[go: up one dir, main page]

 
 
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 (126)

Search Parameters:
Keywords = S. tuberosum

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 8336 KiB  
Article
Morpho-Molecular Identification of Fusarium equiseti and Fusarium oxysporum Associated with Symptomatic Wilting of Potato from Pakistan
by Arsh Bibi, Fathia Mubeen, Ali Rizwan, Irfan Ullah, Masooma Hammad, Muhammad Abu Bakar Waqas, Ayesha Ikram, Zaheer Abbas, Dennis Halterman and Nasir Ahmad Saeed
J. Fungi 2024, 10(10), 701; https://doi.org/10.3390/jof10100701 - 8 Oct 2024
Viewed by 594
Abstract
Potato (Solanum tuberosum L.) is one of the emerging staple crops in Pakistan, with Punjab producing over 95% of the country’s potatoes. Wilt is an emerging threat to the potato crop worldwide, including in Pakistan. We identified and characterized Fusarium species associated [...] Read more.
Potato (Solanum tuberosum L.) is one of the emerging staple crops in Pakistan, with Punjab producing over 95% of the country’s potatoes. Wilt is an emerging threat to the potato crop worldwide, including in Pakistan. We identified and characterized Fusarium species associated with potato wilt in Pakistan through morphological and molecular analyses. Samples were collected during the 2020–2022 potato seasons from five major potato-growing regions: Sahiwal, Chichawatni, Pakpattan, Kamalia, and Faisalabad. Morphological characterization, internal transcribed spacer (ITS) sequencing, specific translation elongation factor 1-alpha (TEF) sequencing, and phylogenetic analysis were used to identify six different Fusarium species: F. oxysporum, F. equiseti, F. incarnatum, F. fujikuroi, F. annulatum and F. thapsinum. Pathogenicity tests in a greenhouse revealed that F. oxysporum and F. equiseti were responsible for Fusarium wilt in all sampled regions, with F. oxysporum being more prevalent in wilted samples. This is the first report of F. equiseti on wilted potatoes in Pakistan. In vitro biocontrol tests using Trichoderma harzianum showed 89% inhibition against F. equiseti and 65% inhibition against F. oxysporum. These findings on F. equiseti will aid in developing future control strategies, including biocontrol measures for Fusarium wilt in potatoes. Full article
(This article belongs to the Special Issue Plant Pathogenic Fungi: Taxonomy, Phylogeny and Morphology)
Show Figures

Figure 1

Figure 1
<p>Front and back images of <span class="html-italic">Fusarium</span> cultures. (<b>A</b>) <span class="html-italic">F. fujikuroi</span>; (<b>B</b>) <span class="html-italic">F. annulatum</span>; (<b>C</b>) <span class="html-italic">F. oxysporum</span>; (<b>D</b>) <span class="html-italic">F. equiseti</span>; (<b>E</b>) <span class="html-italic">F. thapsinum</span>; and (<b>F</b>) <span class="html-italic">F. incarnatum</span>. Images were taken after 14 days of growth.</p>
Full article ">Figure 2
<p>Representative photos of <span class="html-italic">Fusarium</span> mycelia and spores were taken using a binocular compound microscope (Bausch &amp; Lomb Galen). (<b>A</b>) Macroconidia spores of <span class="html-italic">F. equiseti</span>; (<b>B</b>) mycelia of <span class="html-italic">F. equiseti</span>; (<b>C</b>) macroconidia spores of <span class="html-italic">F. equiseti</span>; (<b>D</b>) macroconidia spores of <span class="html-italic">F. oxysporum</span>; (<b>E</b>) macroconidia spores of <span class="html-italic">F. oxysporum</span>; (<b>F</b>) macroconidia spores of <span class="html-italic">F. incarnatum</span>. Scale bars: (<b>A</b>,<b>B</b>) = 100 µm; (<b>C</b>–<b>E</b>) = 25 µm; (<b>F</b>) = 10 µm.</p>
Full article ">Figure 3
<p>(<b>a</b>) Phylogenetic tree of partial <span class="html-italic">TEF</span> gene sequences for the evolutionary relationship of isolates by Maximum Likelihood method. Sequences shown in red are isolates from this study. <span class="html-italic">TEF</span> sequence from <span class="html-italic">Fusarium stilboides</span> was used as an outgroup. (<b>b</b>) Phylogenetic tree of partial <span class="html-italic">ITS</span> gene sequences for the evolutionary relationship of isolates by Maximum Likelihood method. Sequences shown in red are isolates from this study. GenBank accession numbers of <span class="html-italic">ITS</span> sequences are followed by the species name from which the sequences originated. <span class="html-italic">ITS</span> sequence from <span class="html-italic">Fusarium lateritium</span> was used as an outgroup. For both trees, red circles indicate isolation from Pakistan and green circles are database-derived sequences.</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>) Phylogenetic tree of partial <span class="html-italic">TEF</span> gene sequences for the evolutionary relationship of isolates by Maximum Likelihood method. Sequences shown in red are isolates from this study. <span class="html-italic">TEF</span> sequence from <span class="html-italic">Fusarium stilboides</span> was used as an outgroup. (<b>b</b>) Phylogenetic tree of partial <span class="html-italic">ITS</span> gene sequences for the evolutionary relationship of isolates by Maximum Likelihood method. Sequences shown in red are isolates from this study. GenBank accession numbers of <span class="html-italic">ITS</span> sequences are followed by the species name from which the sequences originated. <span class="html-italic">ITS</span> sequence from <span class="html-italic">Fusarium lateritium</span> was used as an outgroup. For both trees, red circles indicate isolation from Pakistan and green circles are database-derived sequences.</p>
Full article ">Figure 4
<p>Symptoms of development of <span class="html-italic">F. oxysporum</span> on potato leaves. Photos were taken of greenhouse-grown plants 7 days after inoculation with <span class="html-italic">F. oxysporum</span>.</p>
Full article ">Figure 5
<p>Series one represents the AGB red variety and series 2 represents the Kuroda variety. Solid bars in both series represent the control plants; the striped bar represents the average tuber mass in grams of plants treated with <span class="html-italic">Fusarium equiseti</span>; and the dotted bar represents the average tuber mass of plants treated with <span class="html-italic">Fusarium oxysporum</span>.</p>
Full article ">Figure 6
<p>Tubers from greenhouse-grown plants inoculated with <span class="html-italic">Fusarium oxysporum</span>. (<b>A</b>) Kuroda; (<b>B</b>) AGB pink; (<b>C</b>) AGB red; (<b>D</b>) AGB CH2.</p>
Full article ">Figure 7
<p>Inhibitory effect of <span class="html-italic">Trichoderma harzianum</span> on <span class="html-italic">Fusarium</span> growth (red circles). The upper row shows control for <span class="html-italic">T. harzianum</span> T.A, T.B and T.C and the left column shows control of <span class="html-italic">Fusarium oxysporum</span> and <span class="html-italic">equiseti</span>. The second row shows <span class="html-italic">Trichoderma harzianum</span> with <span class="html-italic">F. oxysporum</span>, while the lower panel shows <span class="html-italic">T. harzianum</span> with <span class="html-italic">F. equiseti</span>. Control shows the development of fungal pathogens in the absence of <span class="html-italic">T. harzianum</span> and vice versa.</p>
Full article ">
23 pages, 39763 KiB  
Article
Exploring the Molecular Landscape of Nitrogen Use Efficiency in Potato (Solanum tuberosum L.) under Low Nitrogen Stress: A Transcriptomic and Metabolomic Approach
by Rui Xie, Xiaolei Jin, Jing Fang, Shuli Wei, Jie Ma, Ying Liu, Yuchen Cheng, Liyu Chen, Jiawei Liu, Yanan Liu, Zhigang Han, Binyu Guo, Jingshan Guo, Xiaoqing Zhao, Xiangqian Zhang and Zhanyuan Lu
Agronomy 2024, 14(9), 2000; https://doi.org/10.3390/agronomy14092000 - 2 Sep 2024
Viewed by 604
Abstract
Enhancing crop nitrogen use efficiency (NUE) in agricultural sciences is a pivotal challenge, particularly for high-demand crops like potatoes (Solanum tuberosum L.), the world’s third most significant food crop. This study delves into the molecular responses of potatoes to low nitrogen (LN) [...] Read more.
Enhancing crop nitrogen use efficiency (NUE) in agricultural sciences is a pivotal challenge, particularly for high-demand crops like potatoes (Solanum tuberosum L.), the world’s third most significant food crop. This study delves into the molecular responses of potatoes to low nitrogen (LN) stress, employing an integrative approach that combines transcriptomics and metabolomics to compare two cultivars with divergent NUE traits: XS6, known for its high NUE, and NS7, characterized by lower NUE. Our research unveils that XS6 exhibits higher chlorophyll and N content, increased tuber yield, and elevated N assimilation capacity under LN stress conditions compared to NS7. Through transcriptome analysis, we identified critical genes involved in C and N metabolism that had higher expression in XS6. A significant discovery was the high-affinity nitrate transporter 2.7 gene, which showed elevated expression in XS6, suggesting its key role in enhancing NUE. Metabolomics analysis further complemented these findings, revealing a sophisticated alteration of 1252 metabolites under LN stress, highlighting the dynamic interplay between carbon and N metabolism in coping with N scarcity. The integration of transcriptomic and metabolomic data underscored the crucial role of trehalose in mitigating N deficiency and enhancing NUE. This study provides novel insights into the molecular mechanisms governing NUE in potatoes, offering valuable perspectives for molecular breeding to enhance NUE in potatoes and potentially other crops. Full article
(This article belongs to the Section Plant-Crop Biology and Biochemistry)
Show Figures

Figure 1

Figure 1
<p>The agronomic characteristics of two potato varieties, XS6 and NS7, with NUE, evaluated at 45 DAP and at the ripening stages. Measurements of chlorophyll a/b (<b>A</b>,<b>B</b>), total chlorophyll (<b>C</b>), and N content (<b>D</b>) were taken at the seedling stage, 45 days post-sowing. Tuber yield (<b>E</b>) and NUE (<b>F</b>) were assessed at harvest. The data, presented as means ± standard error (SE) for n = 3, underwent statistical analysis via one-way ANOVA, supplemented by Tukey’s honestly significant difference (HSD) post hoc test (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; ns = not significant).</p>
Full article ">Figure 2
<p>Activity levels of key N assimilating enzymes in the leaves and roots of XS6 and NS7 under different N treatments. Enzyme activities for NR (<b>A</b>,<b>E</b>), GS (<b>B</b>,<b>F</b>), GOGAT (<b>C</b>,<b>G</b>), and GDH (<b>D</b>,<b>H</b>) were statistically analyzed using one-way ANOVA followed by Tukey’s HSD post hoc test (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; ns = not significant).</p>
Full article ">Figure 3
<p>Transcriptome analysis of XS6 and NS7 under LN stress. PCA analysis of leaf (<b>A</b>) and root (<b>B</b>) samples. Correlation analysis of 24 samples (<b>C</b>). The numbers of DEGs in the different comparison groups (<b>D</b>). Venn analysis of DEGs that were upregulated (<b>E</b>) and downregulated (<b>F</b>).</p>
Full article ">Figure 4
<p>Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis of DEGs in potato (<span class="html-italic">Solanum tuberosum</span> L.).</p>
Full article ">Figure 5
<p>Weighted Gene Co-expression Network Analysis (WGCNA) reveals the complex interaction between gene expression modules and physiological responses to N availability in potato cultivars XS6 and NS7. (<b>A</b>) The correlation heatmap between co-expression modules and N metabolism-related enzymes highlights the blue module’s significant association with key N assimilation processes. (<b>B</b>) The dendrogram from hierarchical clustering visualizes 18 distinct co-expression modules, with the blue module standing out for further analysis.</p>
Full article ">Figure 6
<p>Diversity of metabolites identified in potato leaves and roots under LN stress conditions. (<b>A</b>) Pie chart illustrating the distribution of the 1252 metabolites across different classes. (<b>B</b>,<b>C</b>) The PCA results for leaves and roots, respectively, illustrating the variation in metabolite profiles across different varieties and N treatments.</p>
Full article ">Figure 7
<p>Overview of the DAMs identified in XS6 and NS7 under LN stress, including the number of DAMs identified (<b>A</b>), Venn diagrams of DAMs in leaves (<b>B</b>) and roots (<b>C</b>), and heatmaps showcasing the expression patterns of shared DAMs in leaves (<b>D</b>) and roots (<b>E</b>).</p>
Full article ">Figure 8
<p>Venn diagrams showcasing shared and unique KEGG pathways between XS6 and NS7 cultivars. (<b>A</b>) Overlap and divergence of KEGG pathways between XS6_L (LN/NN) and NS7_L (LN/NN) in leaf samples, elucidating the common and cultivar-specific metabolic responses to LN conditions. (<b>B</b>) Similar comparison for root samples (XS6_R and NS7_R), highlighting the distinct pathways each cultivar engages in response to N stress. These diagrams emphasize the varietal differences in metabolic strategy and adaptation to LN stress, underpinning the potential for targeted genetic and metabolic engineering to enhance NUE.</p>
Full article ">Figure 9
<p>Co-expression analysis of starch and sucrose metabolism, glycolysis, TCA cycle, and nitrogen (N) metabolism. The heatmap colored in purple and yellow indicates metabolite accumulation. The heatmap colored in green and red indicates gene expression. SUS: sucrose synthase; HK: hexokinase; SPS: sucrose phosphate synthase; SPP: sucrose phosphatase; TPS: trehalose-phosphate synthase; TP: trehalose-phosphatase; SBE: 1,4-alpha-glucan branching enzyme; AMY: beta-amylase; AGPS: glucose-1-phosphate adenylyltransferase; FBA: fructose 1,6 bisphosphate aldolase; PK: pyruvate kinase; PDH: pyruvate dehydrogenase; CS: citrate synthase; ACO: aconitase; IDH: isocitrate dehydrogenase; FUM: fumarase; NRT: nitrate transporter; NR: nitrate reductase; NiR: nitrite reductase; AMT: <math display="inline"><semantics> <msubsup> <mrow> <mi>NH</mi> </mrow> <mn>4</mn> <mo>+</mo> </msubsup> </semantics></math> transporters; GS: glutamine synthetase; GOGAT: glutamate synthase; and PEP: phosphoenolpyruvate.</p>
Full article ">Figure 10
<p>Transcript levels of eight selected differentially expressed genes (DEGs) in the XS6 and NS7 potato cultivars, showcasing both qRT-PCR (represented by bars) and RNA-seq data (indicated by red lines). (<b>A</b>) LOC102577806; (<b>B</b>) LOC102578808; (<b>C</b>) LOC 102580689; (<b>D</b>) LOC102593189; (<b>E</b>) LOC102596437; (<b>F</b>) LOC102604080; (<b>G</b>) NIR; (<b>H</b>) NR3. The graph clearly compares the results obtained by the two techniques, illustrating their overall agreement. Data from qRT-PCR are expressed as means ± standard error (SE) for three biological replicates.</p>
Full article ">Figure 11
<p>The proposed molecular mechanisms contributing to high NUE in potatoes under varying N conditions. The diagram is partitioned by a black dotted line, symbolizing the potato’s physiological and molecular states under LN (<b>left</b>) and NN (<b>right</b>) conditions. Purple arrows depict nitrate uptake and transport, including potential runoff to groundwater, with the arrow thickness reflecting the volume of nitrate uptake and loss under different N treatments. Black arrows represent the influence of specific factors on NUE, providing a visual summary of the interconnected pathways and gene expression that enhance NUE in potatoes.</p>
Full article ">
13 pages, 1415 KiB  
Article
Biological Control of Streptomyces Species Causing Common Scabs in Potato Tubers in the Yaqui Valley, Mexico
by Amelia C. Montoya-Martínez, Roel Alejandro Chávez-Luzanía, Ana Isabel Olguín-Martínez, Abraham Ruíz-Castrejón, Jesús Daniel Moreno-Cárdenas, Fabiola Esquivel-Chávez, Fannie I. Parra-Cota and Sergio de los Santos-Villalobos
Horticulturae 2024, 10(8), 865; https://doi.org/10.3390/horticulturae10080865 - 15 Aug 2024
Viewed by 740
Abstract
Potatoes (Solanum tuberosum L.) represent an important food in the country’s gastronomy due to their cost, nutritional contribution, and versatility. However, many plant diseases such as the common scab—caused by Streptomyces species—reduce its yield and quality. This study aims to determine Streptomyces [...] Read more.
Potatoes (Solanum tuberosum L.) represent an important food in the country’s gastronomy due to their cost, nutritional contribution, and versatility. However, many plant diseases such as the common scab—caused by Streptomyces species—reduce its yield and quality. This study aims to determine Streptomyces species being the causal agent of common scabs in a commercial potato field in the Yaqui Valley, Mexico, while identifying Bacillus strains as a biological control method to mitigate the impact of this disease under field conditions. Thus, three Streptomyces strains were selected from symptomatic samples, and then they were morphologically and molecularly (through sequencing recA and rpoB genes) identified as Streptomyces caniscabiei. After pathogenicity tests, the three strains were found to be pathogenic to potato tubers. In screening assays to identify biocontrol bacteria, strain TSO2T (Bacillus cabrialesii subsp. tritici) and TE3T_UV25 (Bacillus subtilis) had the best in vitro biocontrol effect against S. caniscabiei. Then, a field experiment (1 ha per treatment), under commercial conditions, was carried out to analyze the effectivity of these biocontrol bacteria to mitigate the common scabs on potato crops. After four months, the inoculation of this bacterial consortium decreased common scab incidence from 31% to 21% and increased the potato yield up to almost 5 tons/ha vs. the un-inoculated treatment. These findings demonstrate the effectiveness of the studied bacterial consortium as a potential biological control strategy to control common scabs of potato caused by Streptomyces caniscabiei, as well as increase the potato yield in the Yaqui Valley, Mexico. Full article
Show Figures

Figure 1

Figure 1
<p>Typical lesions of common scabs on potatoes sampled in commercial fields in the Yaqui Valley, Mexico.</p>
Full article ">Figure 2
<p>Maximum likelihood (ML) phylogenetic tree derived from a two-locus dataset. ML bootstrap support (ML-BS) is based on 5000 pseudoreplicates of the data. The outgroup was rooted in sequences of <span class="html-italic">Nocardiopsis dassonvillei</span> NCTC 10488. The bold highlight is used to identify the strains isolated in this study; accession numbers of reference genome assemblies’ sequences are shown in parenthesis.</p>
Full article ">Figure 3
<p><span class="html-italic">Streptomyces caniscabiei</span> pathogenicity test in potato slices. (<b>a</b>) Un-inoculated control; (<b>b</b>) inoculated with the studied <span class="html-italic">Streptomyces caniscabiei</span>.</p>
Full article ">
19 pages, 2771 KiB  
Article
The Influence of Chitosan Derivatives in Combination with Bacillus subtilis Bacteria on the Development of Systemic Resistance in Potato Plants with Viral Infection and Drought
by Liubov Yarullina, Joanna Kalatskaja, Vyacheslav Tsvetkov, Guzel Burkhanova, Ninel Yalouskaya, Katerina Rybinskaya, Evgenia Zaikina, Ekaterina Cherepanova, Kseniya Hileuskaya and Viktoryia Nikalaichuk
Plants 2024, 13(16), 2210; https://doi.org/10.3390/plants13162210 - 9 Aug 2024
Viewed by 650
Abstract
Viral diseases of potatoes are among the main problems causing deterioration in the quality of tubers and loss of yield. The growth and development of potato plants largely depend on soil moisture. Prevention strategies require comprehensive protection against pathogens and abiotic stresses, including [...] Read more.
Viral diseases of potatoes are among the main problems causing deterioration in the quality of tubers and loss of yield. The growth and development of potato plants largely depend on soil moisture. Prevention strategies require comprehensive protection against pathogens and abiotic stresses, including modeling the beneficial microbiome of agroecosystems combining microorganisms and immunostimulants. Chitosan and its derivatives have great potential for use in agricultural engineering due to their ability to induce plant immune responses. The effect of chitosan conjugate with caffeic acid (ChCA) in combination with Bacillus subtilis 47 on the transcriptional activity of PR protein genes and changes in the proteome of potato plants during potato virus Y (PVY) infection and drought was studied. The mechanisms of increasing the resistance of potato plants to PVY and lack of moisture are associated with the activation of transcription of genes encoding PR proteins: the main protective protein (PR-1), chitinase (PR-3), thaumatin-like protein (PR-5), protease inhibitor (PR-6), peroxidase (PR-9), and ribonuclease (PR-10), as well as qualitative and quantitative changes in the plant proteome. The revealed activation of the expression of marker genes of systemic acquired resistance and induced systemic resistance under the influence of combined treatment with B. subtilis and chitosan conjugate indicate that, in potato plants, the formation of resistance to viral infection in drought conditions proceeds synergistically. By two-dimensional electrophoresis of S. tuberosum leaf proteins followed by MALDI-TOF analysis, 10 proteins were identified, the content and composition of which differed depending on the experiment variant. In infected plants treated with ChCA, the synthesis of proteinaceous RNase P 1 and oxygen-evolving enhancer protein 2 was enhanced in conditions of normal humidity, and 20 kDa chaperonin and TMV resistance protein N-like was enhanced in conditions of lack of moisture. The virus coat proteins were detected, which intensively accumulated in the leaves of plants infected with potato Y-virus. ChCA treatment reduced the content of these proteins in the leaves, and in plants treated with ChCA in combination with Bacillus subtilis, viral proteins were not detected at all, both in conditions of normal humidity and lack of moisture, which suggests the promising use of chitosan derivatives in combination with B. subtilis bacteria in the regulation of plant resistance. Full article
(This article belongs to the Special Issue The Role of Signaling Molecules in Plant Stress Tolerance)
Show Figures

Figure 1

Figure 1
<p>The appearance of a leaf of a healthy (1) and PVY-infected (2) potato plant.</p>
Full article ">Figure 2
<p>Detection of PVY by ELISA in the sap of potato leaves after treatment with ChCA and <span class="html-italic">B. subtilis</span> 47 in normal conditions (1) and water deficiency (2). Different letters denote significantly different values.</p>
Full article ">Figure 3
<p>The influence of ChCA and <span class="html-italic">B. subtilis</span> 47 on the proline content (<b>a</b>) and pyrroline-5-carboxylate synthase transcription level (<b>b</b>) in healthy (1) and PVY-infected (2) potato plants on the 10th day after PVY inoculation. Different letters denote significantly different values.</p>
Full article ">Figure 4
<p>The effect of the ChCA and <span class="html-italic">B. subtilis</span> 47 on the relative number of transcripts of the PR-1 ((<b>a</b>), main protective protein) and PR-3 ((<b>b</b>), chitinase) genes in healthy (1) and PVY-infected (2) plants under normal conditions and under conditions of water deficiency. Different letters denote significantly different values.</p>
Full article ">Figure 5
<p>The effect of the ChCA and <span class="html-italic">B. subtilis</span> 47 on the relative number of transcripts of the PR-5 ((<b>a</b>), thaumatin-like protein) and PR-6 ((<b>b</b>), protease inhibitor) genes in healthy (1) and PVY-infected (2) plants under normal conditions and under conditions of water deficiency. Different letters denote significantly different values.</p>
Full article ">Figure 6
<p>The effect of the ChCA and <span class="html-italic">B. subtilis</span> 47 on the relative number of transcripts of the PR-9 ((<b>a</b>), peroxidase) and PR-10 ((<b>b</b>), ribonuclease) genes in healthy (1) and PVY-infected (2) plants under normal conditions and under conditions of water deficiency. Different letters denote significantly different values.</p>
Full article ">Figure 7
<p>The effect of the ChCA and <span class="html-italic">B. subtilis</span> 47 on the relative number of transcripts of the StMT (methyltransferase) gene in healthy (1) and PVY-infected (2) plants under normal conditions and under conditions of water deficiency. Different letters denote significantly different values.</p>
Full article ">
15 pages, 9466 KiB  
Article
Xylem Sap Proteome Analysis Provides Insight into Root–Shoot Communication in Response to flg22
by Romana Kopecká and Martin Černý
Plants 2024, 13(14), 1983; https://doi.org/10.3390/plants13141983 - 20 Jul 2024
Viewed by 661
Abstract
Xylem sap proteomics provides crucial insights into plant defense and root-to-shoot communication. This study highlights the sensitivity and reproducibility of xylem sap proteome analyses, using a single plant per sample to track over 3000 proteins in two model crop plants, Solanum tuberosum and [...] Read more.
Xylem sap proteomics provides crucial insights into plant defense and root-to-shoot communication. This study highlights the sensitivity and reproducibility of xylem sap proteome analyses, using a single plant per sample to track over 3000 proteins in two model crop plants, Solanum tuberosum and Hordeum vulgare. By analyzing the flg22 response, we identified immune response components not detectable through root or shoot analyses. Notably, we discovered previously unknown elements of the plant immune system, including calcium/calmodulin-dependent kinases and G-type lectin receptor kinases. Despite similarities in the metabolic pathways identified in the xylem sap of both plants, the flg22 response differed significantly: S. tuberosum exhibited 78 differentially abundant proteins, whereas H. vulgare had over 450. However, an evolutionarily conserved overlap in the flg22 response proteins was evident, particularly in the CAZymes and lipid metabolism pathways, where lipid transfer proteins and lipases showed a similar response to flg22. Additionally, many proteins without conserved signal sequences for extracellular targeting were found, such as members of the HSP70 family. Interestingly, the HSP70 response to flg22 was specific to the xylem sap proteome, suggesting a unique regulatory role in the extracellular space similar to that reported in mammalians. Full article
(This article belongs to the Section Plant Protection and Biotic Interactions)
Show Figures

Figure 1

Figure 1
<p>Experimental design. (<b>a</b>) Stems of plants grown in soil were cut, and the xylem sap was collected in three consecutive fractions (15–135 min). Representative images of <span class="html-italic">S. tuberosum</span> (<b>left</b>) and <span class="html-italic">H. vulgare</span> (<b>right</b>) plants used in the experiment; (<b>b</b>) a subset of plants was pre-incubated with mock (water) or flg22 applied by spraying the leaves and pouring the solution under the pot as described in Materials and Methods. After 24 h, plants were cut, and the third fraction of the sap (75–135 min) was collected. Each experiment included at least three independent biological replicates.</p>
Full article ">Figure 2
<p>Proteome analysis of <span class="html-italic">S. tuberosum</span> sap. (<b>a</b>) Principal component analysis (PCA) of quantified proteins in three consecutive fractions (F1–F3). Results are based on three biological replicates. (<b>b</b>) Proportion of the proteome extract formed by proteins predicted to have extracellular localization. Proteins were assigned based on homology to <span class="html-italic">Arabidopsis</span> orthologs in the SUBA database (green; SUBA database [<a href="#B40-plants-13-01983" class="html-bibr">40</a>]) and predictions for <span class="html-italic">S. tuberosum</span> proteins from the cropPAL database (gray; cropPAL database [<a href="#B41-plants-13-01983" class="html-bibr">41</a>]). The plots represent the means and standard deviations of three biological replicates. Different letters indicate significant differences (ANOVA, Tukey’s HSD, <span class="html-italic">p</span> &lt; 0.05). See <a href="#app1-plants-13-01983" class="html-app">Supplementary Table S1</a> for details.</p>
Full article ">Figure 3
<p>Sap proteome composition of <span class="html-italic">Solanum tuberosum</span> (<b>left</b>) and <span class="html-italic">Hordeum vulgare</span> (<b>right</b>). (<b>a</b>,<b>b</b>) Estimated abundances of identified enzymes in the third collected fraction; (<b>c</b>,<b>d</b>) differences in categories representing ≥95% of the total identified enzyme abundances visualized on a heat map. The letters represent significant differences (<span class="html-italic">p</span> &lt; 0.05, ANOVA, Tukey’s HSD), the numbers above the connecting lines represent Pearson’s correlation coefficient, and statistically significant correlations (<span class="html-italic">p</span> &lt; 0.05) are marked with asterisks; (<b>e</b>,<b>f</b>) gene ontology enrichment analyses-based annotations of identified Arabidopsis orthologs. Nodes represent enriched GO pathways, with size indicating the number of proteins associated with the pathway and color intensity reflecting enrichment significance. Pathways connected by lines share ≥20% of their protein components. Analyses were performed using ShinyGO 0.8 [<a href="#B42-plants-13-01983" class="html-bibr">42</a>].</p>
Full article ">Figure 4
<p><span class="html-italic">S. tuberosum</span> sap proteome response to flg22. (<b>a</b>) Volcano plot representation of response to flg22 (flg22 vs. mock). Highlighted differentially abundant proteins (DAPs) represent statistically significant differences at 5% FDR. (<b>b</b>) Visualization of functional categories in the ProteoMap. The size of each category corresponds to the estimated protein abundance. (<b>c</b>) The DiVenn visualization depicts DAPs (adjusted <span class="html-italic">p</span>-value &lt; 0.05, absolute fold change &gt; 1.4) and significantly enriched metabolic pathways identified by MetaboAnalyst in DAPs specific to shoot (green), root (gray), sap proteome (orange) and in the overlap of these treatments. Red and blue dots indicate a relative increase and decrease in protein abundances compared to mock-treated control plants, respectively, while yellow dots represent differential responses between the comparisons. <span class="html-italic">S. tuberosum</span> proteins were annotated using the closest <span class="html-italic">Arabidopsis</span> orthologs. Generated using DiVenn online tool [<a href="#B47-plants-13-01983" class="html-bibr">47</a>]. For details on identified DAPs, see <a href="#app1-plants-13-01983" class="html-app">Supplementary Tables S1–S4</a>.</p>
Full article ">Figure 5
<p><span class="html-italic">H. vulgare</span> xylem sap proteome response to flg22. (<b>a</b>) The DiVenn visualization depicts DAPs (adjusted <span class="html-italic">p</span>-value &lt; 0.05) found in the third collected fraction of <span class="html-italic">H. vulgare</span> xylem sap proteome and flg22 response proteins identified in <span class="html-italic">A. thaliana</span> [<a href="#B48-plants-13-01983" class="html-bibr">48</a>]. Red and blue dots indicate a relative increase and decrease in protein abundances compared to mock-treated control plants, respectively, while yellow dots represent differential responses between the comparisons; (<b>b</b>) visualization of functional categories in the ProteoMap. Based on estimated protein abundances of DAPs that are predicted to be extracellular proteins. The size of each category corresponds to the estimated protein abundance. <span class="html-italic">H. vulgare</span> proteins were annotated using the closest <span class="html-italic">Arabidopsis</span> orthologs. For details on identified DAPs, see <a href="#app1-plants-13-01983" class="html-app">Supplementary Tables S5 and S6</a>.</p>
Full article ">
16 pages, 4102 KiB  
Article
The TIR1/AFB Family in Solanum melongena: Genome-Wide Identification and Expression Profiling under Stresses and Picloram Treatment
by Wenchao Du, Umer Karamat, Liuqing Cao, Yunpeng Li, Haili Li, Haoxin Li, Lai Wei, Dongchen Yang, Meng Xia, Qiang Li and Xueping Chen
Agronomy 2024, 14(7), 1413; https://doi.org/10.3390/agronomy14071413 - 28 Jun 2024
Viewed by 637
Abstract
TIR1/AFB proteins are a class of auxin receptors with key roles in plant development and biotic and abiotic stress responses; several have been identified as targets of the auxin-mimicking herbicide picloram. In this study, we identified five putative TIR1/AFB gene family members in [...] Read more.
TIR1/AFB proteins are a class of auxin receptors with key roles in plant development and biotic and abiotic stress responses; several have been identified as targets of the auxin-mimicking herbicide picloram. In this study, we identified five putative TIR1/AFB gene family members in the important vegetable crop Solanum melongena (eggplant) and characterized them using bioinformatics tools and gene expression analyses. Phylogenetic analysis of the TIR1/AFBs classified them into three subgroups based on their Arabidopsis and Solanum lycopersicum homologs. AFB6 homologs were present only in S. melongena and S. lycopersicum, whereas AFB2/3 homologs were found only in Arabidopsis. One pair of S. melongena TIR1 homologs were located in syntenic regions in the genome and appeared to have arisen by segmental duplication. Promoter analysis revealed 898 cis-elements in the TIR1/AFB promoters, 125 of which were related to hormones, stress, light, or growth responses, but only SmAFB5 had a cis-acting regulatory element involved in auxin responsiveness (AuxRR-core). RNA sequencing and expression profiling showed that the TIR1/AFB genes were differentially expressed at different growth stages and in response to light, temperature, and drought. Only SmTIR1A expression was significantly induced by picloram treatment and different growth stages. TIR1/AFB expression is regulated by microRNAs (miRNAs) in other plant species, and we identified 6 or 29 miRNAs that potentially targeted the five TIR1/AFB genes on the basis of comparisons with S. lycopersicum and S. tuberosum miRNAs, respectively. Three-dimensional protein structure predictions revealed that all the TIR1/AFB proteins were very similar in structure, differing only in the numbers of alpha helices and in one angle linking an α helix and a β sheet. For measuring the function of TIR1/AFB genes in response to drought, SmAFB5 was selected, and knockdown by virus-induced gene silence (VIGS) 35S::SmAFB5 lines showed resistance to drought compared to controls. These analyses provide insight into the potential functions of TIR1/AFBs during growth and in response to stress; they highlight differences among the SmTIR1/AFBs that may be useful for eggplant breeding. Full article
Show Figures

Figure 1

Figure 1
<p>Locations of the five <span class="html-italic">SmTIR1/AFB</span> genes on four of the twelve <span class="html-italic">S. melongena</span> chromosomes.</p>
Full article ">Figure 2
<p>Phylogenetic analysis of TIR1/AFB proteins. Phylogenetic tree of TIR1/AFB proteins from <span class="html-italic">Arabidopsis thaliana</span>, <span class="html-italic">S. lycopersicum</span>, and <span class="html-italic">S. melongena.</span> Phylogenetic analysis of five TIR1/AFB proteins identified in <span class="html-italic">S. melongena. S. melongena</span> TIR1/AFBs were present in three clades and named with reference to their <span class="html-italic">Arabidopsis</span> homologs. Clustal W was used to align the full-length protein sequences and MEGA X was used to construct neighbor-joining phylogenetic trees with 1000 bootstrap replicates.</p>
Full article ">Figure 3
<p>Gene structures (<b>left</b>) and conserved protein motifs and domains (<b>right</b>) of the <span class="html-italic">S. melongena TIR1/AFBs.</span></p>
Full article ">Figure 4
<p>Circos plot showing the <span class="html-italic">SmTIR1B</span> and <span class="html-italic">SmTIR1C</span> gene pair (connected in red) that appeared to have arisen via segmental duplication. Gray lines indicate collinear gene blocks.</p>
Full article ">Figure 5
<p>Identification of <span class="html-italic">cis</span>-acting elements in the <span class="html-italic">TIR1/AFB</span> gene promoters and effects of light and temperature on <span class="html-italic">TIR1/AFB</span> gene expression. (<b>A</b>) <span class="html-italic">cis</span>-acting elements identified in the promoters of the <span class="html-italic">TIR1/AFB</span> genes. ARE, auxin-responsive element. LRE, light-responsive element. ABRE, abscisic acid-responsive element. GAE, gibberellic acid-responsive element. SRE, stress-responsive element. MeJA, methyl jasmonate-related element. SAR, salicylic acid-responsive element. GRE, growth-related element. MYC, MYC element. MYB, MYB binding site involved in light response. AuxRR, an auxin-responsive <span class="html-italic">cis</span>-element present only in <span class="html-italic">SmAFB5</span>, highlighted in a red box. (<b>B</b>) Numbers of different <span class="html-italic">cis</span>-acting elements in the <span class="html-italic">TIR1/AFB</span> promoters. (<b>C</b>) Numbers of <span class="html-italic">cis</span>-acting elements related to hormones (HRE, including ARE, MeJA, SAR, ABAR, and GAE), light (LRE), stress (SRE), and growth (GRE) in the promoters of individual <span class="html-italic">TIR1/AFB</span> genes.</p>
Full article ">Figure 6
<p>Predicted 3D structures of the <span class="html-italic">S. melongena</span> TIR1/AFB proteins. Coiled shapes represent α helices, flat ribbon shapes represent β sheets, and red circles highlight a single angle that showed subtle differences among the proteins.</p>
Full article ">Figure 7
<p>Number of miRNAs predicted to target <span class="html-italic">TIR1/AFB</span> genes based on miRNAs identified in <span class="html-italic">S. lycopersicum</span> (<b>A</b>) and <span class="html-italic">S. tuberosum</span> (<b>B</b>).</p>
Full article ">Figure 8
<p>Expression of the <span class="html-italic">SmTIR1/AFB</span> genes. (<b>A</b>) Expression of the <span class="html-italic">TIR1/AFB</span> genes in leaves at the four-leaf stage (W4) and six-leaf stage (W6). (<b>B</b>) Expression of <span class="html-italic">TIR1/AFB</span> genes in leaves 6 h after treatment with water (CK) or the herbicide picloram (P). 1, 2, and 3 after hyphen represent three independent technical replicates of each treatment, |log<sub>2</sub>FC| ≥ 1, FDR ≤ 0.01. (<b>C</b>) Expression of <span class="html-italic">TIR1/AFB</span> genes in the dark at normal temperature and irrigation (BN, 25 °C), in the light at normal temperature (LN, 25 °C), in the dark at low temperature (BL, 10 °C), and in the dark with drought stress (BD, watering ceased for 6 days), as measured by RT–qPCR. ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.1 (Student’s <span class="html-italic">t</span>-test, compared with BN).</p>
Full article ">Figure 9
<p>Expression of <span class="html-italic">SmTIR1/AFB</span> genes at the two-leaf and four-leaf stages (<b>A</b>) and in response to picloram treatment (<b>B</b>), as measured by RT–qPCR (bars) and RNA-seq (red symbols). For RT–qPCR, gene expression levels were normalized to those of <span class="html-italic">Smechr0302615</span> (tubulin gamma). In (<b>B</b>), RT–qPCR expression levels in the CK were set to 1. Data are presented as the means ± SE of three biological replicates. * <span class="html-italic">p</span> &lt; 0.1 (Student’s <span class="html-italic">t</span>-test, RNA-seq data).</p>
Full article ">Figure 10
<p>Effect of silencing <span class="html-italic">SmAFB5</span> on eggplant susceptibility to drought. (<b>A</b>) Expression of <span class="html-italic">SmAFB5</span> in leaves 30 days after VIGS. The asterisks indicate statistically significant differences as determined by Student’s <span class="html-italic">t</span>-test (two-tailed). * <span class="html-italic">p</span> &lt; 0.1. (<b>B</b>) Representative seedlings at 12 h recovery after drought treatment. TRV2::00 (TRV2::00 plants treated with drought), <span class="html-italic">TRV2::SmAFB5</span> (TRV2::<span class="html-italic">SmAFB5</span> plants treated with drought). Bar = 1 cm.</p>
Full article ">
27 pages, 12535 KiB  
Review
Modern Breeding Strategies and Tools for Durable Late Blight Resistance in Potato
by Ioana Virginia Berindean, Abdelmoumen Taoutaou, Soumeya Rida, Andreea Daniela Ona, Maria Floriana Stefan, Alexandru Costin, Ionut Racz and Leon Muntean
Plants 2024, 13(12), 1711; https://doi.org/10.3390/plants13121711 - 20 Jun 2024
Viewed by 1721
Abstract
Cultivated potato (Solanum tuberosum) is a major crop worldwide. It occupies the second place after cereals (corn, rice, and wheat). This important crop is threatened by the Oomycete Phytophthora infestans, the agent of late blight disease. This pathogen was first [...] Read more.
Cultivated potato (Solanum tuberosum) is a major crop worldwide. It occupies the second place after cereals (corn, rice, and wheat). This important crop is threatened by the Oomycete Phytophthora infestans, the agent of late blight disease. This pathogen was first encountered during the Irish famine during the 1840s and is a reemerging threat to potatoes. It is mainly controlled chemically by using fungicides, but due to health and environmental concerns, the best alternative is resistance. When there is no disease, no treatment is required. In this study, we present a summary of the ongoing efforts concerning resistance breeding of potato against this devastating pathogen, P. infestans. This work begins with the search for and selection of resistance genes, whether they are from within or from outside the species. The genetic methods developed to date for gene mining, such as effectoromics and GWAS, provide researchers with the ability to identify genes of interest more efficiently. Once identified, these genes are cloned using molecular markers (MAS or QRL) and can then be introduced into different cultivars using somatic hybridization or recombinant DNA technology. More innovative technologies have been developed lately, such as gene editing using the CRISPR system or gene silencing, by exploiting iRNA strategies that have emerged as promising tools for managing Phytophthora infestans, which can be employed. Also, gene pyramiding or gene stacking, which involves the accumulation of two or more R genes on the same individual plant, is an innovative method that has yielded many promising results. All these advances related to the development of molecular techniques for obtaining new potato cultivars resistant to P. infestans can contribute not only to reducing losses in agriculture but especially to ensuring food security and safety. Full article
(This article belongs to the Section Plant Protection and Biotic Interactions)
Show Figures

Figure 1

Figure 1
<p>Symptoms of late blight disease. The typical symptoms of late blight: on the upper side of the leaf, an oily necrotic spot, surrounded by pale green (<b>A</b>); on the underside of the leaf: a white down is observed (<b>B</b>). This white down is the pathogens sporangiophres and sporanges (<b>C</b>). Stem and petioles could also be attacked (<b>D<sub>1</sub></b>). Advanced disease is manifested by a blight of the leaves and possibly the whole plant (<b>D<sub>2</sub></b>,<b>D<sub>3</sub></b>). Photograph: A. Taoutaou.</p>
Full article ">Figure 2
<p>Breeding strategies of resistance for <span class="html-italic">P. infestans</span>. There are mainly two strategies for breeding: enhance the plant resistance and/or alter its susceptibility to the pathogen. There are two types of resistance: the first one is called qualitative and is total (the plant is or resistant, or susceptible); the second, called quantitative, is partial (the plant could have some degree of resistance/susceptibility). Legend: x: rejected for both qualitative and quantitative resistance breeding (the plant is too susceptible to the disease); +: accepted for quantitative resistance (there is an enhancement of the resistance: the plant is more resistant than the starting material); ✩: Accepted for qualitative resistance: the plant is totally resistant to the pathogen, and there is no sign of the disease.</p>
Full article ">Figure 3
<p>Schematic representation of the structure of a resistance gene of potato to late blight disease. <span class="html-italic">R</span> genes are composed of three major conserved domains: the first one can be a coiled or Toll Interleukin domain. In the middle is a Nucleotide binding site. This domain also contains several conserved regions: P-loop, Kinase2, Kinase 3a, ARC1, and ARC2. On the other end, the Leucine rich-repeat domain is found. Legend: CC: coiled coil, TIR: Tall Interleukin 1 Receptors, NBS: Nucleotide binding site, LR: Leucine rich, LRR: Leucine rich-repeat.</p>
Full article ">
15 pages, 2328 KiB  
Article
Identification, Elucidation and Deployment of a Cytoplasmic Male Sterility System for Hybrid Potato
by Ernst-Jan Eggers, Ying Su, Esmee van der Poel, Martijn Flipsen, Michiel E. de Vries, Christian W. B. Bachem, Richard G. F. Visser and Pim Lindhout
Biology 2024, 13(6), 447; https://doi.org/10.3390/biology13060447 - 18 Jun 2024
Viewed by 1098
Abstract
Recent advances in diploid F1 hybrid potato breeding rely on the production of inbred lines using the S-locus inhibitor (Sli) gene. As a result of this method, female parent lines are self-fertile and require emasculation before hybrid seed production. The [...] Read more.
Recent advances in diploid F1 hybrid potato breeding rely on the production of inbred lines using the S-locus inhibitor (Sli) gene. As a result of this method, female parent lines are self-fertile and require emasculation before hybrid seed production. The resulting F1 hybrids are self-fertile as well and produce many undesirable berries in the field. Utilization of cytoplasmic male sterility would eliminate the need for emasculation, resulting in more efficient hybrid seed production and male sterile F1 hybrids. We observed plants that completely lacked anthers in an F2 population derived from an interspecific cross between diploid S. tuberosum and S. microdontum. We studied the antherless trait to determine its suitability for use in hybrid potato breeding. We mapped the causal locus to the short arm of Chromosome 6, developed KASP markers for the antherless (al) locus and introduced it into lines with T and A cytoplasm. We found that antherless type male sterility is not expressed in T and A cytoplasm, proving that it is a form of CMS. We hybridized male sterile al/al plants with P cytoplasm with pollen from al/al plants with T and A cytoplasm and we show that the resulting hybrids set significantly fewer berries in the field. Here, we show that the antherless CMS system can be readily deployed in diploid F1 hybrid potato breeding to improve hybridization efficiency and reduce berry set in the field. Full article
(This article belongs to the Special Issue Pollination Biology)
Show Figures

Figure 1

Figure 1
<p>Population BC2(P)-1 segregates for anther length. (<b>a</b>) Anther phenotypes in BC<sub>1</sub> population BC2(P)-1 scored from 0 (complete absence of anthers) to 3 (normal anthers). (<b>b</b>) Genetic map of Chromosome 6. (<b>c</b>) QTL analysis for anther score reveals a significant QTL on the top of Chromosome 6. (<b>d</b>) Segregation distortion on the top of Chromosome 6 reduces the number of antherless plants. In red, the fraction of plants heterozygous per locus is shown and in green the fraction of plants homozygous for the antherless donor allele is shown.</p>
Full article ">Figure 2
<p>Flowers of alal homozygous genotypes with A and T cytoplasm have anthers that produce pollen and set self-seed.</p>
Full article ">Figure 3
<p>Berry yield of antherless POC hybrids and male fertile controls in a field trial. Antherless hybrids (n = 26) produce significantly fewer berries than male fertile controls (n = 8) (<span class="html-italic">p</span> &lt; 0.001). The boxes represent the 1st and 3rd quartiles, the horizontal lines in the box represent the medians, and the × represent the means.</p>
Full article ">
19 pages, 8003 KiB  
Article
Economic and Environmental Assessment of Variable Rate Nitrogen Application in Potato by Fusion of Online Visible and Near Infrared (Vis-NIR) and Remote Sensing Data
by Muhammad Qaswar, Danyal Bustan and Abdul Mounem Mouazen
Soil Syst. 2024, 8(2), 66; https://doi.org/10.3390/soilsystems8020066 - 14 Jun 2024
Viewed by 872
Abstract
Addressing within-field spatial variability for nitrogen (N) management to avoid over and under-use of nitrogen is crucial for optimizing crop productivity and ensuring environmental sustainability. In this study, we investigated the economic, environmental, and agronomic benefits of variable rate nitrogen application in potato [...] Read more.
Addressing within-field spatial variability for nitrogen (N) management to avoid over and under-use of nitrogen is crucial for optimizing crop productivity and ensuring environmental sustainability. In this study, we investigated the economic, environmental, and agronomic benefits of variable rate nitrogen application in potato (Solanum tuberosum L.). An online visible and near-infrared (vis-NIR) spectroscopy sensor was utilized to predict soil moisture content (MC), pH, total organic carbon (TOC), extractable phosphorus (P), potassium (K), magnesium (Mg), and cation exchange capacity (CEC) using a partial least squares regression (PLSR) models. The crop’s normalized difference vegetation index (NDVI) from Sentinel-2 satellite images was incorporated into online measured soil data to derive fertility management zones (MZs) maps after homogenous raster and clustering analyses. The MZs maps were categorized into high fertile (VR-H), medium–high fertile (VR-MH), medium–low fertile (VR-ML), and low fertile (VR-L) zones. A parallel strip experiment compared variable rate nitrogen (VR-N) with uniform rate (UR) treatments, adjusting nitrogen levels based on fertility zones as 50% less for VR-H, 25% less for VR-MH, 25% more for VR-ML, and 50% more for VR-L zones compared to the UR treatment. The results showed that the VR-H zone received a 50% reduction in N fertilizer input and demonstrated a significantly higher crop yield compared to the UR treatment. This implies a potential reduction in negative environmental impact by lowering fertilizer costs while maintaining robust crop yields. In total, the VR-N treatment received an additional 1.2 Kg/ha of nitrogen input, resulting in a crop yield increase of 1.89 tons/ha. The relative gross margin for the VR-N treatment compared to the UR treatment is 374.83 EUR/ha, indicating substantial profitability for the farmer. To further optimize environmental benefits and profitability, additional research is needed to explore site-specific applications of all farm resources through precision agricultural technologies. Full article
Show Figures

Figure 1

Figure 1
<p>Location of experimental field in Belgium along with online spectral lines (red) and sampling points (green point).</p>
Full article ">Figure 2
<p>Overall methodology followed in this study. PLSR, partial least square regression; NDVI, normalized difference vegetation index; MZs, management zones; VR-N, variable-rate nitrogen.</p>
Full article ">Figure 3
<p>Online multi-sensor platform used for soil data collection.</p>
Full article ">Figure 4
<p>Strip experiment map comparing variable rate nitrogen (VR-N) fertilization treatment against uniform rate UN treatment. Abbreviations: UR, uniform rate application; VR-H, variable rate high fertile zone; VR-MH, VR medium–high fertile zone; VR-ML, VR medium–low fertile zone; VR-L, VR low fertile zone.</p>
Full article ">Figure 5
<p>Maps of online predicted soil properties [organic carbon (TOC) (<b>a</b>), pH (<b>b</b>), phosphorous (P) (<b>c</b>), potassium (K) (<b>d</b>), magnesium (Mg) (<b>e</b>) and cation exchange capacity (CEC) (<b>f</b>)], crop normalized difference vegetation index (NDVI) (<b>g</b>) and crop yield (<b>h</b>).</p>
Full article ">Figure 6
<p>Management zones (MZs) map developed based on k-means clustering of the online predicted soil properties and normalized difference vegetation index (NDVI). Abbreviations: H, high fertile zone; MH, medium-high fertile zone; ML, medium-low fertile zone; L, low fertile zone; TOC, total organic carbon; P, phosphorous; K, potassium; Mg, magnesium; CEC, cation exchange capacity.</p>
Full article ">Figure 7
<p>Crop yield calculated for the uniform rate nitrogen (N) treatment and the per individual management zone (MZ) variable rate N treatment. Error bars depict the standard deviations (±), and distinct letters positioned above the bars indicate a significant difference (<span class="html-italic">p</span> ≤ 0.05) based on Tukey’s HSD test. UR, uniform rate (control); VR-H, the variable rate in a high fertile zone; VR-MH, VR in medium–high; VR-ML, VR in medium–low; VR-L, VR in a low fertile zone.</p>
Full article ">Figure 8
<p>Variable importance determined by the random forest (RF) model by using spatial raster data as a predictor variables and crop yield as a response. Abbreviations: TOC, total organic carbon; P, phosphorus; K, potassium; Mg, magnesium; CEC, cation exchange capacity.</p>
Full article ">
15 pages, 2428 KiB  
Article
Physiological and Enzymatic Antioxidant Responses of Solanum tuberosum Leaves to Arbuscular Mycorrhizal Fungal Inoculation under Water Stress
by Javiera Nahuelcura, Catalina Bravo, Analía Valdebenito, Sheina Rivas, Christian Santander, Felipe González, Pablo Cornejo, Boris Contreras and Antonieta Ruiz
Plants 2024, 13(8), 1153; https://doi.org/10.3390/plants13081153 - 21 Apr 2024
Viewed by 1230
Abstract
Solanum tuberosum is one of the most widely cropped plant species worldwide; unfortunately, drought is one of the major constraints on potato productivity because it affects the physiology, biochemical processes, and yield. The use of arbuscular mycorrhizal fungi (AMF) has exhibited beneficial effects [...] Read more.
Solanum tuberosum is one of the most widely cropped plant species worldwide; unfortunately, drought is one of the major constraints on potato productivity because it affects the physiology, biochemical processes, and yield. The use of arbuscular mycorrhizal fungi (AMF) has exhibited beneficial effects on plants during drought. The objective of this study was to analyse the effect of AMF inoculation on two genotypes of potato plants exposed to water stress, and the photosynthetic traits, enzymatic antioxidant activity, and exudation of low-molecular-weight organic acids (LMWOAs) of potato plants inoculated with two strains of AMF, Claroideoglomus claroideum (CC) and Claroideoglomus lamellosum (HMC26), were evaluated. Stomatal conductance exhibited a similar trend in the CC and HMC26 treatments for both potato genotypes; moreover, the photosynthetic rate significantly increased by 577.9% between the 100% soil humidity (S0) and 40% soil humidity (S2) stress levels for the VR808 genotype under the CC treatment. The activities of the enzymes catalase (CAT) and ascorbate peroxidase (APX) showed similar trends. In this study, there were different responses among genotypes and treatments. Inoculation with CC under S2 stress levels is a promising potential approach for improving potato growth under drought conditions. Full article
(This article belongs to the Section Plant Response to Abiotic Stress and Climate Change)
Show Figures

Figure 1

Figure 1
<p>Photosynthetic traits measured in leaves of two genotypes of <span class="html-italic">Solanum tuberosum</span> plants inoculated with arbuscular mycorrhizal fungi (AMF) and growing under normal irrigation and two drought conditions. (<b>A</b>) Quantum yield of PSII (ΦPSII); (<b>B</b>) stomatal conductance (gs); (<b>C</b>) photosynthetic rate (A); (<b>D</b>) internal CO<sub>2</sub> concentration (Ci); (<b>E</b>) water use efficiency (WUE). Here, NM: non-inoculated plants, CC: plants inoculated with the fungus <span class="html-italic">Claroideoglomus claroideum</span>, HMC26: plants inoculated with the fungus <span class="html-italic">Claroideoglomus lamellosum</span>, MIX: CC + HMC26; S0: 100%, S1: 70%, S2: 40% of water-holding capacity levels; VR808: yellow skin and yellow flesh genotype, CB2011-104: purple skin and purple flesh genotype. Different letters indicate significant differences according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Concentrations of photosynthetic pigments of two genotypes of <span class="html-italic">Solanum tuberosum</span> plants inoculated with arbuscular mycorrhizal fungi (AMF) and grown under normal irrigation and two drought conditions. (<b>A</b>) Total chlorophyll; (<b>B</b>) chlorophyll a; (<b>C</b>) chlorophyll b; (<b>D</b>) carotenoids. Here, NM: non-inoculated plants, CC: plants inoculated with the fungus <span class="html-italic">Claroideoglomus claroideum</span>, HMC26: plants inoculated with the fungus <span class="html-italic">Claroideoglomus lamellosum</span>, MIX: CC + HMC26; S0: 100%, S1: 70%, S2: 40% of water-holding capacity levels; VR808: yellow skin and yellow flesh genotype, CB2011-104: purple skin and purple flesh genotype. Different letters indicate significant differences according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Antioxidant enzymatic activity levels of two genotypes of <span class="html-italic">Solanum tuberosum</span> plants inoculated with arbuscular mycorrhizal fungi (AMF) and grown under normal irrigation and two drought conditions: (<b>A</b>) Catalase (CAT) enzyme; (<b>B</b>) ascorbate peroxidase (APX) enzyme; (<b>C</b>) glutathione reductase (GR) enzyme. Here, NM: non-inoculated plants, CC: plants inoculated with the fungus <span class="html-italic">Claroideoglomus claroideum</span>, HMC26: plants inoculated with the fungus <span class="html-italic">Claroideoglomus lamellosum</span>, MIX: CC + HMC26; S0: 100%, S1: 70%, S2: 40% of water-holding capacity levels; VR808: yellow skin and yellow flesh genotype, CB2011-104: purple skin and purple flesh genotype. Different letters indicate significant differences according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Concentrations of low-molecular-weight organic acids (LMWOAs) in the rhizosphere of two genotypes of <span class="html-italic">Solanum tuberosum</span> plants inoculated with arbuscular mycorrhizal fungi (AMF) and grown under normal irrigation and two drought conditions: (<b>A</b>) Oxalic acid; (<b>B</b>) citric acid. Here, NM: non-inoculated plants, CC: plants inoculated with the fungus <span class="html-italic">Claroideoglomus claroideum</span>, HMC26: plants inoculated with the fungus <span class="html-italic">Claroideoglomus lamellosum</span>, MIX: CC + HMC26; S0: 100%, S1: 70%, S2: 40% of water-holding capacity levels; VR808: yellow skin and yellow flesh genotype, CB2011-104: purple skin and purple flesh genotype. Different letters indicate significant differences according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Principal components (PCs) for the experimental variables determined in two genotypes of <span class="html-italic">Solanum tuberosum</span> plants inoculated with arbuscular mycorrhizal fungi (AMF) and grown under normal irrigation and two drought conditions. The graph shows the experimental individuals according to PC grouped according to the (<b>A</b>) mycorrhizal treatments of genotype VR808; (<b>B</b>) water stress level of genotype VR808; (<b>C</b>) mycorrhizal treatments of genotype CB2011-104; and (<b>D</b>) water stress level of genotype CB2011-104. Here, NM: non-inoculated plants, CC: plants inoculated with the fungus <span class="html-italic">Claroideoglomus claroideum</span>, HMC26: plants inoculated with the fungus <span class="html-italic">Claroideoglomus lamellosum</span>, MIX: CC + HMC26; S0: 100%, S1: 70%, S2: 40% of water-holding capacity levels; VR808: yellow skin and yellow flesh genotype, CB2011-104: purple skin and purple flesh genotype. ΦPSII: Quantum yield of photosystem II, gs: Stomatic conductance, A: Photosynthetic rate, Ci: Leaf internal CO<sub>2</sub> concentration, WUE: Water use efficiency, TChl: Total chlorophyll, ChlA: Chlorophyll a, ChlB: Chlorophyll b, GR: Glutathione reductase enzyme activity, CAT: Catalase enzyme activity, APX: Ascorbate peroxidase enzyme activity.</p>
Full article ">
11 pages, 3350 KiB  
Article
A Walk on the Wild Side: Genome Editing of Tuber-Bearing Solanum bulbocastanum
by Aristotelis Azariadis, Olga A. Andrzejczak, Frida M. Carlsen, Ida Westberg, Henrik Brinch-Pedersen, Bent L. Petersen and Kim H. Hebelstrup
Plants 2024, 13(7), 1044; https://doi.org/10.3390/plants13071044 - 8 Apr 2024
Viewed by 1098
Abstract
Solanum bulbocastanum is a wild diploid tuber-bearing plant. We here demonstrate transgene-free genome editing of S. bulbocastanum protoplasts and regeneration of gene-edited plants. We use ribonucleoproteins, consisting of Cas9 and sgRNA, assembled in vitro, to target a gene belonging to the nitrate and [...] Read more.
Solanum bulbocastanum is a wild diploid tuber-bearing plant. We here demonstrate transgene-free genome editing of S. bulbocastanum protoplasts and regeneration of gene-edited plants. We use ribonucleoproteins, consisting of Cas9 and sgRNA, assembled in vitro, to target a gene belonging to the nitrate and peptide transporter family. Four different sgRNAs were designed and we observed efficiency in gene-editing in the protoplast pool between 8.5% and 12.4%. Twenty-one plants were re-generated from microcalli developed from individual protoplasts. In three of the plants we found that the target gene had been edited. Two of the edited plants had deletion mutations introduced into both alleles, whereas one only had a mutation in one of the alleles. Our work demonstrates that protocols for the transformation of Solanum tuberosum can be optimized to be applied to a wild Solanum species. Full article
(This article belongs to the Section Plant Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Optimization of protoplast isolation. Protoplast isolated using the non-optimized protocol for <span class="html-italic">S. tuberosum</span> showing viable and burst cells as indicated by green and red arrows, respectively, under white light (<b>A</b>) and fluorescein staining (<b>B</b>). Only very few viable <span class="html-italic">S. bulbocastanum</span> protoplasts could be isolated when using the protocol for <span class="html-italic">S. tuberosum</span>. Protoplast isolated after protocol adjustment, showing viable, spherical shapes with no debris present, under white light (<b>C</b>) and fluorescein staining (<b>D</b>).</p>
Full article ">Figure 2
<p>Sequence of the cDNA of the GOI with the sgRNAs targets depicted.</p>
Full article ">Figure 3
<p>IDAA profiles of edited protoplast cells. The amplicon size of 453 bp of the wild type (WT) was used as reference (<b>A</b>). Small peaks from 440 to 452 bp’s indicated deletions between 1 to 13 bp in batches of protoplasts. The peak areas of the deletions for each sgRNA combination (<b>B</b>–<b>D</b>) was calculated as percentages, as described in [<a href="#B12-plants-13-01044" class="html-bibr">12</a>]. Dashed line indicates amplicon size in WT plants.</p>
Full article ">Figure 4
<p>Timeline of the regeneration protocol. Isolated protoplasts prior to transformation (<b>A</b>). Day 20 post protoplast isolation (ppi) showing protoplast embedded in alginate lenses (<b>B</b>). Day 50 ppi, prior to release from the alginate lenses (<b>C</b>). Day 64 ppi, hardened calli showing initiation of early plant growth (<b>D</b>). Calli with premature root emergence (<b>D1</b>,<b>D2</b>). Day 90 ppi, presence of shoots of ~10–20 mm length (<b>E</b>). Day ~120 ppi, fully regenerated plants grown in propagating medium (<b>F</b>).</p>
Full article ">Figure 5
<p>IDAA profiles of regenerated individual plants. The wild-type control showed a peak at the 453 bp of the two WT alleles (<b>A</b>). Mutant 17 appeared monoallelic with a 2 bp deletion in 86.8% PCR products (<b>B</b>). Mutant 18 appeared to be heterozygous with a deletion of 7 bp in a single allele (<b>C</b>). Mutant 2 appeared to be biallelic with deletions of −8 and −3 bp, respectively (<b>D</b>). Dashed line and yellow colour indicate the size of the WT gene. The blue colour indicates a deletion mutation.</p>
Full article ">Figure 6
<p>Mutations in haplotypes from edited plants 2, 17 and 18. Asterisk at the end of the red region indicates a premature stop codon (for details see <a href="#app1-plants-13-01044" class="html-app">Supplementary Materials S3</a>). The blue arrows indicate a continuation of the sequence.</p>
Full article ">
24 pages, 1752 KiB  
Review
Looking for Resistance to Soft Rot Disease of Potatoes Facing Environmental Hypoxia
by Tomasz Maciag, Edmund Kozieł, Katarzyna Otulak-Kozieł, Sylwia Jafra and Robert Czajkowski
Int. J. Mol. Sci. 2024, 25(7), 3757; https://doi.org/10.3390/ijms25073757 - 28 Mar 2024
Cited by 1 | Viewed by 1369
Abstract
Plants are exposed to various stressors, including pathogens, requiring specific environmental conditions to provoke/induce plant disease. This phenomenon is called the “disease triangle” and is directly connected with a particular plant–pathogen interaction. Only a virulent pathogen interacting with a susceptible plant cultivar will [...] Read more.
Plants are exposed to various stressors, including pathogens, requiring specific environmental conditions to provoke/induce plant disease. This phenomenon is called the “disease triangle” and is directly connected with a particular plant–pathogen interaction. Only a virulent pathogen interacting with a susceptible plant cultivar will lead to disease under specific environmental conditions. This may seem difficult to accomplish, but soft rot Pectobacteriaceae (SRPs) is a group virulent of pathogenic bacteria with a broad host range. Additionally, waterlogging (and, resulting from it, hypoxia), which is becoming a frequent problem in farming, is a favoring condition for this group of pathogens. Waterlogging by itself is an important source of abiotic stress for plants due to lowered gas exchange. Therefore, plants have evolved an ethylene-based system for hypoxia sensing. Plant response is coordinated by hormonal changes which induce metabolic and physiological adjustment to the environmental conditions. Wetland species such as rice (Oryza sativa L.), and bittersweet nightshade (Solanum dulcamara L.) have developed adaptations enabling them to withstand prolonged periods of decreased oxygen availability. On the other hand, potato (Solanum tuberosum L.), although able to sense and response to hypoxia, is sensitive to this environmental stress. This situation is exploited by SRPs which in response to hypoxia induce the production of virulence factors with the use of cyclic diguanylate (c-di-GMP). Potato tubers in turn reduce their defenses to preserve energy to prevent the negative effects of reactive oxygen species and acidification, making them prone to soft rot disease. To reduce the losses caused by the soft rot disease we need sensitive and reliable methods for the detection of the pathogens, to isolate infected plant material. However, due to the high prevalence of SRPs in the environment, we also need to create new potato varieties more resistant to the disease. To reach that goal, we can look to wild potatoes and other Solanum species for mechanisms of resistance to waterlogging. Potato resistance can also be aided by beneficial microorganisms which can induce the plant’s natural defenses to bacterial infections but also waterlogging. However, most of the known plant-beneficial microorganisms suffer from hypoxia and can be outcompeted by plant pathogens. Therefore, it is important to look for microorganisms that can withstand hypoxia or alleviate its effects on the plant, e.g., by improving soil structure. Therefore, this review aims to present crucial elements of potato response to hypoxia and SRP infection and future outlooks for the prevention of soft rot disease considering the influence of environmental conditions. Full article
(This article belongs to the Special Issue Advances in Plant–Pathogen Interactions: 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>The disease triangle for soft rot disease. All three components are necessary for the induction of soft rot disease. (1) Virulent strain of soft rot <span class="html-italic">Pectobacteriaceae</span>, whose virulence depends greatly on the production of cell wall degrading enzymes, but also other virulence factors such as the production of toxins [<a href="#B31-ijms-25-03757" class="html-bibr">31</a>], like necrose inducing protein (Nip) [<a href="#B32-ijms-25-03757" class="html-bibr">32</a>] and the type 4 [<a href="#B33-ijms-25-03757" class="html-bibr">33</a>] and 6 secretion systems [<a href="#B34-ijms-25-03757" class="html-bibr">34</a>]. (2) Susceptible hosts such as potatoes, currently there are no potato varieties resistant to soft rot disease [<a href="#B35-ijms-25-03757" class="html-bibr">35</a>,<a href="#B36-ijms-25-03757" class="html-bibr">36</a>]. (3) Disease-favoring conditions which are high humidity and temperature and, resulting from high humidity, hypoxia [<a href="#B37-ijms-25-03757" class="html-bibr">37</a>].</p>
Full article ">Figure 2
<p>Plant ethylene signaling leads to the expression of genes, leading to waterlogging adaptations. Constantly produced ethylene is trapped in submerged tissues due to the low diffusion of nonpolar gases in water and cannot escape into the atmosphere. Ethylene is then bound by ethylene receptors, inhibiting their activity. In the canonical pathway, ethylene receptors bind ethylene preventing them from activating constitutive triple response 1 (CTR1). Inactive CTR1 kinase no longer phosphorylates ethylene-insensitive 2 (EIN2), leading to its activation and following the activation of hypoxia transcription factors. The non-canonical pathway (marked by a dotted line) starts with a suppressed ETR1 receptor, which no longer activates CTR1 phosphorylates Arabidopsis histidine-containing phosphotransfer proteins (AHPs) (linking the ethylene sensing with cytokine signaling), which in turn activate Arabidopsis response regulators (ARR) by phosphotransfer. ARRs regulate the expression of waterlogging response genes. Prepared based on [<a href="#B43-ijms-25-03757" class="html-bibr">43</a>,<a href="#B45-ijms-25-03757" class="html-bibr">45</a>]. The red arrows demonstrate inhibition and the green arrows represent induction.</p>
Full article ">Figure 3
<p>Soft rot <span class="html-italic">Pectobacteriaceae</span> response to hypoxia. High humidity in soil leads to decreased oxygen penetration into deeper soil. This, in turn, forces the bacteria to switch from aerobic metabolism to fermentation. Low oxygen conditions lead to the increased production of master regulator c-di-GMP. As in many other bacteria groups, hypoxia in pectinolytic bacteria leads to the activation of reactive oxygen species defenses. Pectynolytic bacteria also increase nitrate respiration production of lytic enzymes and virulence factors. Global regulator c-di-GMP favors biofilm formation versus motility, triggering soft rot <span class="html-italic">Pectobacteriaceae</span>’s pathogenicity. Prepared based on [<a href="#B89-ijms-25-03757" class="html-bibr">89</a>,<a href="#B91-ijms-25-03757" class="html-bibr">91</a>,<a href="#B92-ijms-25-03757" class="html-bibr">92</a>,<a href="#B93-ijms-25-03757" class="html-bibr">93</a>,<a href="#B94-ijms-25-03757" class="html-bibr">94</a>,<a href="#B95-ijms-25-03757" class="html-bibr">95</a>]. The red arrows demonstrate inhibition and the green arrows represent induction.</p>
Full article ">Figure 4
<p>Potato tuber response to hypoxia. High humidity leads to decreased gas exchange and hypoxia. During hypoxia, potato tubers decrease aerobic respiration to preserve oxygen and increase anaerobic respiration to provide energy. Increased anaerobic respiration produces acidification and reactive oxygen species (ROS). To protect the cells from the damaging effect of ROS, plants need to activate their defenses against ROS and produce heat shock proteins. A low energetic status abolishes plant defenses, including wound response factors such as extensins, phenylalanine ammonia lyase (PAL), and secondary metabolites such as expansin (EXP), early nodulin 93 (ENOD), 4-coumarate CoA ligase-like 2 (4CL), and HMG-CoA reductase (HMG). The production of proteins is arrested by EF-1α binding to ribosomes in acidic conditions. Plant hormonal response tailored for cell elongation to escape local hypoxic conditions makes tubers prone to the attack of necrotic pathogens. Prepared based on [<a href="#B26-ijms-25-03757" class="html-bibr">26</a>,<a href="#B83-ijms-25-03757" class="html-bibr">83</a>,<a href="#B85-ijms-25-03757" class="html-bibr">85</a>,<a href="#B86-ijms-25-03757" class="html-bibr">86</a>,<a href="#B100-ijms-25-03757" class="html-bibr">100</a>,<a href="#B101-ijms-25-03757" class="html-bibr">101</a>]. The red arrows demonstrate inhibition and the green arrows represent induction.</p>
Full article ">
14 pages, 2378 KiB  
Article
The Growth and Tuber Yield of Potatoes (Solanum tuberosum L.) under Varying LED Light Spectrums in Controlled Greenhouse Conditions
by Md Hafizur Rahman, Md. Jahirul Islam, Umma Habiba Mumu, Byeong-Ryeol Ryu, Jung-Dae Lim, Md Obyedul Kalam Azad, Eun Ju Cheong and Young-Seok Lim
Horticulturae 2024, 10(3), 254; https://doi.org/10.3390/horticulturae10030254 - 7 Mar 2024
Viewed by 1658
Abstract
Plant growing using light-emitting diodes (LEDs) in a controlled environment is a revolutionary and innovative idea, regardless of the external environmental disturbances. Studying the growth and tuber yield of potatoes (Solanum tuberosum L.) in an LED-based plant factory system is a relatively [...] Read more.
Plant growing using light-emitting diodes (LEDs) in a controlled environment is a revolutionary and innovative idea, regardless of the external environmental disturbances. Studying the growth and tuber yield of potatoes (Solanum tuberosum L.) in an LED-based plant factory system is a relatively innovative concept. The current study was conducted in a plant factory to evaluate the effects of different LED spectral compositions on potato tuberization. Potato tuberization was analyzed under six different LED light spectral combinations with irradiances of 300 mol m−2 s−1, with natural light considered the control treatment. The findings stated that the L2 treatment (red70 + blue20 + white10) increased the plant height, branch number, and biomass accumulation, while photosynthetic pigments and photosynthetic activity increased significantly in L5 (red60 + blue20 + green10 + white10). Higher gibberellic acid (GA3) content was recorded in L1 (red70 + blue30), whereas the tuber number and tuber fresh weight were recorded in L3 (red70 + blue20 + green10) and L7 (natural light), respectively. On the other hand, a higher number of smaller-sized tubers were observed in L5, while L2 and L4 (red70 + blue20 + far-red10) resulted in a higher number of medium-sized tubers. In conclusion, a high proportion of red and blue light, along with white and far-red light, increased the plant height, branch number, plant biomass, and production of small- and medium-sized tubers. On the other hand, the inclusion of green light with red and blue enhanced the chlorophyll content, photosynthesis, and leaf expansion, and promoted the production of smaller-sized tubers. Finally, with regard to tuberization, the treatment using L4 followed by L2 outperformed the other treatments. Full article
Show Figures

Figure 1

Figure 1
<p>Photograph of potato plant grown under different LED spectrums: (<b>L1</b>)—red70 + blue30; (<b>L2</b>)—red70 + blue20 + white10; (<b>L3</b>)—red70 + blue20 + green10; (<b>L4</b>)—red70 + blue20 + far-red10; (<b>L5</b>)—red60 + blue20 + green10 + white10; (<b>L6</b>)—red60 + blue20 + far-red10 + white10; (<b>L7</b>)—natural light.</p>
Full article ">Figure 2
<p>Chlorophyll a (<b>A</b>), chlorophyll b (<b>B</b>), carotenoid (<b>C</b>), total chlorophyll (<b>D</b>), and SPAD index of potato plants grown under different LED light spectrums in artificial soil pot culture (<b>E</b>). Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among the genotypes within each parameter. Values are expressed as mean ± SD (<span class="html-italic">n</span> = 3). L1—red70 + blue30; L2—red70 + blue20 + white10; L3—red70 + blue20 + green10; L4—red70 + blue20 + far-red10; L5—red60 + blue20 + green10 + white10; L6—red60 + blue20 + far-red10 + white10; L7—natural light.</p>
Full article ">Figure 3
<p>Net photosynthetic rate (<b>A</b>), transpiration rate (<b>B</b>), stomatal conductance (<b>C</b>), and water use efficiency (<b>D</b>) of potato plants grown under different LED spectrums in artificial soil pot culture. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among the genotypes within each parameter. Values are expressed as mean ± SD (<span class="html-italic">n</span> = 3). L1—red70 + blue30; L2—red70 + blue20 + white10; L3—red70 + blue20 + green10; L4—red70 + blue20 + far-red10; L5—red60 + blue20 + green10 + white10; L6—red60 + blue20 + far-red10 + white10; L7—natural light.</p>
Full article ">Figure 4
<p>Tuber number (TN/plant) and tuber fresh weight (TFW/plant) under different LED spectrums in artificial soil pot culture. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among the genotypes within each parameter. Values are expressed as mean ± SD (<span class="html-italic">n</span> = 3). L1—red70 + blue30; L2—red70 + blue20 + white10; L3—red70 + blue20 + green10; L4—red70 + blue20 + far-red10; L5—red60 + blue20 + green10 + white10; L6—red60 + blue20 + far-red10 + white10; L7—natural light.</p>
Full article ">Figure 5
<p>Tuber grading of potato grown under different LED light spectrums in artificial soil pot culture. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among the genotypes within each parameter. Values are expressed as mean ± SD (<span class="html-italic">n</span> = 3). L1—red70 + blue30; L2—red70 + blue20 + white10; L3—red70 + blue20 + green10; L4—red70 + blue20 + far-red10; L5—red60 + blue20 + green10 + white10; L6—red60 + blue20 + far-red10 + white10; L7—natural light.</p>
Full article ">Figure 6
<p>Effect of different light spectrums on GA3 content of potato plants. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) among the genotypes within each parameter. Values are expressed as mean ± SD (<span class="html-italic">n</span> = 3). L1—red70 + blue30; L2—red70 + blue20 + white10; L3—red70 + blue20 + green10; L4—red70 + blue20 + far-red10; L5—red60 + blue20 + green10 + white10; L6—red60 + blue20 + far-red10 + white10; L7—natural light.</p>
Full article ">Figure 7
<p>Principal component analysis (PCA) represents patterns and associations among the treatments. L1—red70 + blue30; L2—red70 + blue20 + white10; L3—red70 + blue20 + green10; L4—red70 + blue20 + far-red10; L5—red60 + blue20 + green10 + white10; L6—red60 + blue20 + far-red10 + white10; L7—natural light. Lines started from the ballot’s center show negative or positive correlations between distinct light treatments. (Stem L, stem length); (Stem dia., stem diameter); (leaf N, leaf number); (Leaf L, leaf length); (Leaf W, leaf width); (Branch N, branch number); (Root L, root length); (PFW, plant fresh weight); (PDW, plant dry weight); (A, photosynthetic rate); (E, transpiration rate); (gs, stomatal conductance); (WUE, water use efficiency); (Chl a, chlorophyll a); (Chl b, chlorophyll b); (Tch, total chlorophyll); (Car, carotenoid); (SPAD index); (GA3, gibberellic acid content); (Tuber N, tuber number); (Tuber FW, tuber fresh weight); (Tuber N, tuber number); (TFW, tuber fresh weight); (&lt;1 g, less than 1 g); (&gt;1 g, more than 1 g); (&gt;3 g, more than 3 g).</p>
Full article ">
19 pages, 1437 KiB  
Article
Assessment of Fertility Dynamics and Nutritional Quality of Potato Tubers in a Compost-Amended Mars Regolith Simulant
by Antonio Giandonato Caporale, Roberta Paradiso, Mario Palladino, Nafiou Arouna, Luana Izzo, Alberto Ritieni, Stefania De Pascale and Paola Adamo
Plants 2024, 13(5), 747; https://doi.org/10.3390/plants13050747 - 6 Mar 2024
Viewed by 1058
Abstract
Mars exploration will foresee the design of bioregenerative life support systems (BLSSs), in which the use/recycle of in situ resources might allow the production of food crops. However, cultivation on the poorly-fertile Mars regolith will be very challenging. To pursue this goal, we [...] Read more.
Mars exploration will foresee the design of bioregenerative life support systems (BLSSs), in which the use/recycle of in situ resources might allow the production of food crops. However, cultivation on the poorly-fertile Mars regolith will be very challenging. To pursue this goal, we grew potato (Solanum tuberosum L.) plants on the MMS-1 Mojave Mars regolith simulant, pure (R100) and mixed with green compost at 30% (R70C30), in a pot in a cold glasshouse with fertigation. For comparison purposes, we also grew plants on a fluvial sand, pure (S100) and amended with 30% of compost (S70C30), a volcanic soil (VS) and a red soil (RS). We studied the fertility dynamics in the substrates over time and the tuber nutritional quality. We investigated nutrient bioavailability and fertility indicators in the substrates and the quality of potato tubers. Plants completed the life cycle on R100 and produced scarce but nutritious tubers, despite many critical simulant properties. The compost supply enhanced the MMS-1 chemical/physical fertility and determined a higher tuber yield of better nutritional quality. This study demonstrated that a compost-amended Mars simulant could be a proper substrate to produce food crops in BLSSs, enabling it to provide similar ecosystem services of the studied terrestrial soils. Full article
Show Figures

Figure 1

Figure 1
<p>pH values (<b>A</b>) and electrical conductivity (EC) (<b>B</b>) of volcanic soil (VS), red soil (RS), fluvial sand, pure (S100) and mixed with green compost (70:30 <span class="html-italic">v</span>:<span class="html-italic">v</span>; S70C30), and Mojave Mars regolith simulant MMS-1, pure (R100) and amended with green compost (70:30 <span class="html-italic">v</span>:<span class="html-italic">v</span>; R70C30), separated (after potato plant growth) in potato tuberosphere/rhizo (RH) and bulk (BK) soils. Bars indicate mean values of 5 replicates ± standard errors. Soil (S), RH vs. BK (RB) and their interaction (S × RB) were compared by two-way ANOVA, Duncan’s multiple range test (* <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001; ns: not significant). Different lowercase letters among bars indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Concentration (mg kg<sup>−1</sup> DW) of available P extracted by 0.5M NaHCO<sub>3</sub> (buffered at pH 8.5, Olsen method) from volcanic soil (VS), red soil (RS), fluvial sand, pure (S100) and mixed with green compost (70:30 <span class="html-italic">v</span>:<span class="html-italic">v</span>; S70C30), and Mojave Mars regolith simulant MMS-1, pure (R100) and amended with green compost (70:30 <span class="html-italic">v</span>:<span class="html-italic">v</span>; R70C30), separated (after potato plant growth) in potato tuberosphere/rhizo (RH) and bulk (BK) soils. Bars indicate mean values ± standard errors (n = 5). Soil (S), RH vs. BK (RB) and their interaction (S × RB) were compared by two-way ANOVA, Duncan’s multiple range test (*** <span class="html-italic">p</span> &lt; 0.001; ns: not significant).</p>
Full article ">Figure 3
<p>Pore frequency distribution of volcanic soil (VS), red soil (RS), fluvial sand, pure (S100) and mixed with green compost (70:30 <span class="html-italic">v</span>:<span class="html-italic">v</span>; S70C30), and Mojave Mars regolith simulant MMS-1, pure (R100) and amended with green compost (70:30 <span class="html-italic">v</span>:<span class="html-italic">v</span>; R70C30) with respect to suction, derived from first derivative of respective water retention curves.</p>
Full article ">
15 pages, 2315 KiB  
Article
Effect of Light Quality on Seed Potato (Solanum tuberose L.) Tuberization When Aeroponically Grown in a Controlled Greenhouse
by Md Hafizur Rahman, Md. Jahirul Islam, Umma Habiba Mumu, Byeong Ryeol Ryu, Jung-Dae Lim, Md Obyedul Kalam Azad, Eun Ju Cheong and Young-Seok Lim
Plants 2024, 13(5), 737; https://doi.org/10.3390/plants13050737 - 6 Mar 2024
Cited by 2 | Viewed by 1483
Abstract
A plant factory equipped with artificial lights is a comparatively new concept when growing seed potatoes (Solanum tuberosum L.) for minituber production. The shortage of disease-free potato seed tubers is a key challenge to producing quality potatoes. Quality seed tuber production all [...] Read more.
A plant factory equipped with artificial lights is a comparatively new concept when growing seed potatoes (Solanum tuberosum L.) for minituber production. The shortage of disease-free potato seed tubers is a key challenge to producing quality potatoes. Quality seed tuber production all year round in a controlled environment under an artificial light condition was the main purpose of this study. The present study was conducted in a plant factory to investigate the effects of distinct spectrum compositions of LEDs on potato tuberization when grown in an aeroponic system. The study was equipped with eight LED light combinations: L1 = red: blue: green (70 + 25 + 5), L2 = red: blue: green (70 + 20 + 10), L3 = red: blue: green (70 + 15 + 15), L4 = red: blue: green (70 + 10 + 20), L5 = red: blue: far-red (70 + 25 + 5), L6 = red: blue: far-red (70 + 20 + 10), L7 = red: blue: far-red (70 + 15 + 15), L8 = red: blue: far-red (70 + 10 + 20), and L9 = natural light with 300 µmol m−2 s−1 of irradiance, 16/8 h day/night, 65% relative humidity, while natural light was used as the control treatment. According to the findings, treatment L4 recorded a higher tuber number (31/plant), tuber size (>3 g); (9.26 ± 3.01), and GA3 content, along with better plant growth characteristics. Moreover, treatment L4 recorded a significantly increased trend in the stem diameter (11.08 ± 0.25), leaf number (25.32 ± 1.2), leaf width (19 ± 0.81), root length (49 ± 2.1), and stolon length (49.62 ± 2.05) compared to the control (L9). However, the L9 treatment showed the best performance in plant fresh weight (67.16 ± 4.06 g) and plant dry weight (4.46 ± 0.08 g). In addition, photosynthetic pigments (Chl a) (0.096 ± 0.00 mg g−1, 0.093 ± 0.00 mg g−1) were found to be the highest in the L1 and L2 treatments, respectively. However, Chl b and TCL recorded the best results in treatment L4. Finally, with consideration of the plant growth and tuber yield performance, treatment L4 was found to have the best spectral composition to grow quality seed potato tubers. Full article
(This article belongs to the Special Issue Light and Its Influence on the Growth and Quality of Plants)
Show Figures

Figure 1

Figure 1
<p>Photographs of potato plants grown under different artificial LED light spectrums.</p>
Full article ">Figure 2
<p>Tuber yield of potatoes grown under different LEDs light spectra in the aeroponic culture system. Significant differences (<span class="html-italic">p</span> &lt; 0.05) are indicated by different letters in each bar graph. L1 = red: blue: green (70 + 25 + 5), L2 = red: blue: green (70 + 20 + 10), L3 = red: blue: green (70 + 15 + 15), L4 = red: blue: green (70 + 10 + 20), L5 = red: blue: far-red (70 + 25 + 5), L6 = red: blue: far-red (70 + 20 + 10), L7 = red: blue: far-red (70 + 15 + 15), L8 = red: blue: far-red (70 + 10 + 20), L9 = natural light.</p>
Full article ">Figure 3
<p>Tuber grading of potato grown under different LEDs light spectra in the aeroponic culture system. Significant differences (<span class="html-italic">p</span> &lt; 0.05) are indicated by different letters in each bar graph. L1 = red: blue: green (70 + 25 + 5), L2 = red: blue: green (70 + 20 + 10), L3 = red: blue: green (70 + 15 + 15), L4 = red: blue: green (70 + 10 + 20), L5 = red: blue: far-red (70 + 25 + 5), L6 = red: blue: far-red (70 + 20 + 10), L7 = red: blue: far-red (70 + 15 + 15), L8 = red: blue: far-red (70 + 10 + 20), L9 = natural light.</p>
Full article ">Figure 4
<p>Net photosynthetic rate (A), transpiration rate (E), stomatal conductance (gs), and water use efficiency (WUE) of potato plants grown different LEDs light spectra in an aeroponic culture system. Significant differences (<span class="html-italic">p</span> &lt; 0.05) are indicated by different letters in each bar graph. L1 = red: blue: green (70 + 25 + 5), L2 = red: blue: green (70 + 20 + 10), L3 = red: blue: green (70 + 15 + 15), L4 = red: blue: green (70 + 10 + 20), L5 = red: blue: far-red (70 + 25 + 5), L6 = red: blue: far-red (70 + 20 + 10), L7 = red: blue: far-red (70 + 15 + 15), L8 = red: blue: far-red (70 + 10 + 20), L9 = natural light.</p>
Full article ">Figure 5
<p>Chlorophyll <span class="html-italic">a</span> (Chl <span class="html-italic">a</span>), chlorophyll <span class="html-italic">b</span> (Chl <span class="html-italic">b</span>), total chlorophyll (TCL), carotenoid (car) and SPAD index of potato plants grown under different LED light spectra in an aeroponic culture system. Significant differences (<span class="html-italic">p</span> &lt; 0.05) are indicated by different letters in each bar graph. L1 = red: blue: green (70 + 25 + 5), L2 = red: blue: green (70 + 20 + 10), L3 = red: blue: green (70 + 15 + 15), L4 = red: blue: green (70 + 10 + 20), L5 = red: blue: far-red (70 + 25 + 5), L6 = red: blue: far-red (70 + 20 + 10), L7 = red: blue: far-red (70 + 15 + 15), L8 = red: blue: far-red (70 + 10 + 20), L9 = natural light.</p>
Full article ">Figure 6
<p>The GA3 content of potato plants grown in an aeroponic system. Significant differences (<span class="html-italic">p</span> &lt; 0.05) are indicated by different letters in each bar graph. L1 = red: blue: green (70 + 25 + 5), L2 = red: blue: green (70 + 20 + 10), L3 = red: blue: green (70 + 15 + 15), L4 = red: blue: green (70 + 10 + 20), L5 = red: blue: far-red (70 + 25 + 5), L6 = red: blue: far-red (70 + 20 + 10), L7 = red: blue: far-red (70 + 15 + 15), L8 = red: blue: far-red (70 + 10 + 20), L9 = natural light.</p>
Full article ">Figure 7
<p>Patterns and associations between treatments are represented by principal component analysis (PCA). Stem L. (Stem length); stem dia. (stem diameter); leaf N. (leaf number); leaf L. (leaf length); leaf W. (leaf width); branch N. (branch number); root L. (root length); stolon L. (stolon length); PFW (plant fresh weight); PDW (plant dry weight); A (photosynthetic rate); E (transpiration rate); gs (stomatal conductance); WUE (water use efficiency); Chl <span class="html-italic">a</span> (chlorophyll <span class="html-italic">a</span>); Chl <span class="html-italic">b</span> (chlorophyll <span class="html-italic">b</span>); Tch (total chlorophyll); Car (carotenoid); SPAD index; GA3 (gibberellic acid content) tuber N. (tuber number); tuber FW. (tuber fresh weight); tuber N. (tuber number); TFW (tuber fresh weight); &lt;1 g, (less than 1 g); &gt;1 g, (more than 1 g); &gt;3 g (more than 3 g).</p>
Full article ">
Back to TopTop