italic">TaOPR2</span>); and as well as (<b>F</b>) JA content in XY54 and J411 responding to leaf senescence induced by HL. The data are represented as mean ± SE. The different letters indicate a significant difference at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Expressional responses of six ABA biosynthesis-related genes encoding: (<b>A</b>) <span class="html-italic">β-carotene hydroxylase (TaBCH1)</span>; (<b>B</b>) <span class="html-italic">zeaxanthin epoxidase (TaABA1)</span>; (<b>C</b>) <span class="html-italic">ABA DEFICIENT 4 (TaABA4)</span>; (<b>D</b>) <span class="html-italic">9-cis-epoxy carotenoid dioxygenase (TaNCED1)</span>; (<b>E</b>) <span class="html-italic">short-chain alcohol dehydrogenase (TaABA2)</span>; and (<b>F</b>) <span class="html-italic">abscisic aldehyde oxidase (TaAAO3)</span> in XY54 and J411 to leaf senescence induced by HL. The data are represented as mean ± SE. The different letters indicate a significant difference at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Expressional responses of CK and ET biosynthesis-related genes encoding: (<b>A</b>) <span class="html-italic">tRNA isopentenyl transferase 2 (TaIPT2)</span>; (<b>B</b>) <span class="html-italic">cytokinin hydroxylase (TaCYP735A1)</span>; (<b>C</b>) <span class="html-italic">cytokinin riboside 5′-monophosphate phosphoribohydrolase 1 (TaLOG1)</span>; (<b>D</b>) <span class="html-italic">S-adenosylmethionine synthetase 4 (TaSAM4)</span>; (<b>E</b>) <span class="html-italic">1-aminocyclopropane-1-carboxylic acid synthase 7 (TaACS7)</span>; and (<b>F</b>) <span class="html-italic">1-aminocyclopropane-1-carboxylic acid oxidase 1 (TaACO1)</span> in XY54 and J411 to leaf senescence induced by HL. The data are represented as mean ± SE. The different letters indicate a significant difference at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Expressional responses of SA and BR biosynthesis-related genes encoding: (<b>A</b>) <span class="html-italic">isochorismate synthase (TaICS2)</span>; (<b>B</b>) <span class="html-italic">chorismate mutase (TaCM1)</span>; (<b>C</b>) <span class="html-italic">phenylalanine ammonia-lyase (TaPAL1);</span> (<b>D</b>) <span class="html-italic">cytochrome P450 90A1 (TaCYP90A1)</span>; (<b>E</b>) <span class="html-italic">steroid 5-α-reductase (TaDET2)</span>; and (<b>F</b>) <span class="html-italic">cytochrome P450 90D2 (TaCYP90D2)</span> in XY54 and J411 to leaf senescence induced by HL. The data are represented as mean ± SE. The different letters indicate a significant difference at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Expressional responses of indole-3-acetic acid, nitric oxide, and strigolactone biosynthesis-related genes encoding: (<b>A</b>) <span class="html-italic">L-tryptophan--pyruvate aminotransferase 1</span> (<span class="html-italic">TAA1</span>); (<b>B</b>) <span class="html-italic">indole-3-pyruvate monooxygenase YUCCA9 (TaYUC9</span>); (<b>C</b>) <span class="html-italic">nitric oxide synthase 1</span> (<span class="html-italic">TaNOS1</span>); (<b>D</b>) <span class="html-italic">β-carotene isomerase D27</span> (<span class="html-italic">TaD27</span>); (<b>E</b>) <span class="html-italic">carotenoid cleavage dioxygenase 7</span> (<span class="html-italic">TaMAX3</span>); and (<b>F</b>) <span class="html-italic">cytochrome P450 711A1</span> (<span class="html-italic">TaMAX1</span>) in XY54 and J411 to leaf senescence induced by HL. The data are represented as mean ± SE. The different letters indicate a significant difference at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 9
<p>Hierarchical clustering of the expression of 35 phytohormones biosynthesis genes in XY54 and J411 subjected to HL for 2–8 d. The yellow color indicates higher expression levels of the corresponding genes, while blue indicates lower expression levels and black represents no significant change in gene expression. The clustering of genes into groups 1, 2, and 3 further highlights patterns of expression that are consistent within these groups under the experimental conditions.</p>
Full article ">
29 pages, 8791 KiB  
Article
Leaf Physiological Responses and Early Senescence Are Linked to Reflectance Spectra in Salt-Sensitive Coastal Tree Species
by Steven M. Anderson, Emily S. Bernhardt, Jean-Christophe Domec, Emily A. Ury, Ryan E. Emanuel, Justin P. Wright and Marcelo Ardón
Forests 2024, 15(9), 1638; https://doi.org/10.3390/f15091638 - 17 Sep 2024
Abstract
Salt-sensitive trees in coastal wetlands are dying as forests transition to marsh and open water at a rapid pace. Forested wetlands are experiencing repeated saltwater exposure due to the frequency and severity of climatic events, sea-level rise, and human infrastructure expansion. Understanding the [...] Read more.
Salt-sensitive trees in coastal wetlands are dying as forests transition to marsh and open water at a rapid pace. Forested wetlands are experiencing repeated saltwater exposure due to the frequency and severity of climatic events, sea-level rise, and human infrastructure expansion. Understanding the diverse responses of trees to saltwater exposure can help identify taxa that may provide early warning signals of salinity stress in forests at broader scales. To isolate the impacts of saltwater exposure on trees, we performed an experiment to investigate the leaf-level physiology of six tree species when exposed to oligohaline and mesohaline treatments. We found that species exposed to 3–6 parts per thousand (ppt) salinity had idiosyncratic responses of plant performance that were species-specific. Saltwater exposure impacted leaf photochemistry and caused early senescence in Acer rubrum, the most salt-sensitive species tested, but did not cause any impacts on plant water use in treatments with <6 ppt. Interestingly, leaf spectral reflectance was correlated with the operating efficiency of photosystem II (PSII) photochemistry in A. rubrum leaves before leaf physiological processes were impacted by salinity treatments. Our results suggest that the timing and frequency of saltwater intrusion events are likely to be more detrimental to wetland tree performance than salinity concentrations. Full article
(This article belongs to the Special Issue Coastal Forest Dynamics and Coastline Erosion, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Plant communities in coastal, non-tidal freshwater wetlands are comprised of salt-sensitive forest, shrub–scrub, and salt-tolerant marsh species. Soil salinization varies in response to heterogeneous microtopography and response to gradients in elevation and vegetation abundance from estuarine waters to tree-dominated wetlands. Coastal tree species that dominate forested wetlands have species-specific sensitivity to salinization and are therefore susceptible to ecosystem transition. Several of these tree species included in this study are deciduous hardwoods (<span class="html-italic">Acer rubrum</span>, <span class="html-italic">Nyssa sylvatica</span>, and <span class="html-italic">Quercus nigra</span>) and evergreen or deciduous conifers (<span class="html-italic">Juniperus virginiana</span>, <span class="html-italic">Taxodium distichum</span>, and <span class="html-italic">Pinus taeda</span>). The background image is a modification of a photograph taken by S. Anderson in Swan Quarter National Wildlife Refuge in North Carolina, USA.</p>
Full article ">Figure 2
<p>(<b>a</b>) Conceptual drawing of one basin with one tree of each of the six tree species and the groundwater level (8–16 cm) that was maintained in each of the 22 basins from the greenhouse study. (<b>b</b>) A schematic of the number of replicate trees within each experimental treatment (in parts per thousands; ppt) for each species. The two primary treatments (control and 3 ppt) highlighted in grey ranging from 4–8 replicates per species and individuals for intermediate and high salinity treatments for regression analysis only. All 118 trees were grown in the same greenhouse and the same basin setup.</p>
Full article ">Figure 3
<p>Specific conductance (µs/cm) of (<b>a</b>) basin groundwater and (<b>b</b>,<b>c</b>) tree pot soils measured from 1 May to 26 September 2018, averaged across all experimental basins within each treatment over the entire 6-month experiment. Soil-specific conductance measured at (<b>b</b>) 5 cm and (<b>c</b>) 10 cm depths from the soil surface. The two replicated treatments are in color: control (green) and orange (3 ppt). All other intermediate (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) treatments are in a gradient of grey circles and dotted lines (0.5 and 4 ppt), squares and dashed lines (1 and 5 ppt), or triangles and solid lines (2 and 6 ppt). Non-replicated intermediate treatments (in grey) are shown to highlight trends only. Error bars are the standard deviation for both replicated treatments (control and 3 ppt). The vertical shaded bars in the soil conductance (<b>b</b>,<b>c</b>) noted by the asterisks above the figure indicate salt additions, or pulses, added to the groundwater of each treatment basin.</p>
Full article ">Figure 3 Cont.
<p>Specific conductance (µs/cm) of (<b>a</b>) basin groundwater and (<b>b</b>,<b>c</b>) tree pot soils measured from 1 May to 26 September 2018, averaged across all experimental basins within each treatment over the entire 6-month experiment. Soil-specific conductance measured at (<b>b</b>) 5 cm and (<b>c</b>) 10 cm depths from the soil surface. The two replicated treatments are in color: control (green) and orange (3 ppt). All other intermediate (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) treatments are in a gradient of grey circles and dotted lines (0.5 and 4 ppt), squares and dashed lines (1 and 5 ppt), or triangles and solid lines (2 and 6 ppt). Non-replicated intermediate treatments (in grey) are shown to highlight trends only. Error bars are the standard deviation for both replicated treatments (control and 3 ppt). The vertical shaded bars in the soil conductance (<b>b</b>,<b>c</b>) noted by the asterisks above the figure indicate salt additions, or pulses, added to the groundwater of each treatment basin.</p>
Full article ">Figure 4
<p>Summary table of effect sizes (Hedge’s g) between control and 3 ppt salinity treatments for plant performance parameters including leaf mass and area (LM and LA), specific leaf area (SLA), leaf dry matter content (LDMC), leaf percent carbon and nitrogen (%C and %N), leaf carbon–nitrogen ratio (C:N), aboveground biomass (AGB), belowground biomass (BGB), height and diameter growth rates (HGR and DGR), and root-to-shoot ratio/root mass fraction (RSR). Dark-shaded cells indicate a large mean increase or decrease in effect sizes compared to controls. Light-shaded cells indicate a small or medium effect size compared to control. If differences in means were significant (<span class="html-italic">p</span> ≤ 0.05), values are bold. White cells indicate no difference.</p>
Full article ">Figure 5
<p>(<b>a</b>) The relationship between osmotic potential (Ψ<sub>π</sub>, MPa) of the groundwater treatment solutions (NaCl) after the second salinity dose and the difference between Ψ<sub>π</sub> and predawn leaf water potential (Ψ<sub>pd</sub>, MPa) across all species and salinity treatments. Each color and shape combination represents Ψ<sub>π</sub> and Ψ<sub>pd</sub> for each species. The horizontal arrow (color gradient from green to red) represents the direction of salinity increase as Ψ<sub>π</sub> becomes more negative. The vertical arrow in the panel shows that as Ψ<sub>pd</sub> decreases in tandem with a decrease in Ψ<sub>π</sub>, hydrological flow between the soil and roots reaches equilibrium and ceases to move resulting in leaf loss and mortality. The points closest to zero on the x-axis are controls, and salinity treatments increase as they move to the right (more negative Ψ<sub>π</sub>) noted by the green-to-red gradient arrow. (<b>b</b>) The relationship between osmotic potential (Ψ<sub>π</sub>) of the groundwater treatment solutions (NaCl) and the difference between midday water potential (Ψ<sub>md</sub>) and Ψ<sub>pd</sub> across all species and salinity treatments. As Ψ<sub>π</sub> increases (salinity treatments become more saline), hydrological flow between the soil (Ψ<sub>pd</sub>) and light-adapted leaves (Ψ<sub>md</sub>) reach equilibrium (Ψ<sub>md</sub> − Ψ<sub>pd</sub> = 0; noted by “No H<sub>2</sub>O Transport”). Equilibrium is expected at high soil salinity meaning water would cease to move to more negative pressure potential in the leaves.</p>
Full article ">Figure 6
<p>Reflectance spectra from <span class="html-italic">Acer rubrum</span> leaves in October 2018 for select salinity treatments (control = dark green, 1 ppt = light green, 3 ppt = yellow, 4 ppt = orange, 6 ppt = red) showing the wavelengths expected to correlate with salinity and drought stress (531, 570, 680, and 800 nanometers). Photochemical reflectance index (PRI) is derived from narrowband reflectance at 531 and 570 nanometers (nm), and normalized difference vegetation index (NDVI) derived from red and near-infrared calculated as NDVI = (NIR − Red)/(NIR + Red), which is reflected over the incoming radiation. Here, we used the reflectance values at 680 (red) and 800 nm (infrared) to calculated NDVI. An NDVI close to 0–0.1 corresponds to no vegetation, while NDVI close to 0.8–0.9, as seen here in the control treatment (NDVI = 0.827), indicates the highest possible density of green leaves.</p>
Full article ">Figure A1
<p>Soil nutrients (NO<sub>3</sub><sup>−</sup>, NH<sub>4</sub><sup>+</sup>, and PO<sub>4</sub><sup>−</sup>) across binned intermediate salinity treatments. Non-replicated treatments were put into two groups, low (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) salinity levels. Controls are green and become more orange with increase in salinity. Ammonium concentrations in soils are significantly greater at salinity ≥3 ppt. Bars are colored by control (dark green), low (light green), 3 ppt (tan), and high (orange) salinity treatments. Asterisks indicate statistically significant differences of salinity treatments to controls Asterisks indicate statistical significance (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure A2
<p>(<b>a</b>) Total whole-plant leaf area (cm<sup>2</sup>) by species and binned treatments. Log transformed whole-plant leaf area of (<b>b</b>) <span class="html-italic">Acer rubrum</span> (<span class="html-italic">p</span> = 0.039) and (<b>c</b>) <span class="html-italic">Pinus taeda</span> (<span class="html-italic">p</span> = 0.046) along the soil sodium gradient. Non-replicated treatments were put into two groups, low (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) salinity levels. Bars are colored by control (dark green), low (light green), 3 ppt (tan), and high (orange) salinity treatments. Asterisks indicate statistical significance (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure A3
<p>(<b>a</b>,<b>b</b>) Cumulative leaf loss by species and (<b>c</b>) proportion of <span class="html-italic">Acer rubrum</span> litter fall over the course of this experiment compared to (<b>d</b>) groundwater-specific conductance (µs/cm). Non-replicated treatments were put into two groups, low (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) salinity levels (<b>a</b>,<b>b</b>).</p>
Full article ">Figure A3 Cont.
<p>(<b>a</b>,<b>b</b>) Cumulative leaf loss by species and (<b>c</b>) proportion of <span class="html-italic">Acer rubrum</span> litter fall over the course of this experiment compared to (<b>d</b>) groundwater-specific conductance (µs/cm). Non-replicated treatments were put into two groups, low (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) salinity levels (<b>a</b>,<b>b</b>).</p>
Full article ">Figure A4
<p>Comparison of (<b>a</b>) specific leaf area (SLA, leaf area cm<sup>3</sup>/g dry mass) and (<b>b</b>) leaf dry matter content (LDMC, mg dry leaf mass/g water-saturated fresh leaf mass) within and across all six species and binned salinity treatments (control, low, mid, and high). Non-replicated treatments were put into two groups, low (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) salinity levels. Treatments statistically significant (<span class="html-italic">p</span> ≤ 0.05) compared to controls are indicated with an asterisk above the boxplot. Black asterisks below the boxplot indicator significance to other salinity treatment within the bracket. Letters below each set of boxplots per species indicate the statistical significance across species controls. Black points are outliers (values &gt; 1.5 times the interquartile range) included in the statistical analysis.</p>
Full article ">Figure A5
<p>(<b>a</b>) The relationship between osmotic potential (Ψ<sub>π</sub>) of the groundwater treatment solutions (NaCl) after the second salinity dose and the difference between Ψ<sub>π</sub> and predawn leaf water potentials (Ψ<sub>pd</sub>) across the hardwood species and salinity treatments. Each color and shape combination represent Ψ<sub>π</sub> and Ψ<sub>pd</sub> for each species. The horizontal arrow (color gradient from green to red) represents the direction of salinity increase as Ψ<sub>π</sub> becomes more negative. The vertical arrow in panel (<b>a</b>) shows that as Ψ<sub>pd</sub> decreases in tandem with a decrease in Ψ<sub>π</sub>, hydrological flow between the soil and roots reaches equilibrium and ceases to move resulting in leaf loss and mortality. The points closest to zero on the x-axis are controls and salinity treatments increase as they move to the right (more negative Ψ<sub>π</sub>), noted by the green to red gradient arrow. (<b>b</b>) The relationship between osmotic potential (Ψ<sub>π</sub>) of the groundwater treatment solutions (NaCl) and the difference between midday water potential (Ψ<sub>md</sub>) and Ψ<sub>pd</sub> across the hardwood species and salinity treatments. As Ψ<sub>π</sub> increases (salinity treatments become more saline) hydrological flow between the soil (Ψ<sub>pd</sub>) and light-adapted leaves (Ψ<sub>md</sub>), equilibrium would be expected (Ψ<sub>md</sub> − Ψ<sub>pd</sub> = 0; noted by “No H<sub>2</sub>O Transport”) at high soil salinity, meaning water would ceases to move to more negative pressure potential in the leaves.</p>
Full article ">Figure A6
<p>CO<sub>2</sub> assimilation, i.e., photosynthetic rate (A<sub>net</sub>, μmol CO<sub>2</sub> m<sup>−2</sup> s<sup>−1</sup>), for six tree species from June to October. Significant differences in salinity treatments mean (<span class="html-italic">p</span> ≤ 0.05) from the controls are marked with asterisks, and marginally significant means (<span class="html-italic">p =</span> 0.05–0.1) with hats just above the x-axis at zero, colored by the treatment. Green squares and orange diamonds are the primary replicated salinity treatments (control and 3 ppt, respectively). Non-replicated treatments were put into two groups, low (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) salinity levels. Grey triangles and black circles are the low (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) treatments. Measurements were taken every month ±5 days from the first day of each month. Points of each measurement are offset for ease of visualization, but all measurements for each species were taken in 1–2 days between 10:00–15:00 h.</p>
Full article ">Figure A7
<p>Stomatal conductance (<span class="html-italic">g<sub>s</sub></span>, μmol H<sub>2</sub>O m<sup>−2</sup> s<sup>−1</sup>) for six tree species from June to October. Significant differences in salinity treatments mean (<span class="html-italic">p</span> ≤ 0.05) from the controls are marked with asterisks, and marginally significant means (<span class="html-italic">p =</span> 0.05–0.1) with hats just above the x-axis at zero, colored by the treatment. Green squares and orange diamonds are the primary replicated salinity treatments (control and 3 ppt, respectively). Non-replicated treatments were put into two groups, low (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) salinity levels. Measurements were taken every month ±5 days from the first day of each month. Points of each measurement are offset for ease of visualization, but all measurements for each species were taken in 1–2 days between 10:00–15:00 h.</p>
Full article ">Figure A8
<p>Quantum efficiency of photosystem II (ΦPSII) of light-adapted leaves all six tree species from June–October 2018 using a pulse-amplitude modulated (PAM) fluorometer chamber for measuring leaf chlorophyll fluorescence and gas exchange simultaneously. Marginal significance in each salinity treatment means (<span class="html-italic">p =</span> 0.05–0.1) from the controls are marked with hats just above the x-axis at zero, colored by the treatment significantly affected by salinity. Green squares and orange diamonds are the primary replicated salinity treatments (control and 3 ppt, respectively). Non-replicated treatments were put into two groups, low (0.5, 1, and 2 ppt) and high (4, 5, and 6 ppt) salinity levels. Measurements were taken every month roughly ±5 days from the first day of each month. Points of each measurement are offset for ease of visualization, but all measurements for each species were taken in 1–2 days between 10:00–15:00 h.</p>
Full article ">Figure A9
<p>(<b>a</b>) Linear relationship (R<sup>2</sup> = 0.52, <span class="html-italic">p</span> = &lt;0.001) between stomatal conductance (g<sub>wv</sub>) and photosynthesis rate (A<sub>net</sub>) across all 6 species (<span class="html-italic">Acer rubrum</span>, <span class="html-italic">Juniper virginiana</span>, <span class="html-italic">Nyssa sylvatica</span>, <span class="html-italic">Pinus taeda</span>, <span class="html-italic">Quercus nigra</span>, and <span class="html-italic">Taxodium distichum</span>) and binned by salinity treatments (control, low (0.5–2 ppt), mid (3 ppt), and high (4–6 ppt)). Control trees are indicated by green circles, 0.5–2 ppt salinity by light green hollow circles, 3 ppt salinity by orange triangles, and 4–6 ppt by red hollow triangles. (<b>b</b>) Linear relationship (R<sup>2</sup> = 0.59, <span class="html-italic">p</span> = &lt;0.01) between stomatal conductance (g<sub>wv</sub>) and photosynthesis rate (A<sub>net</sub>) of <span class="html-italic">Acer rubrum</span> in the two replicated salinity treatments: controls (green circles) and 3 ppt (orange triangles).</p>
Full article ">Figure A9 Cont.
<p>(<b>a</b>) Linear relationship (R<sup>2</sup> = 0.52, <span class="html-italic">p</span> = &lt;0.001) between stomatal conductance (g<sub>wv</sub>) and photosynthesis rate (A<sub>net</sub>) across all 6 species (<span class="html-italic">Acer rubrum</span>, <span class="html-italic">Juniper virginiana</span>, <span class="html-italic">Nyssa sylvatica</span>, <span class="html-italic">Pinus taeda</span>, <span class="html-italic">Quercus nigra</span>, and <span class="html-italic">Taxodium distichum</span>) and binned by salinity treatments (control, low (0.5–2 ppt), mid (3 ppt), and high (4–6 ppt)). Control trees are indicated by green circles, 0.5–2 ppt salinity by light green hollow circles, 3 ppt salinity by orange triangles, and 4–6 ppt by red hollow triangles. (<b>b</b>) Linear relationship (R<sup>2</sup> = 0.59, <span class="html-italic">p</span> = &lt;0.01) between stomatal conductance (g<sub>wv</sub>) and photosynthesis rate (A<sub>net</sub>) of <span class="html-italic">Acer rubrum</span> in the two replicated salinity treatments: controls (green circles) and 3 ppt (orange triangles).</p>
Full article ">Figure A10
<p>Operating efficiency of photosystem II (Φ PSII) in light-adapted <span class="html-italic">Acer rubrum</span> leaves as a function of photochemical reflectance index (PRI) (<b>top row</b>) and normalized difference vegetation index (NDVI) (<b>bottom row</b>) in June, September, and October 2018. Each point is the mean of each of the eight salinity treatments, with horizontal and vertical error bars as standard errors (0.95 CI). Non-linear regression lines (dark grey) are significant relationship (October 2018).</p>
Full article ">
25 pages, 5793 KiB  
Article
Prolonged Post-Harvest Preservation in Lettuce (Lactuca sativa L.) by Reducing Water Loss Rate and Chlorophyll Degradation Regulated through Lighting Direction-Induced Morphophysiological Improvements
by Jingli Yang, Jinnan Song, Jie Liu, Xinxiu Dong, Haijun Zhang and Byoung Ryong Jeong
Plants 2024, 13(18), 2564; https://doi.org/10.3390/plants13182564 - 12 Sep 2024
Abstract
To investigate the relationship between the lighting direction-induced morphophysiological traits and post-harvest storage of lettuce, the effects of different lighting directions (top, T; top + side, TS; top + bottom, TB; side + bottom, SB; and top + side + bottom, TSB; the [...] Read more.
To investigate the relationship between the lighting direction-induced morphophysiological traits and post-harvest storage of lettuce, the effects of different lighting directions (top, T; top + side, TS; top + bottom, TB; side + bottom, SB; and top + side + bottom, TSB; the light from different directions for a sum of light intensity of 600 μmol·m−2·s−1 photosynthetic photon flux density (PPFD)) on the growth morphology, root development, leaf thickness, stomatal density, chlorophyll concentration, photosynthesis, and chlorophyll fluorescence, as well as the content of nutrition such as carbohydrates and soluble proteins in lettuce were analyzed. Subsequently, the changes in water loss rate, membrane permeability (measured as relative conductivity and malondialdehyde (MDA) content), brittleness (assessed by both brittleness index and β-galactosidase (β-GAL) activity), and yellowing degree (evaluated based on chlorophyll content, and activities of chlorophyllase (CLH) and pheophytinase (PPH)) were investigated during the storage after harvest. The findings indicate that the TS treatment can effectively reduce shoot height, increase crown width, enhance leaves’ length, width, number, and thickness, and improve chlorophyll fluorescence characteristics, photosynthetic capacity, and nutrient content in lettuce before harvest. Specifically, lettuce’s leaf thickness and stomatal density showed a significant increase. Reasonable regulation of water loss in post-harvested lettuce is essential for delaying chlorophyll degradation. It was utilized to mitigate the increase in conductivity and hinder the accumulation of MDA in lettuce. The softening speed of leafy vegetables was delayed by effectively regulating the activity of the β-GAL. Chlorophyll degradation was alleviated by affecting CLH and PPH activities. This provides a theoretical basis for investigating the relationship between creating a favorable light environment and enhancing the post-harvest preservation of leafy vegetables, thus prolonging their post-harvest storage period through optimization of their morphophysiological phenotypes. Full article
(This article belongs to the Section Crop Physiology and Crop Production)
Show Figures

Figure 1

Figure 1
<p>The morphology of lettuce (<span class="html-italic">Lactuca sativa</span> L.) ‘Caesar Green’ plants under various lighting directions for 45 days. The bar indicates 5 cm.</p>
Full article ">Figure 2
<p>The thickness of lettuce leaves under various lighting directions during 45 days of cultivation. (<b>A</b>) The micrograph of the cross-section of the middle section of the mature fourth leaf from the top; the bars indicate 0.2 mm. (<b>B</b>) The leaf thickness at the same site under different treatments. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 3
<p>The growth and development of lettuce roots under various lighting directions during 45 days of cultivation. (<b>A</b>) The raw state of the lettuce root after removing the pot. (<b>B</b>) The state of the cleaned lettuce roots; the bar indicates 5 cm. (<b>C</b>) Root length (average most extended root length). (<b>D</b>) Root fresh weight. (<b>E</b>) Root dry weight. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 4
<p>The trait of lettuce leaf epidermal cells under various lighting directions during 45 days of cultivation. Micrographs of epidermal cells (20×). (<b>A</b>) The micrograph of the upper and lower epidermal cells in the mature fourth leaf from the top; the bars indicate 10 μm. (<b>B</b>) The average length and width of upper and lower epidermal cells. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 5
<p>The trait of stomatal cells in the lower surface of lettuce leaves under various lighting directions during 45 days of cultivation. Micrographs of stomatal density and morphology (20×). (<b>A</b>) The micrograph of stomatal cells in the lower surface of the mature fourth leaf from the top; the bars indicate 10 μm. (<b>B</b>) The stomatal density per 100 mm<sup>2</sup>. (<b>C</b>) The length and width of guard cell pairs. (<b>D</b>) The length and width of stomatal pores. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 6
<p>The photosynthetic and chlorophyll fluorescence characteristics of lettuce plants under various lighting directions during 45 days of cultivation. (<b>A</b>) Net photosynthetic rate. (<b>B</b>) Transpiration rate. (<b>C</b>) Stomatal conductance. (<b>D</b>) Intercellular CO<sub>2</sub> concentration. (<b>E</b>) The maximal PSII quantum yield. (<b>F</b>) The photochemical efficiency of PSII. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 7
<p>The content of carbohydrates and soluble proteins of lettuce plants under various lighting directions during 45 days of cultivation. (<b>A</b>) Total soluble sugar content. (<b>B</b>) Starch content. (<b>C</b>) Soluble protein content. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9). Different lowercase letters indicate significant differences within treatments by Duncan’s multiple range test at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 8
<p>Cultured under various lighting directions for 45 days, after picking, the (<b>A</b>) chlorophyll content, (<b>B</b>) chlorophyllase (CLH), and (<b>C</b>) pheophytinase (PPH) contents in lettuce leaves during storage at room temperature (22 °C, 60% RH (Relative Humidity)) for 0, 3, 6, 9, and 12 days, respectively. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9).</p>
Full article ">Figure 9
<p>Cultured under various lighting directions for 45 days, after picking, the water loss rate in lettuce leaves during storage at room temperature (22 °C, 60% RH (Relative Humidity)) for 3, 6, 9, and 12 days, respectively. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9).</p>
Full article ">Figure 10
<p>Cultured under various lighting directions for 45 days, after picking, the (<b>A</b>) relative conductivity and (<b>B</b>) malondialdehyde (MDA) content in lettuce leaves during storage at room temperature (22 °C, 60% RH (Relative Humidity)) for 3, 6, 9, and 12 days, respectively. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9).</p>
Full article ">Figure 11
<p>Cultured under various lighting directions for 45 days, after picking, the (<b>A</b>) β-Galactosidase (β-GAL) activity and (<b>B</b>) brittleness in lettuce leaves during storage at room temperature (22 °C, 60% RH (Relative Humidity)) for 0, 3, 6, 9, and 12 days, respectively. The above parameters of the mature leaves were measured by selecting from the top to the fourth round. Vertical bars represent means ± standard error (n = 9).</p>
Full article ">Figure 12
<p>The experimental layout and design of the lighting direction combinations. (<b>A</b>) One of the plant culture shelves in the closed-type plant factory. The T, TS, TB, SB, and TSB refer to the top (1/1), top (1/2) + side (1/2), top (1/2) + bottom (1/2), side (1/2) + bottom (1/2), and top (1/3) + side (1/3) + bottom (1/3) lighting, respectively; please see the detailed information in <a href="#plants-13-02564-t002" class="html-table">Table 2</a>. (<b>B</b>) The experimental light treatments utilized white LEDs with a spectral distribution of ~400–750 nm, peaking at 452 nm. The light period followed a 12 h day/night cycle starting at 8:00 a.m. daily. (<b>C</b>) Shading treatments between layers treated with different light directions.</p>
Full article ">Figure 13
<p>Freehand slice cutting range (the middle section of the mature fourth leaf from the top of the treated plants).</p>
Full article ">
16 pages, 2087 KiB  
Article
Dynamics of Mulatto Grass Regrowth Depending on Rotational Cattle Grazing Management
by Carlindo Santos Rodrigues, Márcia Cristina Teixeira da Silveira, Leandro Martins Barbero, Salim Jacaúna Sousa Júnior, Veridiana Aparecida Limão, Guilherme Pontes Silva, Sila Carneiro da Silva and Domicio do Nascimento Júnior
Grasses 2024, 3(3), 174-189; https://doi.org/10.3390/grasses3030013 - 3 Sep 2024
Viewed by 275
Abstract
This study was carried out to characterize the dynamics of forage accumulation during the regrowth of Mulatto grass submitted to rotational grazing strategies. The treatments corresponded to combinations between two pre-grazing conditions (95% and a maximum light interception during regrowth—LI95% and LI [...] Read more.
This study was carried out to characterize the dynamics of forage accumulation during the regrowth of Mulatto grass submitted to rotational grazing strategies. The treatments corresponded to combinations between two pre-grazing conditions (95% and a maximum light interception during regrowth—LI95% and LIMax) and two post-grazing conditions (post-grazing heights of 15 and 20 cm), according to a 2 × 2 factorial arrangement and randomized complete block design, with four replications. Rates of leaf growth (LGR), stems growth (SGR), total growth (TGR), leaf senescence (LSR), grass accumulation (GAR) (kg·ha−1·day−1), and the senescence/canopy growth ratio during different stages of regrowth. There was no difference between the management strategies for TGR. However, a higher GAR was reported for pastures managed with LI95% relative to LIMax, of 161.7 and 120.2 kg DM ha−1·day−1, respectively. Pastures managed with LI95% have a lower SGR in the intermediate and final regrowth period, reflecting the efficient control in the stalks production. On the other hand, in pastures managed, the LIMax showed higher SGR and LSR in the final regrowth phase. Thus, the LAI was higher in pastures managed at LI95% compared to those managed at LIMax, of 163.9 and 112.7 kg DM ha−1·day−1, respectively. Mulatto grass pastures, which were managed at LI95% pre-grazing, corresponded to approximately 30 cm in height, showed higher LAI, and ensured a low SGR throughout the regrowth period, constituting a more efficient management strategy. Full article
Show Figures

Figure 1

Figure 1
<p>Monthly averages of precipitation, maximum temperature, average temperature, and minimum temperature, January to April 2009, in the experimental area.</p>
Full article ">Figure 2
<p>Monthly water balance extract in the experimental area from January to April 2009.</p>
Full article ">Figure 3
<p>Relationship between light interception and leaf area index during regrowth depending on rotational cattle grazing management.</p>
Full article ">Figure 4
<p>Pre-grazing height during regrowth of Mulatto grass depending on rotational cattle grazing management characterized by pre-grazing targets LI<sub>95%</sub> and LI<sub>Max</sub> from January to April 2009. Averages followed by the same letter do not differ from each other (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 5
<p>Total growth rate (TGR) during regrowth of Mulatto depending on rotational cattle grazing management. Averages followed by the same letter do not differ from each other (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 6
<p>Leaf growth rate (LGR) of Mulatto grass managed with pre-grazing targets LI<sub>95%</sub> and LI<sub>Max</sub> (<b>A</b>) during the regrowth period, (<b>B</b>) depending on rotational cattle grazing management. Averages followed by the same letter do not differ from each other (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 7
<p>Mulatto grass accumulation rate (GAR) managed with pre-grazing targets LI<sub>95%</sub> and LI<sub>Max</sub> (<b>A</b>) during the regrowth period (<b>B</b>) depending on rotational cattle grazing management. Averages followed by the same letter did not differ (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">
21 pages, 4155 KiB  
Article
Regulating Leaf Photosynthesis and Soil Microorganisms through Controlled-Release Nitrogen Fertilizer Can Effectively Alleviate the Stress of Elevated Ambient Ozone on Winter Wheat
by Nanyan Zhu, Yinsen Qian, Lingqi Song, Qiaoqiao Yu, Haijun Sheng, Ying Li and Xinkai Zhu
Int. J. Mol. Sci. 2024, 25(17), 9381; https://doi.org/10.3390/ijms25179381 - 29 Aug 2024
Viewed by 210
Abstract
The mitigation mechanisms of a kind of controlled-release nitrogen fertilizer (sulfur-coated controlled-release nitrogen fertilizer, SCNF) in response to O3 stress on a winter wheat (Triticum aestivum L.) variety (Nongmai-88) were studied in crop physiology and soil biology through the ozone-free-air controlled [...] Read more.
The mitigation mechanisms of a kind of controlled-release nitrogen fertilizer (sulfur-coated controlled-release nitrogen fertilizer, SCNF) in response to O3 stress on a winter wheat (Triticum aestivum L.) variety (Nongmai-88) were studied in crop physiology and soil biology through the ozone-free-air controlled enrichment (O3-FACE) simulation platform and soil microbial metagenomics. The results showed that SCNF could not delay the O3-induced leaf senescence of winter wheat but could enhance the leaf size and photosynthetic function of flag leaves, increase the accumulation of nutrient elements, and lay the foundation for yield by regulating the release rate of nitrogen (N). By regulating the soil environment, SCNF could maintain the diversity and stability of soil bacterial and archaeal communities, but there was no obvious interaction with the soil fungal community. By alleviating the inhibition effects of O3 on N-cycling-related genes (ko00910) of soil microorganisms, SCNF improved the activities of related enzymes and might have great potential in improving soil N retention. The results demonstrated the ability of SCNF to improve leaf photosynthetic function and increase crop yield under O3-polluted conditions in the farmland ecosystem, which may become an effective nitrogen fertilizer management measure to cope with the elevated ambient O3 and achieve sustainable production. Full article
(This article belongs to the Special Issue Genetic Engineering of Plants for Stress Tolerance)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>–<b>d</b>) Photosynthetic functions of wheat flag leaves in the flowering period (Z60), (<b>e</b>–<b>g</b>) nutrient accumulation of functional leaves, and (<b>h</b>–<b>k</b>) yield components at the maturity stage (Z92). A_CK, normal atmospheric environment + urea; A_S, normal atmospheric environment + sulfur-coated controlled release nitrogen fertilizer (SCNF); E_CK, ozone-free-air controlled enrichment (O<sub>3</sub>-FACE) + urea; E_S, O<sub>3</sub>-FACE + SCNF. Error bars mean the standard error and the different lowercase letters indicate significant differences between various treatments based on a one-way ANOVA followed by Duncan’s multiple-range tests (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Soil chemical properties at the (<b>a</b>) flowering (Z60) and (<b>b</b>) mature period in different treatments and (<b>c</b>) activities of soil urease and soil nitrate reductase at the flowering period (Z60). OM, organic matter; NO<sub>3</sub><sup>−</sup>–N, nitrate nitrogen; NH<sub>4</sub><sup>+</sup>–N, ammonium nitrogen; AK, available potassium; AP, available phosphorus. Error bars mean the standard error and the different lowercase letters indicate significant differences between various treatments based on a one-way ANOVA followed by Duncan’s multiple-range tests (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Influence of O<sub>3</sub>-FACE combined with SCNF on the relative abundance of soil (<b>a</b>) bacterial; (<b>b</b>) fungal; and (<b>c</b>) archaeal phyla. Only the phyla with RPKM ≥ 1% are presented in this figure.</p>
Full article ">Figure 4
<p>The principal component analysis (PCoA) and PERMANOVA at 99% level based on Bray-Curtis distance of soil (<b>a</b>) bacterial; (<b>b</b>) fungal; and (<b>c</b>) archaeal communities at the species level in the various treatments.</p>
Full article ">Figure 5
<p>Redundancy analysis (RDA) of soil (<b>a</b>) bacterial; (<b>b</b>) fungal; and (<b>c</b>) archaeal genera with soil chemical properties. The soil biochemical properties were fitted to the ordination plots using a 999–permutation test (P_value). OM, soil organic matter; NO<sub>3</sub><sup>−</sup>–N, nitrate nitrogen; NH<sub>4</sub><sup>+</sup>–N, ammonium nitrogen; AP, available phosphorus; AK, available potassium; SUE, soil urease; SNR, soil nitrate reductase. Asterisks indicate significant differences at * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Correlation network between soil (<b>a</b>) bacterial; (<b>b</b>) fungal; and (<b>c</b>) archaeal genera and soil properties based on Gephi 0.9.2 software. Each network node represents a genus; its color and size correspond to the phylum to which it belongs and the relative abundance, respectively. The color and thickness of the network edge are expressed in the Spearman correlation and <span class="html-italic">r</span> value between the genus and the environmental factor, respectively.</p>
Full article ">Figure 7
<p>Pairwise comparisons of soil variables, agronomic parameters, and wheat yield. LS, leaf size; LMA, leaf mass per area; <span class="html-italic">Pn</span>, net photosynthetic rate. Mantel tests depict the association between soil microbial phyla and metabolic functions at the KEGG level 2 with environmental factors, respectively. The width of each edge matches Mantel’s <span class="html-italic">r</span> statistic, and the color represents Mantel’s P value (<a href="#app1-ijms-25-09381" class="html-app">Table S2</a>). Asterisks indicate significant differences at * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Variance analysis of 14 key functional genes involved in N-cycling. DNRA, dissimilatory nitrate reduction ammonia; ANRA, assimilatory nitrate reduction ammonia. Different lowercase letters mean significant difference based on a one-way ANOVA followed by Duncan’s multiple-range tests (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>(<b>a</b>) The ozone-free-air controlled enrichment (O<sub>3</sub>-FACE) simulation platform. (<b>b</b>) AOT40 value during the treatment phase of elevated ozone. (<b>c</b>) The average temperature and total precipitation at the experimental field during the wheat growing season in 2022–2023.</p>
Full article ">
18 pages, 3728 KiB  
Article
Jackfruit Genotypes in Southern Nayarit: A Comparative Study of Morphological, Physiological, Physicochemical, Phytochemical, and Molecular Assessments
by David Antonio Morelos-Flores, Efigenia Montalvo-González, Martina Alejandra Chacon-López, Amalio Santacruz-Varela, Víctor Manuel Zamora-Gasga, Guillermo Berumen-Varela and María de Lourdes García-Magaña
Horticulturae 2024, 10(9), 918; https://doi.org/10.3390/horticulturae10090918 - 28 Aug 2024
Viewed by 354
Abstract
Jackfruit, primarily cultivated in Nayarit, Mexico, has four notable genotypes: “Agüitada”, “Rumina”, “Licenciada”, and “Karlita”, which require thorough characterization. This study aimed to provide a comprehensive characterization of these genotypes through an integration of morphological, physiological, physicochemical, phytochemical, and DNA fingerprinting analyses. Measurements [...] Read more.
Jackfruit, primarily cultivated in Nayarit, Mexico, has four notable genotypes: “Agüitada”, “Rumina”, “Licenciada”, and “Karlita”, which require thorough characterization. This study aimed to provide a comprehensive characterization of these genotypes through an integration of morphological, physiological, physicochemical, phytochemical, and DNA fingerprinting analyses. Measurements were taken from physiological maturity to senescence. SSR and SRAP markers were employed for DNA fingerprinting, and a complete randomized design followed by multivariate analysis was used to observe variable relationships. The results revealed that “Rumina” had the largest leaf size, while “Karlita” had the largest fruit size and the highest respiration rate (117.27 mL of CO2·kg−1·h−1). “Licenciada” showed the highest ethylene production (265.45 µL·kg−1·h−1). “Agüitada” and “Licenciada” were associated with orange bulbs, whereas “Rumina” and “Karlita” were associated with yellow ones. Additionally, “Agüitada” demonstrated higher levels of soluble phenols and carotenoids, indicating greater antioxidant capacity. The Jaccard index suggested moderate genetic diversity among the genotypes, and the dendrogram revealed two genetic clusters. “Licenciada” emerged as a promising genotype, combining high genetic diversity with desirable physicochemical traits. This study highlights the need to broaden future genetic analyses to include a wider range of jackfruit genotypes from various regions, offering a more comprehensive understanding of genetic diversity. Full article
(This article belongs to the Section Fruit Production Systems)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Jackfruit genotypes of Nayarit (fruits, leaves, and bulbs): “Agüitada” (<b>a</b>), “Rumina” (<b>b</b>), “Licenciada” (<b>c</b>), and “Karlita” (<b>d</b>).</p>
Full article ">Figure 2
<p>Scatter plot of the morphological analysis results in four jackfruit genotypes (n = 50 on leaves, n = 15 on fruits): leaf vertical length (<b>a</b>), leaf horizontal length (<b>b</b>), polar diameter of the fruit (<b>c</b>), and equatorial diameter of the fruit (<b>d</b>). A horizontal line is used to indicate the mean value between each data spread. Different letters indicate statistically significant differences (Fisher LSD test; <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 3
<p>Physiological analysis of four jackfruit genotypes (n = 11): respiration rate (<b>a</b>), ethylene production rate (<b>b</b>), and physiological weight loss (<b>c</b>). Columns indicate means of each genotype per analysis; different letters denote statistically significant differences (Fisher LSD test; <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 4
<p>Radar chart of physicochemical analysis of four jackfruit genotypes (n = 12): peel color (<b>a</b>), bulb color (<b>b</b>), peel firmness (<b>c</b>), bulb firmness (<b>d</b>), titratable acidity (<b>e</b>), pH (<b>f</b>), and total soluble solids (<b>g</b>). Different letters indicate statistically significant differences between genotypes (Fisher LSD test; <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 5
<p>Boxplot of the phytochemical analysis in four jackfruit genotypes (n = 12): total soluble phenols (<b>a</b>), total carotenoids (<b>b</b>), DPPH total soluble phenols (<b>c</b>), DPPH total carotenoids (<b>d</b>), FRAP total soluble phenols (<b>e</b>), FRAP total carotenoids (<b>f</b>), ABTS total soluble phenols (<b>g</b>), and ABTS total carotenoids (<b>h</b>). The horizontal line denotes the mean value of the assessed variable. The values within the box correspond to the 50% of measurements nearest to the mean. The whiskers depict the measurements furthest from the mean, indicating dispersion. Different letters denote statistically significant differences between genotypes (Fisher LSD test; <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 6
<p>Principal component analysis of morphological, physiological, physicochemical, and phytochemical analysis in jackfruit genotypes.</p>
Full article ">Figure 7
<p>Dendrogram illustrating the genetic distances among the analyzed genotypes of jackfruit.</p>
Full article ">
19 pages, 5162 KiB  
Article
The AtMINPP Gene, Encoding a Multiple Inositol Polyphosphate Phosphatase, Coordinates a Novel Crosstalk between Phytic Acid Metabolism and Ethylene Signal Transduction in Leaf Senescence
by Xiaoyun Peng, Haiou Li, Wenzhong Xu, Qian Yang, Dongming Li, Tingting Fan, Bin Li, Junhui Ding, Wenzhen Ku, Danyi Deng, Feiying Zhu, Langtao Xiao and Ruozhong Wang
Int. J. Mol. Sci. 2024, 25(16), 8969; https://doi.org/10.3390/ijms25168969 - 17 Aug 2024
Viewed by 507
Abstract
Plant senescence is a highly coordinated process that is intricately regulated by numerous endogenous and environmental signals. The involvement of phytic acid in various cell signaling and plant processes has been recognized, but the specific roles of phytic acid metabolism in Arabidopsis leaf [...] Read more.
Plant senescence is a highly coordinated process that is intricately regulated by numerous endogenous and environmental signals. The involvement of phytic acid in various cell signaling and plant processes has been recognized, but the specific roles of phytic acid metabolism in Arabidopsis leaf senescence remain unclear. Here, we demonstrate that in Arabidopsis thaliana the multiple inositol phosphate phosphatase (AtMINPP) gene, encoding an enzyme with phytase activity, plays a crucial role in regulating leaf senescence by coordinating the ethylene signal transduction pathway. Through overexpressing AtMINPP (AtMINPP–OE), we observed early leaf senescence and reduced chlorophyll contents. Conversely, a loss-of-function heterozygous mutant (atminpp/+) exhibited the opposite phenotype. Correspondingly, the expression of senescence-associated genes (SAGs) was significantly upregulated in AtMINPP–OE but markedly decreased in atminpp/+. Yeast one-hybrid and chromatin immunoprecipitation assays indicated that the EIN3 transcription factor directly binds to the promoter of AtMINPP. Genetic analysis further revealed that AtMINPP–OE could accelerate the senescence of ein3–1eil1–3 mutants. These findings elucidate the mechanism by which AtMINPP regulates ethylene-induced leaf senescence in Arabidopsis, providing insights into the genetic manipulation of leaf senescence and plant growth. Full article
(This article belongs to the Special Issue Transcription Factors in Plant Gene Expression Regulation)
Show Figures

Figure 1

Figure 1
<p>Identification of <span class="html-italic">AtMINPP</span> in <span class="html-italic">Arabidopsis thaliana</span>. (<b>A</b>) Phylogenetic relationship between <span class="html-italic">Arabidopsis thaliana</span> multiple inositol phosphate phosphatases (MINPPs) and other known histidine acid phosphatase (HAP), acid phosphatase (AP), and phosphatases from plants, microbes, animals, and humans. (<b>B</b>) Multiple alignment of amino acid sequences between AtMINPP and other multiple inositol phosphate phosphatases, HAP phytases, and acid phosphatases around the RHGXRXP, HAE, and HD motifs of the active site. Yellow highlight indicates the same amino acid sequence among the twelve species, red font indicates the amino acid sequence of AtMINPP. (<b>C</b>,<b>D</b>) Acid phosphatase activity (<b>C</b>) and phytase activity assay (<b>D</b>) of the recombinant protein. MBP protein as control. Error bars represent the means ± SD (n = 3). The different lowercase letters above the columns indicate significant differences according to Student’s <span class="html-italic">t</span>-test (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>Expression patterns of the <span class="html-italic">AtMINPP</span> gene. (<b>A</b>–<b>F</b>) Tissue-specific expression patterns of <span class="html-italic">AtMINPP::GUS.</span> (<b>A</b>) A 3-day-old seedling; (<b>B</b>) 7-day-old seedling; (<b>C</b>) 14-day-old seedling; (<b>D</b>) 22-day-old mature plant; (<b>E</b>) flower buds, stems, and cauline leaves; (<b>F</b>) siliques. (<b>G</b>) Expression levels of the <span class="html-italic">AtMINPP</span> gene in different tissues were detected by RT-qPCR. Error bars represent the means ± SD (n = 3). The different lowercase letters above the columns indicate significant differences according to one-way ANOVA (<span class="html-italic">p</span> &lt; 0.05). <span class="html-italic">TUB2</span> was used as an internal reference. (<b>H</b>) The subcellular localization of AtMINPP in Arabidopsis protoplasts.</p>
Full article ">Figure 3
<p>Phytic acid metabolism regulated by AtMINPP is involved in leaf senescence. (<b>A</b>) The leaf senescence phenotypes of 5-week-old plants. (<b>B</b>) The leaf senescence phenotypes of detached rosette leaves from 6-week-old plants. (<b>C</b>) Relative chlorophyll content in fifth to eighth leaves of the three genotypes at the indicated ages; those in 4-week-old plants were designated as 100%. (<b>D</b>,<b>E</b>) Analysis of ion leakage (<b>D</b>) and Fv/Fm values (<b>E</b>) from detached fifth to eighth leaves of 5-week-old plants. (<b>F</b>,<b>G</b>) Analysis of the phytic content from detached fifth to eighth leaves of 5-week-old plants (<b>F</b>) and those leaves at three different developmental stages (<b>G</b>). NS, nonsenescent leaves; ES, early senescent leaves; LS, late senescent leaves. Error bars in C-G represent the means ± SD (n = 3). Different lowercase letters above the columns indicate significant differences according to one-way ANOVA (<b>C</b>–<b>E</b>): <span class="html-italic">p</span> &lt; 0.05; (<b>F</b>,<b>G</b>), <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>AtMINPP plays a key role in the regulation of SAGs. (<b>A</b>) Expression levels of SAGs in the fifth to eighth leaves of 5-week-old Col–0, <span class="html-italic">AtMINPP–OE</span>, and <span class="html-italic">atminpp/+</span> plants detected by RT-qPCR. (<b>B</b>,<b>C</b>) Expression levels of <span class="html-italic">AtMINPP</span> in different parts (<b>B</b>) and different developmental stages (<b>C</b>) of the sixth leaves of Col–0. <span class="html-italic">SAG13</span> was used as a positive control. B, base part; M, middle part; T, tip part. NS, nonsenescent leaves; ES, early senescent leaves; LS, late senescent leaves. Error bars represent the means ± SD (n = 3). Different lowercase letters above the columns indicate significant differences according to one-way ANOVA (<span class="html-italic">p</span> &lt; 0.05). <span class="html-italic">TUB2</span> was used as an internal reference in (<b>A</b>–<b>C</b>).</p>
Full article ">Figure 5
<p>AtMINPP participates in ethylene-induced leaf senescence. (<b>A</b>,<b>G</b>) The senescent phenotypes of detached leaves (fifth to eighth) of Col–0, <span class="html-italic">AtMINPP–OE</span>, and <span class="html-italic">atminpp/+</span> treated with water (CK), 50 μM AgNO<sub>3</sub>, 50 μM AVG, and 100 μM ACC under dark conditions. (<b>B</b>,<b>H</b>) Chlorophyll content of the three genotype leaves with indicated treatment (FW, fresh weight). (<b>C</b>,<b>I</b>) Chlorophyll content ratio of the three genotype leaves with indicated treatment. AgNO<sub>3</sub>/CK, AVG/CK, and ACC/CK indicate the relative ratio between leaves treated with AgNO<sub>3</sub>, AVG, or ACC and the controls for each genotype, respectively. (<b>D</b>,<b>J</b>) Ion leakage of the three genotype leaves with indicated treatment. (<b>E</b>,<b>K</b>) Fv/Fm values of the three genotype leaves with indicated treatment. (<b>F</b>) Expression level of <span class="html-italic">AtMINPP</span> in the 2-week-old Col–0 seedlings treated with ACC for indicated time. Error bars represent the means ± SD (n = 3). Different lowercase letters above the columns indicate significant differences according to two-way ANOVA (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>EIN3 is associated with the <span class="html-italic">AtMINPP</span> promoter and transcriptionally activates its expression. (<b>A</b>) Effector and reporter constructs used in the yeast one-hybrid assay. (<b>B</b>) Yeast one-hybrid assay detected the interaction of the <span class="html-italic">AtMINPP</span> promoter with EIN3. (<b>C</b>) The promoter structure of <span class="html-italic">AtMINPP.</span> Primers used for ChIP-PCR were specific to the promoter regions containing EIN3-binding sites. <span class="html-italic">Locus1</span> to <span class="html-italic">Locus3</span> with short lines indicate fragments for ChIP analysis. (<b>D</b>) The ChIP-quantitative PCR analysis for EIN3. Col–0 and EIN3–MYC represent enrichment abundance from Col–0 and <span class="html-italic">35S::EIN3–MYC</span>. <span class="html-italic">Locus1</span> to <span class="html-italic">Locus3</span> indicated the detected individual fragments of <span class="html-italic">AtMINPP</span> promoters. A sequence with no predicted binding sites of <span class="html-italic">AtMINPP</span> was used as negative control. (<b>E</b>) Transcript levels of <span class="html-italic">AtMINPP</span> between Col–0 and <span class="html-italic">ein3–1eil1–3</span> by RT-qPCR. <span class="html-italic">SAG13</span> was used as a positive control. Error bars represent the means ± SD (n = 3). Different lowercase letters above the columns indicate the significant differences according to two-way ANOVA ((<b>D</b>): <span class="html-italic">p</span> &lt; 0.01) and Student’s <span class="html-italic">t</span>-test ((<b>E</b>): <span class="html-italic">p</span> &lt; 0.05), respectively.</p>
Full article ">Figure 7
<p>EIN3 promotes the expression of <span class="html-italic">AtMINPP</span>. (<b>A</b>) The senescent phenotypes of rosette leaves from 5-week-old plants. (<b>B</b>) Chlorophyll contents in the leaves of 5-week-old plants (FW, fresh weight). (<b>C</b>) Ion leakage rates of the detached leaves of 5-week-old plants. (<b>D</b>) Fv/Fm values in the detached leaves of 5-week-old plants. In (<b>B</b>–<b>D</b>), the fifth to eighth leaves from Col–0, <span class="html-italic">AtMINPP–OE</span>, and <span class="html-italic">atminpp/+</span> were used for each experiment. Error bars in (<b>B</b>–<b>D</b>) represent the means ± SD (n = 3). Different lowercase letters above the columns indicate the significant differences according to one-way ANOVA ((<b>B</b>,<b>C</b>): <span class="html-italic">p</span> &lt; 0.05; (<b>D</b>): <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 8
<p>A working model of the functioning of AtMINPP in ethylene-mediated leaf senescence. Yellow arrowheads represent positive regulation; black blunt arrows represent negative regulation; green dashed lines represent indirect regulation.</p>
Full article ">
16 pages, 8325 KiB  
Article
Transcriptome-Wide Identification of Dark- and Salt-Induced Senescence-Related NAC Gene Family Members in Alfalfa
by Xiangxue Duan, Daicai Tian, Peiran Gao, Yue Sun, Xiaojing Peng, Jiangqi Wen, Hongli Xie, Zeng-Yu Wang and Maofeng Chai
Int. J. Mol. Sci. 2024, 25(16), 8908; https://doi.org/10.3390/ijms25168908 - 15 Aug 2024
Viewed by 592
Abstract
Leaves are a key forage part for livestock, and the aging of leaves affects forage biomass and quality. Preventing or delaying premature leaf senescence leads to an increase in pasture biomass accumulation and an improvement in alfalfa quality. NAC transcription factors have been [...] Read more.
Leaves are a key forage part for livestock, and the aging of leaves affects forage biomass and quality. Preventing or delaying premature leaf senescence leads to an increase in pasture biomass accumulation and an improvement in alfalfa quality. NAC transcription factors have been reported to affect plant growth and abiotic stress responses. In this study, 48 NAC genes potentially associated with leaf senescence were identified in alfalfa under dark or salt stress conditions. A phylogenetic analysis divided MsNACs into six subgroups based on similar gene structure and conserved motif. These MsNACs were unevenly distributed in 26 alfalfa chromosomes. The results of the collinearity analysis show that all of the MsNACs were involved in gene duplication. Some cis-acting elements related to hormones and stress were screened in the 2-kb promoter regions of MsNACs. Nine of the MsNAC genes were subjected to qRT-PCR to quantify their expression and Agrobacterium-mediated transient expression to verify their functions. The results indicate that Ms.gene031485, Ms.gene032313, Ms.gene08494, and Ms.gene77666 might be key NAC genes involved in alfalfa leaf senescence. Our findings extend the understanding of the regulatory function of MsNACs in leaf senescence. Full article
(This article belongs to the Special Issue Plant Response to Abiotic Stress—3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Identification of leaf senescence phenotypes in alfalfa under dark- and salt-stress-induced conditions. Senescence process of detached alfalfa leaves treated with CK (light-control), dark, and 150 mM NaCl for 0, 2, 4, 6, 8, and 10 days. Scale bar, 1 cm.</p>
Full article ">Figure 2
<p>Screening of 48 aging-related <span class="html-italic">MsNACs</span> under dark and 150 mM NaCl treatment. (<b>a</b>) Venn diagram of <span class="html-italic">MsNAC</span> genes under dark (Orange) and 150 mM NaCl treatment (Blue). (<b>b</b>) Heatmap of expression levels for the 48 screened <span class="html-italic">MsNAC</span> genes. The horizontal row represents the gene, while the vertical row represents the treatment period. The far left of the heat map is a dendrogram of <span class="html-italic">MsNACs</span>, and different color bars represent different subclusters. The color bar on the right side of the heat map indicates the gene’s expression level after standardized treatment. Red indicates a high expression level of the gene, while blue represents a low expression level. The gradient from blue to red signifies a change from low to high expression.</p>
Full article ">Figure 3
<p>A phylogenetic analysis of <span class="html-italic">NACs</span> from <span class="html-italic">Medicago sativa</span> and <span class="html-italic">Arabidopsis thaliana</span> (At) was conducted. Subgroups are marked with different colors, and the subgroups’ (I–VI) color is marked in the upper right corner.</p>
Full article ">Figure 4
<p>Phylogenetic relationships and motifs of <span class="html-italic">NAC</span> genes from <span class="html-italic">Medico sativa</span>. (<b>a</b>) Phylogenetic tree of 48 <span class="html-italic">MsNACs</span>. (<b>b</b>) Analysis of conserved elements of 48 <span class="html-italic">MsNACs</span>. Boxes of different colors represent different motifs. (<b>c</b>) Exon–intron organizations of 48 <span class="html-italic">MsNACs</span>. Green boxes indicate exons; black lines indicate introns.</p>
Full article ">Figure 5
<p>Localization analysis of 48 <span class="html-italic">MsNACs</span> on chromosomes. Green bars represent chromosomes; red font represents genes’ ID.</p>
Full article ">Figure 6
<p>Collinear loop diagram of the interior of 48 <span class="html-italic">MsNACs.</span> Red lines indicate duplicated <span class="html-italic">MsNAC</span> gene pairs.</p>
Full article ">Figure 7
<p>Analysis of cis-acting elements in the <span class="html-italic">MsNACs</span> promoter. Different cis-acting elements are represented by circles of different colors.</p>
Full article ">Figure 8
<p>Selected <span class="html-italic">MsNACs</span> based on response to leaf senescence on days 4 and 6. (<b>a</b>) Heatmap of the 17 <span class="html-italic">MsNACs</span> identified on days 4 and 6 of dark- and salt-induced leaf senescence. The horizontal row represents the gene, while the vertical row represents the treatment period. The far left of the heat map is a dendrogram of <span class="html-italic">MsNACs</span>, and different color bars represent different subclusters. The color bar on the right side of the heat map indicates the gene’s expression level after standardized treatment. Red indicates a high expression level of the gene, while blue represents a low expression level. The gradient from blue to red signifies a change from low to high expression. The qRT-PCR verification of Nine <span class="html-italic">MsNACs</span> expression under dark stress (<b>b</b>), salt stress (<b>c</b>) and natural aging (<b>d</b>). Data in (<b>b</b>–<b>d</b>) represent mean values (±SD; n = 3) and were analyzed using Student’s <span class="html-italic">t</span>-test (NS; not significant, * <span class="html-italic">p</span> &lt; 0.05) against D0 or X0.</p>
Full article ">Figure 9
<p>Functional validation of selected <span class="html-italic">MsNACs</span> using an <span class="html-italic">Agrobacterium</span>-mediated transient expression assay. Positive control: <span class="html-italic">SGR</span>. Negative control: empty vector with YFP.</p>
Full article ">
Back to TopTop