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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (236)

Search Parameters:
Keywords = Avena sativa

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 3430 KiB  
Article
Glauconite-Based Nanocomposites with Zn/Cu/B: Multifunctional Micronutrient Fertilizers
by Ivan Khitrin, Prokopiy Maximov, Evan Dasi, Kanipa Ibraeva, Konstantin Ponomarev, Natalia Maximova, Peter Belousov, Alexey Ruban and Maxim Rudmin
Minerals 2025, 15(1), 12; https://doi.org/10.3390/min15010012 - 26 Dec 2024
Viewed by 270
Abstract
The full potential of glauconite-based nanocomposites as micronutrient fertilizers remains underexplored, particularly their interaction with Zn, Cu, and B. Despite the promising applications, the mechanisms of nutrient sorption and their effects on plant growth require further investigation, especially concerning structural changes and nutrient [...] Read more.
The full potential of glauconite-based nanocomposites as micronutrient fertilizers remains underexplored, particularly their interaction with Zn, Cu, and B. Despite the promising applications, the mechanisms of nutrient sorption and their effects on plant growth require further investigation, especially concerning structural changes and nutrient delivery efficiency. This study investigates the modification of glauconite with Zn, Cu, and B solutions to create multifunctional nanocomposites with enhanced properties. It was established that the activation process preserves the primary globular–lamellar morphology of glauconite while introducing structural changes. Nanocomposites were synthesized using chemical activation and characterized using XRD, SEM-EDS, TEM, FTIR, and BET analyses. Agrochemical tests evaluated their effects on oat growth under controlled conditions. Nanocomposites with zinc sulfate exhibited an increase in specific surface area and mesoporosity, enhancing sorption capacity and facilitating the formation of inner-sphere complexes on the mineral’s basal surface. Modification with copper led to the formation of secondary phases, such as sulfates, on the surfaces of microflakes and globules while preserving the crystalline structure with inner-sphere coordination of Cu2+. Boron-modified nanocomposites were characterized by localized restructuring, pore channeling, and an increase in mesopore diameter, along with the formation of outer-sphere complexes relative to the basal surface of glauconite. Thermogravimetric and calorimetric analyses with mass spectrometry revealed specific endothermic and exothermic effects, particularly in Zn-modified samples, confirming changes in dehydration energetics. Agricultural tests on oats (Avena sativa) demonstrated the effectiveness of Cu- and B-modified nanocomposites in improving plant growth parameters, including a 7% increase in plant height and a 6.4% increase in dry weight. Zn-modified nanocomposites showed high germination rates (up to 100%) at low dosages but require optimization to avoid phytotoxicity at higher concentrations. The findings highlight the potential of adapting nanocomposites for targeted nutrient release. Additionally, glauconite nanocomposites have potential applications in restoring degraded soils, treating polluted runoff, and developing slow-release agrochemical systems. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>X-ray diffraction patterns of activated nanocomposites and the initial glauconite concentrate: Glt—glauconite, Qz—quartz.</p>
Full article ">Figure 2
<p>High-resolution TEM images (<b>A</b>,<b>C</b>,<b>D</b>,<b>F</b>,<b>G</b>,<b>I</b>) and SAED patterns (<b>B</b>,<b>E</b>,<b>H</b>) of activated nanocomposites: (<b>A</b>–<b>C</b>) Gk3Zn2-20and (<b>D</b>–<b>F</b>) Gk3B2-20. High-magnification TEM images (<b>D,F,I</b>) reveal structural changes in the smectite layers within the glauconite nanocomposite caused by interlayer expansion and modifications of the crystalline layers. Abbreviations: Glt—glauconite crystalline layers, Glt-Sme—smectite layers within glauconite, t—tetrahedral sheet, o—octahedral sheet, int—interlayer space.</p>
Full article ">Figure 3
<p>FTIR spectra of activated nanocomposites and the initial glauconite. Abbreviations: δ—deformation vibrations, ν—stretching vibrations.</p>
Full article ">Figure 4
<p>TG (dashed lines) and DSC curves (solid lines) of prepared nanocomposites and original glauconite.</p>
Full article ">Figure 5
<p>Selected representative curves for studied nanocomposites from mass spectrometry (MS) multi-ion detection are showcased, specifically for ions with mass-to-charge ratios (<span class="html-italic">m</span>/<span class="html-italic">z</span>) of 17 (OH<sup>+</sup>), 18 (H<sub>2</sub>O), 30 (NO<sup>+</sup>), 42 (NCO<sup>+</sup>), 44 (CO<sub>2</sub><sup>+</sup>), and (SO<sub>3</sub><sup>+</sup>).</p>
Full article ">Figure 6
<p>SEM images in a secondary electron detector showing different morphological features of the nanocomposites: (<b>A</b>) Gk3Zn2-20, (<b>B</b>) Gk3Cu2-20, (<b>C</b>) Gk3B2-20, and (<b>D</b>) original glauconite.</p>
Full article ">
28 pages, 3258 KiB  
Review
Research Progress in the Extraction, Structural Characteristics, Bioactivity, and Commercial Applications of Oat β-Glucan: A Review
by Xiaolu Li, Yicheng Wu, Ruilin Duan, Haoran Yu, Siyao Liu and Yulong Bao
Foods 2024, 13(24), 4160; https://doi.org/10.3390/foods13244160 - 22 Dec 2024
Viewed by 770
Abstract
Oats (Avena sativa L.) are an important cereal crop with diverse applications in both food and forage. Oat β-glucan has gained attention for its beneficial biological activities, such as reducing cardiovascular risk, preventing diabetes, and enhancing intestinal health. Despite its potential, more [...] Read more.
Oats (Avena sativa L.) are an important cereal crop with diverse applications in both food and forage. Oat β-glucan has gained attention for its beneficial biological activities, such as reducing cardiovascular risk, preventing diabetes, and enhancing intestinal health. Despite its potential, more comprehensive research is required to explore its preparation, modification, bioactivities, and applications. This review highlights recent advancements in the determination and preparation of oat β-glucan, explores its biological activities and mechanisms, and examines the impact of food processing techniques on its properties. This review is intended to provide a theoretical foundation and reference for the development and application of oat β-glucan in the functional food industry. Full article
Show Figures

Figure 1

Figure 1
<p>Structural representation of the oat grain presenting different oat tissues and the nutrient distribution [<a href="#B3-foods-13-04160" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>Principle of the enzymatic analysis of β-glucan [<a href="#B18-foods-13-04160" class="html-bibr">18</a>].</p>
Full article ">Figure 3
<p>Extraction and purification scheme of β-d-glucan [<a href="#B41-foods-13-04160" class="html-bibr">41</a>].</p>
Full article ">Figure 4
<p>Summary of the mechanisms of biological activities of oat β-glucan [<a href="#B65-foods-13-04160" class="html-bibr">65</a>].</p>
Full article ">Figure 5
<p>Scanning electron microscopy and FT-IR results for oat β-glucan processed using different thermal processes. BG, extracted β-glucan from whole grain oats without thermal processing; MBG, β-glucan from microwave-processed whole grain oats; SBG, extracted β-glucan from steam-processed whole grain oats [<a href="#B148-foods-13-04160" class="html-bibr">148</a>].</p>
Full article ">Figure 6
<p>Extrusion increases oat β-glucan extractability [<a href="#B151-foods-13-04160" class="html-bibr">151</a>].</p>
Full article ">Figure 7
<p>Contour plots for the effects of the addition of breadcrumbs (X1), inulin (X2), and β-glucan (X3) on (<b>a</b>) cooking yield, (<b>b</b>) moisture retention, (<b>c</b>) fat retention, and (<b>d</b>) reduction in the diameter of the low-fat beef burgers containing a canola and olive oil blend [<a href="#B170-foods-13-04160" class="html-bibr">170</a>].</p>
Full article ">
17 pages, 4164 KiB  
Article
Sterol Profile in Leaves of Spring Oats (Avena sativa L.) Under Conditions of the Cryolithozone
by Vasiliy V. Nokhsorov, Valentina E. Sofronova, Igor V. Sleptsov, Svetlana V. Senik, Lidia V. Petrova and Klim A. Petrov
Int. J. Plant Biol. 2024, 15(4), 1304-1320; https://doi.org/10.3390/ijpb15040090 - 12 Dec 2024
Viewed by 304
Abstract
Plant sterols (STs) are essential for the regulation of fluidity and permeability of cell membranes, which have a wide structural diversity. The dynamics of changes in sterol molecular species in leaves of a valuable cereal crop, spring oat (Avena sativa L.), as [...] Read more.
Plant sterols (STs) are essential for the regulation of fluidity and permeability of cell membranes, which have a wide structural diversity. The dynamics of changes in sterol molecular species in leaves of a valuable cereal crop, spring oat (Avena sativa L.), as a function of different sowing dates were studied. In particular, 11 molecular species of sterols (STs) and triterpenoids in A. sativa leaves were identified by GC-MS. Triterpenoids Ψ-taraxasterol, cyclolaudenol, and betulin were identified in A. sativa leaves for the first time, which may be related to adaptation to extreme climatic conditions of the cryolithozone. The dynamics of STs and triterpenoids changes were revealed during growth and development of the standard term and late summer sowing term during A. sativa hardening to low ambient temperatures. The ratio of β-sitosterol to campesterol was found to increase in response to low positive air temperatures, while the ratio of stigmasterol to β-sitosterol remained constant from mid-September to the end of October. Overall, leaves of standard-seeded A. sativa plants maintained higher levels of absolute STs and triterpenoids by 1.9-fold than leaves of late-seeded A. sativa plants. It is suggested that the ability of A. sativa plants to synthesize β-sitosterol and stigmasterol may be part of an evolutionary adaptation process to cope with wide temperature fluctuations and to maintain important membrane-bound metabolic processes. Full article
(This article belongs to the Section Plant Physiology)
Show Figures

Figure 1

Figure 1
<p>The air temperature course during the period of experiments.</p>
Full article ">Figure 2
<p><span class="html-italic">A. sativa</span> cultivar Vilensky, (<b>A</b>) standard sowing date (photo taken on 5 July) and (<b>B</b>) late sowing date (photo taken on 18 October).</p>
Full article ">Figure 3
<p>Representative GC-MS chromatogram of silylated free sterols and triterpens from <span class="html-italic">A. sativa</span> (4 July, leaves of the first sowing). Peaks: cholesterol (1), campesterol (2), campestanol (3), stigmasterol (4), β-sitosterol (5), sitostanol (6), Δ<sup>5</sup>-avenasterol (7), Δ<sup>7</sup>-avenasterol (8), Ψ-taraxasterol (9), cyclolaudenol (10), betulin (11).</p>
Full article ">Figure 4
<p>Mass spectra and structural formulas of the first identified triterpenoid compounds (betulin-TMS ether (<b>a</b>), cyclolaudenol-TMS ether (<b>b</b>), Ψ-taraxasterol-TMS ether (<b>c</b>)) in spring oat (<span class="html-italic">A. sativa</span>) leaves.</p>
Full article ">Figure 5
<p>Hierarchical clustering heatmap of STs and triterpenoids in the leaves of <span class="html-italic">A. sativa</span> (I sowing date, 3 June) performed on the normalized data (using the Euclidean distance matrix with Ward’s method). Colors represent different concentrations indicated by the color bar.</p>
Full article ">Figure 6
<p>Hierarchical clustering heatmap of STs and triterpenoids in leaves of <span class="html-italic">A. sativa</span> (II sowing date, 27 July) performed on the normalized data (using the Euclidean distance matrix with Ward’s method). Colors represent different concentrations indicated by the color bar.</p>
Full article ">Figure 7
<p>Dynamics of sterol C22-desaturase enzymes (<b>a</b>) and sterol molecular species ratio and activity of sterol methyltransferase (SMT) (<b>b</b>) in leaves of <span class="html-italic">A. sativa</span>. Means ± SDs. The experiments were performed with at least three biological replicates (<span class="html-italic">n</span> = 3); the different letters indicate significant differences (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 8
<p>The graph of estimations of PCA obtained on the basis of the analysis of the main components of the sterol profile of <span class="html-italic">A. sativa</span> leaves of I (blue arrow) and II crops (green arrow) in different phenological phases.</p>
Full article ">
17 pages, 2016 KiB  
Article
Different Species and Cultivars of Broad Beans, Lupins, and Clovers Demonstrated Varying Environmental Adaptability and Nitrogen Fixation Potential When Cultivated as Green Manures in Northeastern Portugal
by Peltier Aguiar, Margarida Arrobas, Ezar Alfredo Nharreluga and Manuel Ângelo Rodrigues
Sustainability 2024, 16(23), 10725; https://doi.org/10.3390/su162310725 - 6 Dec 2024
Viewed by 586
Abstract
The success of growing legumes as green manure depends on their spatial and temporal integration within agroecosystems, which minimizes competition with cash crops, and on their nitrogen (N) fixation potential. This study evaluated seven legume species for biomass production, N fixation, and suitability [...] Read more.
The success of growing legumes as green manure depends on their spatial and temporal integration within agroecosystems, which minimizes competition with cash crops, and on their nitrogen (N) fixation potential. This study evaluated seven legume species for biomass production, N fixation, and suitability for use in cropping systems in northern Portugal. Oats (Avena sativa L.) were grown to estimate the N fixation using the difference method, as a non-legume reference crop is required for this purpose, and oats are widely grown in the region. The study was conducted over four cropping cycles (2021–2024) in two climate zones across four land plots. The results indicated that the biomass production and N fixation varied by the species/cultivar and cropping cycle, which was significantly influenced by spring precipitation. Broad beans (Vicia faba L.) failed to develop in one cycle on highly acidic soil (pH 4.9), showing negative N fixation values when calculated by the difference method. Conversely, the lupins maintained a relatively high level of N fixation across all the conditions, demonstrating strong environmental adaptability. Thus, the N fixation values across the four cycles ranged from −5.4 to 419.4 kg ha−1 for broad bean (cv. Favel), while yellow lupin (Lupinus luteus L.) exhibited average values between 204.0 and 274.0 kg ha−1. The percentage of N derived from the atmosphere (%Ndfa) ranged from −13.3 to 91.6, −39.4 to 85.8, 83.8 to 94.7, 74.9 to 94.3, 72.8 to 92.2, 23.1 to 75.8, and 11.7 to 21.7 for these species/cultivars. Due to their environmental adaptability, biomass production, and N fixation capacity, these legumes could be used as green manure in inter-rows of woody crops or in summer annual crops like tomatoes and maize, grown in winter as an alternative to fallow land. The lupins showed strong promise due to their environmental resilience. Full article
(This article belongs to the Section Sustainable Agriculture)
Show Figures

Figure 1

Figure 1
<p>Climatological normal (N) and annual values (Y) of monthly temperature (T) and precipitation (P) in Mirandela (<b>left</b>) and Bragança (<b>right</b>) for the respective study periods.</p>
Full article ">Figure 2
<p>Dry matter yield (DMY) of eight species/cultivars grown under four location/year combinations [Bragança 2022 (L1/22), Mirandela 2023 (L2/23), Bragança 2023 (L3/23), and Bragança 2024 (L4/24)]. The <span class="html-italic">X</span>-axis represents Julian dates (starting from 1 January) to express the dates as a continuous variable. Error bars represent standard deviations (n = 3).</p>
Full article ">Figure 2 Cont.
<p>Dry matter yield (DMY) of eight species/cultivars grown under four location/year combinations [Bragança 2022 (L1/22), Mirandela 2023 (L2/23), Bragança 2023 (L3/23), and Bragança 2024 (L4/24)]. The <span class="html-italic">X</span>-axis represents Julian dates (starting from 1 January) to express the dates as a continuous variable. Error bars represent standard deviations (n = 3).</p>
Full article ">Figure 3
<p>Tissue nitrogen (N) concentrations in eight species/cultivars grown under four location/year combinations [Bragança 2022 (L1/22), Mirandela 2023 (L2/23), Bragança 2023 (L3/23), and Bragança 2024 (L4/24)]. The <span class="html-italic">X</span>-axis represents Julian dates (starting from 1 January) to express the dates as a continuous variable. Error bars represent standard deviations (n = 3).</p>
Full article ">Figure 4
<p>Nitrogen (N) recovery in aboveground biomass in eight species/cultivars grown under four location/year combinations [Bragança 2022 (L1/22), Mirandela 2023 (L2/23), Bragança 2023 (L3/23), and Bragança 2024 (L4/24)]. The <span class="html-italic">X</span>-axis represents Julian dates (starting from January 1) to express the dates as a continuous variable. Error bars represent standard deviations (n = 3).</p>
Full article ">Figure 4 Cont.
<p>Nitrogen (N) recovery in aboveground biomass in eight species/cultivars grown under four location/year combinations [Bragança 2022 (L1/22), Mirandela 2023 (L2/23), Bragança 2023 (L3/23), and Bragança 2024 (L4/24)]. The <span class="html-italic">X</span>-axis represents Julian dates (starting from January 1) to express the dates as a continuous variable. Error bars represent standard deviations (n = 3).</p>
Full article ">
11 pages, 1106 KiB  
Article
Changes in Morphometric and Physiological Parameters of Oat (Avena sativa L.) Caused by Fluoride Contamination
by Eugenia Krasavtseva and Dmitriy Makarov
Int. J. Plant Biol. 2024, 15(4), 1277-1287; https://doi.org/10.3390/ijpb15040088 - 4 Dec 2024
Viewed by 383
Abstract
The article presents the results of the study of the effect of fluoride on the morphometric and physiological parameters of higher plants. The test culture was the seeds of oat Avena sativa L. Phytotesting was carried out according to standard methods in eluate [...] Read more.
The article presents the results of the study of the effect of fluoride on the morphometric and physiological parameters of higher plants. The test culture was the seeds of oat Avena sativa L. Phytotesting was carried out according to standard methods in eluate and contact versions. Four different levels (0.09, 0.9, 9 and 90 mgF/L) of NaF solution for eluate phytotesting and five levels (10, 100, 500, 1000 and 2000 mgF⸱kg−1 dry soil) for contact phytotesting were applied. The decrease in root length, plant height and biomass at the maximum pollution level (90 mgF/L and 2000 mgF⸱kg−1 dry soil, respectively) relative to the control was 35.5, 23.86 and 62.47%, respectively. Statistical data processing was conducted. In addition, using a portable mini-spectrometer for leaves CI-710S, indices characterizing changes in chlorophyll content in plants were determined: Chlorophyll Content Index, Green Chlorophyll Index, Red-Edge Chlorophyll Index, Leaf Chlorophyll Index, Soil–Plant Analysis Development. The decrease in CCI, CI Green, CI Red, LCI, and SPAD indices at the maximum pollution level (2000 mgF⸱kg−1 dry soil) relative to the control was 86.2, 42.0, 57.9, 32.8 and 70.4%, respectively. Correlation analysis using the Pearson coefficient made it possible to establish a significant relationship between individual morphometric and physiological indicators. It was found that high levels of soil fluoride contamination cause significant changes in the morphometric and physiological parameters of Avena sativa L. The results of the study may have implications for agriculture or environmental protection in areas exposed to fluoride. Full article
(This article belongs to the Section Plant Response to Stresses)
Show Figures

Figure 1

Figure 1
<p>Phytoeffect provided by sodium fluoride solution of different concentrations.</p>
Full article ">Figure 2
<p>Height and aboveground biomass of plants. Letters above the columns indicate reliable presence (letters are different), absence of differences (letters are the same) between experimental variants at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Appearance of sprouts grown in control soil (<b>a</b>) and in sample V (<b>b</b>).</p>
Full article ">
24 pages, 12110 KiB  
Article
Genetic Variation Analysis of Avenin Components in the Population of F2 Common Oat Hybrids
by Anna V. Lyubimova, Dmitry I. Eremin and Evgeny P. Renev
Agronomy 2024, 14(12), 2899; https://doi.org/10.3390/agronomy14122899 - 4 Dec 2024
Viewed by 438
Abstract
The use of avenins as biochemical markers successfully complements the use of molecular markers in oat breeding. Currently, the genes controlling the synthesis of oat prolamins are insufficiently studied. The purpose of the work was to study the genetic variation of avenin components [...] Read more.
The use of avenins as biochemical markers successfully complements the use of molecular markers in oat breeding. Currently, the genes controlling the synthesis of oat prolamins are insufficiently studied. The purpose of the work was to study the genetic variation of avenin components in populations of F2 common oat hybrids and to describe new allelic variants of component blocks. The avenins component of F2 grain in 19 hybrid oat populations was studied using the native electrophoresis method. Cultivars with new combinations of avenin components were used as parental genotypes to produce hybrids. The protein separation was conducted in vertical plates of 13.2% polyacrylamide gel. The number of avenin components in the spectra of cultivars varied from 8 to 12. The observed ratio of the grain number that compose the phenotypic classes for allele pairs at each of the loci corresponded to the theoretically expected one for codominant monohybrid inheritance. Our results confirm the assumption that avenin synthesis is controlled by three independent gene clusters located on three chromosomes. In the course of the studies, hybrid combinations were not identified in the spectra of which avenin components were manifested that were absent in both parents. The prolamin component blocks in oat are formed by 2–5 components, are characterized by high stability, and are inherited unchanged. Fifteen new allelic variants of blocks of components of the avenin electrophoretic spectrum have been identified: six for the Avn A locus, six for the Avn B locus, and three for the Avn C locus. This expands the possibilities of using prolamins as biochemical markers of economically valuable oat traits and certification of new cultivars and valuable breeding lines. Full article
(This article belongs to the Section Crop Breeding and Genetics)
Show Figures

Figure 1

Figure 1
<p>Conducting hybridization. (<b>a</b>) Hybridization nursery in protected ground conditions; (<b>b</b>) oat flower with anthers.</p>
Full article ">Figure 2
<p>Electrophoretic spectra of avenins of the parent cultivars Tyumensky golozerny (P1), Megion 1 biotype (P2), and their F<sub>2</sub> hybrids. St—standard, a1–a4—electrophoregrams of F<sub>2</sub> hybrids illustrating the dose effects of avenin components controlled by the <span class="html-italic">Avn C</span> locus; 1–11—avenin component numbers.</p>
Full article ">Figure 3
<p>Electrophoretic spectra of avenins of the parent cultivars Tyumensky golozerny (P1), Hull-less (P2), and their F<sub>2</sub> hybrids. St—standard, a1–a4—electrophoregrams of F<sub>2</sub> hybrids illustrating the dose effects of avenin components controlled by the <span class="html-italic">Avn A</span> locus; 1–10—numbers of avenin components. The dashed line indicates the expected component blocks.</p>
Full article ">Figure 4
<p>Electrophoretic spectra of avenins of the parent cultivars Foma (P1), Otrada 2 biotype (P2), and their F<sub>2</sub> hybrids. St—standard, 1–14—avenin component numbers.</p>
Full article ">Figure 5
<p>Electrophoretic spectra of avenins of the parent cultivars Pobeda 2 biotype (P1), Foma (P2), and their F<sub>2</sub> hybrids. St—standard, 1–13—avenin component numbers.</p>
Full article ">Figure 6
<p>Electrophoretic spectra of avenins of the parent cultivars Otrada 2 biotype (P1), Orel (P2), and their F<sub>2</sub> hybrids. St—standard; 1–14—avenin component numbers.</p>
Full article ">Figure 7
<p>Electrophoretic spectra of avenins of parent cultivars Megion 1st biotype (P1), MF 9116-150 (P2), and their F<sub>2</sub> hybrids. H1, H2—homozygotes F<sub>2</sub>; St—standard; 1–11—numbers of avenin components.</p>
Full article ">Figure 8
<p>Electrophoretic spectra of avenins of the parent cultivars Tyumensky golozerny (P1), Megion 2nd biotype (P2), and their F<sub>2</sub> hybrids. St—standard, a1–a4—electrophoregrams of F<sub>2</sub> hybrids illustrating the dose effects of avenin components controlled by the <span class="html-italic">Avn A</span> locus; 1–14—avenin component numbers.</p>
Full article ">Figure 9
<p>Electrophoretic spectra of avenins of the parent cultivars Severyanin (1st biotype) (P1), Tyumensky golozerny (P2), and their F<sub>2</sub> hybrids. St—standard; 1–14—avenin component numbers.</p>
Full article ">Figure 10
<p>Electrophoretic spectra of avenins of the parent cultivars Tyumensky golozerny (P1), Otrada 1 biotype (P2), and their F<sub>2</sub> hybrids. St—standard; H1, H2—homozygotes F<sub>2</sub>; 1–21—numbers of avenin components.</p>
Full article ">Figure 11
<p>Electrophoretic spectra of avenins of the parent cultivars Megion 1 biotype (P1), Otrada 2 biotype (P2), and their F<sub>2</sub> hybrids. St—standard; H1—homozygote F<sub>2</sub> with <span class="html-italic">11.4.5</span> genotype formula.</p>
Full article ">Figure 12
<p>Electrophoretic spectra of avenins of the parent cultivars Foma (F), Tyumensky golozerny (Tg), Megion 1 and 2 biotype (M1, M2), and their F<sub>2</sub> hybrids. St—standard; a—spectrum of F<sub>2</sub> hybrid from crosses ♀ Foma × ♂ Tyumensky golozerny; b—spectrum of F<sub>2</sub> hybrid from crosses ♀ Foma × ♂ Megion (1st biotype); 1–20—numbers of avenin components. The formulas of avenin genotypes F<sub>2</sub> are indicated in italics.</p>
Full article ">Figure 13
<p>Electrophoretic spectra of avenins of the parent cultivars Bucephal (P1), MF 9224-164 (P2), and their F<sub>2</sub> hybrids. St—standard; 1–16—avenin component numbers.</p>
Full article ">Figure 14
<p>Electrophoretic spectra of avenins of the parent cultivars Local (P1), Hull-less (P2), and their F<sub>2</sub> hybrids. St—standard; 1–17—avenin component numbers.</p>
Full article ">Figure 15
<p>Electrophoretic spectra of avenins of the parent cultivars MF 9116-150 (P1), Otrada 1 biotype (P2), and their F<sub>2</sub> hybrids. St—standard; 1–20—avenin component numbers; the avenin formula of F<sub>2</sub> genotype is indicated in italics. The arrow marks the component that is part of the block controlled by the <span class="html-italic">Avn B</span> locus.</p>
Full article ">Figure 16
<p>Electrophoretic spectra of avenins of the parent cultivars MF 9224-164 (P1), MF 9116-150 (P2), and their F<sub>2</sub> hybrids. St—standard; 1–18—avenin component numbers; H—homozygote F<sub>2</sub> with the avenin formula <span class="html-italic">Avn A14 B6 C10</span>. The arrow marks the lower component of block B6.</p>
Full article ">Figure 17
<p>Catalogue of allelic variants of blocks of components of the avenin electrophoretic spectrum controlled by the <span class="html-italic">Avn A</span>, <span class="html-italic">Avn B</span>, and <span class="html-italic">Avn C</span> loci [<a href="#B50-agronomy-14-02899" class="html-bibr">50</a>,<a href="#B51-agronomy-14-02899" class="html-bibr">51</a>]. The new blocks identified by us are marked in red. 1–13—titles of allelic variants of blocks of components of the avenin electrophoretic spectrum.</p>
Full article ">
17 pages, 1859 KiB  
Article
Genistein and Aphid Probing Behavior: Case Studies on Polyphagous Aphid Species
by Anna Wróblewska-Kurdyk, Bożena Kordan, Katarzyna Stec, Jan Bocianowski and Beata Gabryś
Molecules 2024, 29(23), 5715; https://doi.org/10.3390/molecules29235715 - 3 Dec 2024
Viewed by 477
Abstract
(1) Background: Genistein is a naturally occurring flavonoid with a rich spectrum of biological activities, including plant-herbivore interactions. The aim of the study was to evaluate the effect of exogenous application of genistein on aphid behavior during probing in plant tissues. (2) Methods: [...] Read more.
(1) Background: Genistein is a naturally occurring flavonoid with a rich spectrum of biological activities, including plant-herbivore interactions. The aim of the study was to evaluate the effect of exogenous application of genistein on aphid behavior during probing in plant tissues. (2) Methods: Vicia faba, Brassica rapa ssp. pekinensis, and Avena sativa were treated transepidermally with a 0.1% ethanolic solution of genistein, and the probing behavior of generalist aphid species Aphis fabae, Myzus persicae, and Rhopalosiphum padi was monitored on their respective treated and untreated host plants using electropenetrography (=electrical penetration graph technique, EPG); (3) Results: Genistein did not deter aphid probing activities in non-phloem tissues. In A. fabae and R. padi, a trend towards reduction and in M. persicae a trend towards increase in phloem sap ingestion occurred on genistein-treated plants, but these trends were not statistically significant. (4) Conclusions: Genistein is not a deterrent chemical against generalist aphid species studied; therefore, it is not recommended for practical application. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of genistein [<a href="#B2-molecules-29-05715" class="html-bibr">2</a>].</p>
Full article ">Figure 2
<p>Visualization of aphid stylet activities in plant tissues recorded using electropenetrography (a sample derived from EPG recording of <span class="html-italic">Myzus persicae</span> on <span class="html-italic">Brassica rapa</span> ssp. <span class="html-italic">pekinensis</span> treated with 0.1% genistein). Upper panel illustrates a 1 h section of the 8 h EPG. Lower panels present the details of individual EPG waveforms corresponding with the display in the upper panel. ‘G’—stylets in the xylem (EPG waveform visualizes the active uptake of xylem sap); ‘C’—stylets in epidermis and mesophyll (EPG waveform visualizes the progressive stylet movements within the apoplast and occasional uptake of sap from cells adjacent to the stylet track represented as potential drops ‘pd’); ‘E1’—stylets in phloem (egestion of saliva into sieve elements); ‘E2’—stylets in phloem (passive ingestion of phloem sap from sieve elements).</p>
Full article ">Figure 3
<p>Individual variation in probing behavior of aphids on control untreated plants and plants treated transepidermally with 0.1% genistein: (<b>a</b>,<b>b</b>) <span class="html-italic">Aphis fabae</span> on <span class="html-italic">Vicia faba</span>; (<b>c</b>,<b>d</b>) <span class="html-italic">Myzus persicae</span> on <span class="html-italic">Brassica rapa</span> ssp. <span class="html-italic">pekinensis;</span> and (<b>e</b>,<b>f</b>) <span class="html-italic">Rhopalosiphum padi</span> on <span class="html-italic">Avena sativa</span>. Panels (<b>a</b>–<b>f</b>) represent the proportion of time (percentage of cumulative time for individual aphids and the mean of the group) devoted to Np—no probing, C + F + G—pathway + derailed stylet activities + xylem phase, and E—phloem phase E1 (salivation) + E2 (sap ingestion) activities recorded during the 8 h EPG experiments.</p>
Full article ">Figure 4
<p>Sequential changes in EPG-recorded stylet penetration activities of aphids on control untreated plants and plants treated transepidermally with 0.1% genistein: (<b>a</b>,<b>b</b>) <span class="html-italic">Aphis fabae</span> on <span class="html-italic">Vicia faba</span>; (<b>c</b>,<b>d</b>) <span class="html-italic">Myzus persicae</span> on <span class="html-italic">Brassica rapa</span> ssp. <span class="html-italic">pekinensis</span>; and (<b>e</b>,<b>f</b>) <span class="html-italic">Rhopalosiphum padi</span> on <span class="html-italic">Avena sativa</span>. Panels (<b>a</b>–<b>f</b>) represent the proportion of time (average percentage of cumulative time for aphids in the group) devoted to Np—no probing, C + F + G—pathway + derailed stylet activities + xylem phase, and E—phloem phase E1 (salivation) + E2 (sap ingestion) activities during the consecutive hours of 8 h EPG recording.</p>
Full article ">Figure 4 Cont.
<p>Sequential changes in EPG-recorded stylet penetration activities of aphids on control untreated plants and plants treated transepidermally with 0.1% genistein: (<b>a</b>,<b>b</b>) <span class="html-italic">Aphis fabae</span> on <span class="html-italic">Vicia faba</span>; (<b>c</b>,<b>d</b>) <span class="html-italic">Myzus persicae</span> on <span class="html-italic">Brassica rapa</span> ssp. <span class="html-italic">pekinensis</span>; and (<b>e</b>,<b>f</b>) <span class="html-italic">Rhopalosiphum padi</span> on <span class="html-italic">Avena sativa</span>. Panels (<b>a</b>–<b>f</b>) represent the proportion of time (average percentage of cumulative time for aphids in the group) devoted to Np—no probing, C + F + G—pathway + derailed stylet activities + xylem phase, and E—phloem phase E1 (salivation) + E2 (sap ingestion) activities during the consecutive hours of 8 h EPG recording.</p>
Full article ">Figure 5
<p>Cumulative percentage of aphids that attained phloem phase (=reached sieve elements) during the 8 h EPG monitoring on control untreated plants and plants treated transepidermally with 0.1% genistein: (<b>a</b>,<b>b</b>) <span class="html-italic">Aphis fabae</span> on <span class="html-italic">Vicia faba</span>; (<b>c</b>,<b>d</b>) <span class="html-italic">Myzus persicae</span> on <span class="html-italic">Brassica rapa</span> ssp. <span class="html-italic">pekinensis</span>; and (<b>e</b>,<b>f</b>) <span class="html-italic">Rhopalosiphum padi</span> on <span class="html-italic">Avena sativa</span>. E1—phloem phase salivation represents any contact with sieve elements; E2—phloem phase sap ingestion represents actual feeding, i.e., the uptake of phloem sap.</p>
Full article ">Figure 5 Cont.
<p>Cumulative percentage of aphids that attained phloem phase (=reached sieve elements) during the 8 h EPG monitoring on control untreated plants and plants treated transepidermally with 0.1% genistein: (<b>a</b>,<b>b</b>) <span class="html-italic">Aphis fabae</span> on <span class="html-italic">Vicia faba</span>; (<b>c</b>,<b>d</b>) <span class="html-italic">Myzus persicae</span> on <span class="html-italic">Brassica rapa</span> ssp. <span class="html-italic">pekinensis</span>; and (<b>e</b>,<b>f</b>) <span class="html-italic">Rhopalosiphum padi</span> on <span class="html-italic">Avena sativa</span>. E1—phloem phase salivation represents any contact with sieve elements; E2—phloem phase sap ingestion represents actual feeding, i.e., the uptake of phloem sap.</p>
Full article ">
20 pages, 7487 KiB  
Article
Genome-Wide Identification of the Lectin Receptor-like Kinase Gene Family in Avena sativa and Its Role in Salt Stress Tolerance
by Gui Xiong, Dongli Cui, Yaqi Tian, Trude Schwarzacher, John Seymour Heslop-Harrison and Qing Liu
Int. J. Mol. Sci. 2024, 25(23), 12754; https://doi.org/10.3390/ijms252312754 - 27 Nov 2024
Viewed by 608
Abstract
Lectin receptor-like kinases (LecRLKs) are membrane-bound receptor genes found in many plant species. They are involved in perceiving stresses and responding to the environment. Oat (Avena sativa; 2n = 6x = 42) are an important food and forage crop [...] Read more.
Lectin receptor-like kinases (LecRLKs) are membrane-bound receptor genes found in many plant species. They are involved in perceiving stresses and responding to the environment. Oat (Avena sativa; 2n = 6x = 42) are an important food and forage crop with potential in drought, saline, or alkaline soils. Here, we present a comprehensive genome-wide analysis of the LecRLK gene family in A. sativa and the crop’s wild relatives A. insularis (4x) and A. longiglumis (2x), unveiling a rich diversity with a total of 390 LecRLK genes identified, comprising 219 G-types, 168 L-types, and 3 C-types in oats. Genes were unevenly distributed across the oat chromosomes. GFP constructs show that family members were predominantly located in the plasma membrane. Expression under salt stress demonstrated functional redundancy and differential expression of LecRLK gene family members in oats: 173 members of this family were involved in the response to salt stress, and the expression levels of three C-type genes in the root and leaf were significantly increased under salt stress. The results show the diversity, evolutionary dynamics, and functional implications of the LecRLK gene family in A. sativa, setting a foundation for defining its roles in plant development and stress resilience, and suggesting its potential agricultural application for crop improvement. Full article
(This article belongs to the Special Issue Genetic Engineering of Plants for Stress Tolerance)
Show Figures

Figure 1

Figure 1
<p>Chromosome location of <span class="html-italic">LecRLK</span> gene family of <span class="html-italic">Avena sativa</span>; L-type (red), C-type (yellow), and G-type (blue) subfamilies are shown. Chromosome numbers are shown at the left. Center of chromosomes shows the overall gene density. <span class="html-italic">LecRLK</span> gene locations are shown on the right.</p>
Full article ">Figure 2
<p>Phylogenetic relationships of LecRLK proteins in <span class="html-italic">Avena sativa</span> and three RLK Pelle family proteins in animals. The phylogenetic trees were constructed using the maximum-likelihood method based on predicted protein sequences. L-type (red), C-type (yellow), G-type (blue), Pelle (green).</p>
Full article ">Figure 3
<p>Domain structure prediction of AsaLecRLKs showing the number of genes (right column) with each structure. G-type AsaLecRLKs contain bulb lectin domain, S-locus glycoprotein domain, and PAN domain at the N-terminus and protein kinase domain and DUF3403 domain at the C-terminus; L-type AsaLecRLKs contain the legume lectin domain at the N-terminus and protein kinase domain and adh_short domain at the C-terminus; C-type AsaLecRLK contains the calcium-binding lectin domain at the N-terminus and protein kinase domain at the C-terminus.</p>
Full article ">Figure 4
<p>Synteny analyses of <span class="html-italic">LecRLK</span> gene family between <span class="html-italic">Arabidopsis thaliana</span> and <span class="html-italic">Avena sativa</span> (<b>A</b>), <span class="html-italic">Oryza sativa</span> and <span class="html-italic">A. sativa</span> (<b>B</b>), <span class="html-italic">A. longiglumis</span> and <span class="html-italic">A. sativa</span> (<b>C</b>), and <span class="html-italic">A. insularis</span> and <span class="html-italic">A. sativa</span> (<b>D</b>). Lines represent collinear gene pairs between genomes of <span class="html-italic">A. sativa</span> and other species. Blue line: A genome; red line: C genome; green line: D genome.</p>
Full article ">Figure 5
<p>Expression profiles of <span class="html-italic">AsaLecRLK</span> genes under different levels of salt stress in root and leaf. RNA-sequencing data on salt stress for <span class="html-italic">AsaLecRLKs</span>. The heatmap was generated on the Log2 of (FPKM+1) values using TBtools. Color bar represents normalized FPKM values: red, high expression level; blue, low expression level.</p>
Full article ">Figure 6
<p>Relative expression level of <span class="html-italic">AsaLecRLK</span> genes under salt stress after 0 h, 6 h, 12 h, 24 h and 48 h in root and leaf tissue. Expression level of each gene at 0 h is set as reference. The data represent the mean values of three replicates ± SD. Statistical significance of differences was tested by one-way ANOVA analysis (<span class="html-italic">p</span> &lt; 0.05) and is indicated by lowercase letters.</p>
Full article ">Figure 7
<p>The subcellular localization of <span class="html-italic">AsaLecRLK-L-type-43</span>, <span class="html-italic">AsaLecRLK-L-type-44</span>, and <span class="html-italic">AsaLecRLK-G-type-45</span>. Bars = 25 μm. The figures show confocal images of GFP fluorescence, plasmalemma localization (mCherry), bright field, and composite field.</p>
Full article ">
19 pages, 4898 KiB  
Article
Molecular Regulation of Photosynthetic Carbon Assimilation in Oat Leaves Under Drought Stress
by Yiqun Xu, Liling Jiang, Jia Gao, Wei Zhang, Meijun Zhang, Changlai Liu and Juqing Jia
Plants 2024, 13(23), 3317; https://doi.org/10.3390/plants13233317 - 26 Nov 2024
Viewed by 492
Abstract
Common oat (Avena sativa L.) is one of the important minor grain crops in China, and drought stress severely affects its yield and quality. To investigate the drought resistance characteristics of oat seedlings, this study used Baiyan 2, an oat cultivar at [...] Read more.
Common oat (Avena sativa L.) is one of the important minor grain crops in China, and drought stress severely affects its yield and quality. To investigate the drought resistance characteristics of oat seedlings, this study used Baiyan 2, an oat cultivar at the three-leaf stage, as the experimental material. Drought stress was simulated using polyethylene glycol (PEG) to treat the seedlings. The photosynthetic parameters and physicochemical indices of the treatment groups at 6 h and 12 h were measured and compared with the control group at 0 h. The results showed that drought stress did not significantly change chlorophyll content, but it significantly reduced net photosynthetic rate and other photosynthetic parameters while significantly increasing proline content. Transcriptome analysis was conducted using seedlings from both the control and treatment groups, comparing the two treatment groups with the control group using Tbtool software (v2.136). This analysis identified 344 differentially expressed genes. Enrichment analysis of these differentially expressed genes revealed significant enrichment in physiological pathways such as photosynthesis and ion transport. Ten differentially expressed genes related to the physiological process of photosynthetic carbon assimilation were identified, all of which were downregulated. Additionally, seven differentially expressed genes were related to ion transport. Through gene co-expression analysis combined with promoter region structure analysis, 11 transcription factors (from MYB, AP2/ERF, C2C2-dof) were found to regulate the expression of 10 genes related to photosynthetic carbon assimilation. Additionally, five transcription factors regulate the expression of two malate transporter protein-related genes (from LOB, zf-HD, C2C2-Dof, etc.), five transcription factors regulate the expression of two metal ion transporter protein-related genes (from MYB, zf-HD, C2C2-Dof), five transcription factors regulate the expression of two chloride channel protein-related genes (from MYB, bZIP, AP2/ERF), and two transcription factors regulate the expression of one Annexin-related gene (from NAC, MYB). This study provides a theoretical foundation for further research on the molecular regulation of guard cells and offers a molecular basis for enhancing drought resistance in oats. Full article
(This article belongs to the Section Plant Response to Abiotic Stress and Climate Change)
Show Figures

Figure 1

Figure 1
<p>The inhibitory effects of different concentrations of PEG-6000 on the root and plant height of oats. (<b>A</b>) The germination status of oat seeds after being treated with PEG-6000 for 9 d. (<b>B</b>) Plant height of oat seedlings after treatment with four concentrations of PEG. (<b>C</b>) Root length of oat seedlings after treatment with four concentrations of PEG. Asterisks indicate statistical significance between the 0% treatment group and the 10%, 20%, and 30% treatment groups (****, <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 2
<p>Chlorophyll and proline content in oat leaves after drought treatment 0 h, 6 h and 12 h. (<b>A</b>) Chlorophyll content of oat leaves. (<b>B</b>) Proline content of oat leaves. Asterisks indicate statistical significance between the 0 h treatment group and the 6 h and 12 h treatment groups (ns, <span class="html-italic">p</span> &gt; 0.05; ****, <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 3
<p>Photosynthetic parameters of oat leaves after drought treatment 0 h,6 h and 12 h. (<b>A</b>) Net photosynthetic rate, (<b>B</b>) stomatal conductance, (<b>C</b>) transpiration rate, (<b>D</b>) intracellular carbon dioxide concentration. Asterisks indicate statistical significance between the 6 h and 12 h treatment groups (***, <span class="html-italic">p</span> &lt; 0.001; ****, <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 4
<p>The Pearson correlation of gene expression levels in 12 oat leaf samples, with baiyan2_0_1, baiyan2_0_2, and baiyan2_0_3 representing the 0 h sampling group after drought treatment. baiyan2_6_1, baiyan2_6_2, and baiyan2_6_3 were the 6 h sampling group after drought treatment, and baiyan2_12_1, baiyan2_12_2, and baiyan2_12_3 were the 12 h sampling group after drought treatment.</p>
Full article ">Figure 5
<p>Venn diagram showing the overlap of differentially expressed genes among the three different groups. baiyan2_6vsbaiyan2_0 represents the differentially expressed genes in oat seedlings under drought stress at 6 h compared to 0 h. baiyan2_12vsbaiyan2_0 represents the differentially expressed genes in oat seedlings under drought stress at 6 h compared to 0 h. baiyan2_12vsbaiyan2_6 represents the differentially expressed genes in oat seedlings under drought stress at 6 h compared to 0 h.</p>
Full article ">Figure 6
<p>The expression patterns and clustering heatmap of the 344 DEGs. (<b>A</b>) Group of expression patterns. (<b>B</b>) Heatmap of the 344 DEGs. (<b>C</b>) Number of genes in each group. (<b>D</b>) Gene expression pattern of the sample. The Z-score represents the relative expression level of a gene, which is a normalized value; red indicates high expression, while blue indicates low expression.</p>
Full article ">Figure 7
<p>Gene ontology (GO) enrichment of differentially expressed genes (DEGs). (<b>A</b>) GO enrichment of upregulated genes in oat seedlings treated with PEG-6000 for 6 h (middle group), (<b>B</b>) GO enrichment of upregulated genes in oat seedlings treated with PEG-6000 for 12 h (up group), (<b>C</b>) GO enrichment of all DEGs (344 DEGs), (<b>D</b>) GO enrichment of upregulated genes in oat seedlings treated with PEG-6000 for 0 h (down group). The legend colors represents the −log10 (<span class="html-italic">p</span>-value) of the enrichment test; the size of the circle represents the number of genes.</p>
Full article ">Figure 8
<p>KEGG enrichment of differentially expressed genes. (<b>A</b>) KEGG enrichment of 344 DEGs, (<b>B</b>) KEGG enrichment of 242 downregulated genes (down group genes), (<b>C</b>) KEGG enrichment of 73 upregulated genes (up group genes). The legend colors represents the −log10 (<span class="html-italic">p</span>-value) of the enrichment test.</p>
Full article ">Figure 9
<p>Chord diagram of differentially expressed genes associated with photosynthetic carbon fixation and ion transport and their corresponding GO terms. (<b>A</b>) Chord diagram of key genes in photosynthetic carbon fixation and their corresponding GO terms, (<b>B</b>) chord diagram of key genes in ion transport and their corresponding GO terms. Gene expression levels are represented by logFC, with logFC values of −1 for downregulated genes and 1 for upregulated genes.</p>
Full article ">Figure 10
<p>Correlation analysis between 17 key genes and 53 transcription factors. The numerical values in the small cells represent the correlation coefficients, and the colors represent the <span class="html-italic">p</span>-values of the correlation coefficient tests (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01; ***, <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 11
<p>Transcription factor and gene regulatory network: the relationship between transcription factors and genes is based on gene co-expression analysis and transcription factor binding analysis in gene promoter regions. (<b>A</b>) Network diagram of key genes involved in photosynthetic carbon assimilation. (<b>B</b>) Network diagram of key genes in ion transport.</p>
Full article ">
18 pages, 2206 KiB  
Article
Effects of Rocky Desertification Stress on Oat (Avena sativa L.) Seed Germination and Seedling Growth in the Karst Areas of Southwest China
by Haiyan Huang, Yuting Yang, Junqin Li, Yang Gao, Xiangtao Wang, Rui Wang, Zijun Zhou, Puchang Wang and Lili Zhao
Plants 2024, 13(22), 3260; https://doi.org/10.3390/plants13223260 - 20 Nov 2024
Viewed by 702
Abstract
Oat is an important crop widely distributed in temperate zones and is also commonly planted in the karst areas of southwest China. However, due to severe rocky desertification, the complex soil in this area is characterized by high calcium content, alkaline conditions, and [...] Read more.
Oat is an important crop widely distributed in temperate zones and is also commonly planted in the karst areas of southwest China. However, due to severe rocky desertification, the complex soil in this area is characterized by high calcium content, alkaline conditions, and drought, which significantly negatively impact the growth of oat seedlings. To study the adaptability of oats to rocky desertification stress at the seedling stage, we investigated the effects of CaCl2 (0–150 mM), the pH (3–9), and drought stress (PEG-6000 solution at 0 to −0.79 MPa) on seed germination and seedling growth. The results showed that (1) calcium stress had dual effects on seed germination within the range of 5–150 mM CaCl2. Low concentrations of CaCl2 (5 mM) promoted the germination potential, germination rate, germination index, and vigor index of oats, as well as the growth and biomass accumulation of radicles in oat seedlings; however, high concentrations of CaCl2 inhibited these germination parameters. (2) Under drought stress, moderate concentrations of a PEG-6000 solution significantly improved the germination potential and germination rate of oat seeds, but the germination index and vigor index decreased with an increasing PEG-6000 concentration. When the PEG-6000 concentration corresponded to −0.06 MPa, the root growth and fresh weight accumulation of oat seedlings were significantly promoted; however, as the concentration increased to −0.53 MPa and –0.79 MPa, seed germination and seedling growth were significantly inhibited. (3) pH treatments had no significant effect on oat seed germination, but all growth indexes of oats showed a downward trend under alkaline conditions. These results suggest that suitable conditions for oat planting in karst rocky desertification areas are 5 mM CaCl2, pH levels of 5–8, and drought stress between 0 and −0.32 MPa. This study provides a theoretical basis for oat introduction, cultivation, and stress-resistant breeding in this area. Full article
(This article belongs to the Special Issue Ecophysiology and Quality of Crops)
Show Figures

Figure 1

Figure 1
<p>Effects of calcium stress on oat (<span class="html-italic">Avena sativa</span> L.) seedlings: (<b>A</b>) root length; (<b>B</b>) shoot length. Different lowercase letters in the same column indicate significant differences at <span class="html-italic">p</span> &lt; 0.05; the same letter indicates no significant difference (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 2
<p>Effects of pH on oat (<span class="html-italic">Avena sativa</span> L.) seedlings: (<b>A</b>) root length; (<b>B</b>) shoot length. Different lowercase letters in the same column indicate significant differences at <span class="html-italic">p</span> &lt; 0.05; the same letter indicates no significant difference (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 3
<p>Effects of drought stress on oat (<span class="html-italic">Avena sativa</span> L.) seedlings: (<b>A</b>) root length; (<b>B</b>) shoot length. Different lowercase letters in the same column indicate significant differences at <span class="html-italic">p</span> &lt; 0.05; the same letter indicates no significant difference (<span class="html-italic">p</span> &gt; 0.05). Note: the unit of PEG-6000 (0 to −0.79) is MPa.</p>
Full article ">Figure 4
<p>Effects of different stresses on oat (<span class="html-italic">Avena sativa</span> L.) biomass and water content: (<b>A</b>) calcium stress; (<b>B</b>) drought stress; (<b>C</b>) pH. FW: fresh weight; DW: dry weight. Data are presented as the mean ± standard error. Different letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Pearson correlation analysis of the effects of calcium stress on oat (<span class="html-italic">Avena sativa</span> L.) seed germination. RL: root length; SL: shoot length; GI: germination index; GP: germination potential; VI: vigor index; GR: germination rate; FW: fresh weight; DW: dry weight; TWC: tissue water content. Note: **** indicates highly significant correlation at <span class="html-italic">p</span> ≤ 0.01; ***, ** indicates significant correlation at <span class="html-italic">p</span> ≤ 0.01, * indicates significant correlation at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 6
<p>Pearson correlation analysis of the effects of pH stress on oat (<span class="html-italic">Avena sativa</span> L.) seed germination. RL: root length; SL: shoot length; GI: germination index; GP: germination potential; VI: vigor index; GR: germination rate; FW: fresh weight; DW: dry weight; TWC: tissue water content. Note: **** indicates highly significant correlation at <span class="html-italic">p</span> ≤ 0.01; ***, ** indicates significant correlation at <span class="html-italic">p</span> ≤ 0.01, * indicates significant correlation at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 7
<p>Pearson correlation analysis of the effects of drought stress on oat (<span class="html-italic">Avena sativa</span> L.) seed germination. RL: root length; SL: shoot length; GI: germination index; GP: germination potential; VI: vigor index; GR: germination rate; FW: fresh weight; DW: dry weight; TWC: tissue water content. Note: **** indicates highly significant correlation at <span class="html-italic">p</span> ≤ 0.01; ***, ** indicates significant correlation at <span class="html-italic">p</span> ≤ 0.01, * indicates significant correlation at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">
13 pages, 991 KiB  
Article
Fatty Acid Composition, Oxidative Status, and Content of Biogenic Elements in Raw Oats Modified Through Agricultural Practices
by Michał Wojtacki, Krystyna Żuk-Gołaszewska, Robert Duliński, Joanna Giza-Gołaszewska, Barbara Kalisz and Janusz Gołaszewski
Foods 2024, 13(22), 3622; https://doi.org/10.3390/foods13223622 - 13 Nov 2024
Viewed by 598
Abstract
The chemical composition of raw oat grain is responsible for the high dietary value and health-promoting properties of oat products. This article presents the results of a study investigating the biofortification of grain in two oat genotypes—hulless and hulled—through agronomic treatments: chemical plant [...] Read more.
The chemical composition of raw oat grain is responsible for the high dietary value and health-promoting properties of oat products. This article presents the results of a study investigating the biofortification of grain in two oat genotypes—hulless and hulled—through agronomic treatments: chemical plant protection against weeds and fungi and mineral nitrogen fertilization. The applied agronomic treatments induced different changes in the fatty acid profiles, content of tocopherols, macronutrients, and micronutrients in the grain of hulled and hulless oats. Plant health contributed to higher concentrations of unsaturated fatty acids and potassium in oat grain. In turn, nitrogen fertilization decreased the content of unsaturated fatty acids, potassium, and copper and increased the content of saturated fatty acids, calcium, and manganese in oat grain. At the same time, agronomic treatments reduced the tocopherol content of the grain, which implies that the nutritional value of oats increases in the absence of chemical plant protection agents. The correlations between the content of desirable chemical compounds and agronomic treatments were stronger in hulless oat grain, which may suggest that the agronomic modification of oat-based foods is more effective in this genotype. The content of exogenous alpha-linoleic acid C18:3 n-3 and alpha-tocopherol was higher in grain harvested from the control treatment (without chemical plant protection), whereas grain harvested from fully protected treatments accumulated more essential gamma-linolenic acid C18:3 n-6. The content of gamma-tocopherol and copper in oat grain was higher in the absence of nitrogen fertilization. Full article
(This article belongs to the Section Food Nutrition)
Show Figures

Figure 1

Figure 1
<p>Groups of fatty acids in the grain of different oat morphotypes, and <span class="html-italic">oat morphotype</span> x <span class="html-italic">plant protection</span> and <span class="html-italic">oat morphotype</span> x <span class="html-italic">nitrogen fertilization</span> interactions. C—control without plant protection; H—herbicide; HF—herbicide and fungicide; N0, N60, N120—nitrogen rate of 0, 60, and 120 kg ha<sup>−1</sup>, respectively; SFAs—saturated fatty acids; MUFAs—monounsaturated fatty acids; PUFAs—polyunsaturated fatty acids.</p>
Full article ">Figure 2
<p>Interaction effects of agronomic treatments on the content of tocopherol in the grain of hulled and hulless oats. C—control without plant protection; H—herbicide; HF—herbicide and fungicide; N0, N60, N120—nitrogen rates of 0, 60, and 120 kg ha<sup>−1</sup>, respectively.</p>
Full article ">Figure 3
<p>Significant effects of (<b>A</b>) the <span class="html-italic">nitrogen fertilization</span> x <span class="html-italic">oat morphotype</span> interaction on the calcium content of oat grain and (<b>B</b>) the <span class="html-italic">plant protection</span> x <span class="html-italic">oat morphotype</span> interaction on the zinc content of oat grain. The same lower-case letters next to the bars indicate a statistically insignificant difference according to Tukey’s test.</p>
Full article ">Figure 4
<p>Principal component analysis (PCA)—the first two principal components (PC1 and PC2) describe the relationships between agronomic treatments and the chemical properties of hulless (<b>left</b>) and hulled oat (<b>right</b>) grain. Inner and outer dotted circles represent correlation coefficients of 0.7 and 1, respectively. Black circles denote agronomic treatments: {C}—control without chemical plant protection, {H}—herbicide application, {HF}—herbicide and fungicide application, {N0}—without mineral nitrogen fertilization, {N60} and {N120}—nitrogen fertilization applied at 60 and 120 kg ha<sup>−1</sup>, respectively. Blue circles—macronutrients and micronutrients. Yellow circles—saturated fatty acids. Red circles—unsaturated fatty acids. White circles—proximates. Gray circles—tocopherols.</p>
Full article ">
17 pages, 3842 KiB  
Article
Metallurgical Waste for Sustainable Agriculture: Converter Slag and Blast-Furnace Sludge Increase Oat Yield in Acidic Soils
by Olga V. Zakharova, Peter A. Baranchikov, Svetlana P. Chebotaryova, Gregory V. Grigoriev, Nataliya S. Strekalova, Tatiana A. Grodetskaya, Igor N. Burmistrov, Sergey V. Volokhov, Denis V. Kuznetsov and Alexander A. Gusev
Agronomy 2024, 14(11), 2642; https://doi.org/10.3390/agronomy14112642 - 9 Nov 2024
Viewed by 698
Abstract
The study is the first to examine the combined use of blast-furnace sludge as a source of microelements and converter slag as a soil-deoxidizing agent in oat (Avena sativa L.) cultivation in sod-podzolic soils. It has been established that blast-furnace sludge is [...] Read more.
The study is the first to examine the combined use of blast-furnace sludge as a source of microelements and converter slag as a soil-deoxidizing agent in oat (Avena sativa L.) cultivation in sod-podzolic soils. It has been established that blast-furnace sludge is a highly dispersed waste, which contains about 50% iron, 7% zinc, and a small amount of calcium, silicon, magnesium, aluminum, and sulfur. Hazardous components such as lead, arsenic, etc., are not detected. Converter slag comprises porous granules up to 3 mm in size, consisting mainly of calcium compounds (CaO, Ca(CO)3, CaSiO3, CaFe2O4) and a small amount of Mn, Al, and Mg trace elements. In a laboratory experiment, blast-furnace sludge increased the germination of oats by 5–10%, regardless of the addition of a deoxidizer (slag), but at the same time suppressed the growth of stem length by a maximum of 18% at 1 g∙kg−1. The addition of slag raised substrate pH and increased the index by 8% at a sludge concentration of 0.1 g∙kg−1. Root length in deoxidizer-free variants increased by 50–60% and with the addition of slag by 27–47%. Root dry mass also increased under the addition of sludge by 85–98%; however, the addition of slag reduced the indicator to the control level. In a field experiment with the combined application of waste, an increase in yield by more than 30% was shown. When soil was treated with slag and sludge, the height of plants increased by an average of 18%. It should be noted that the introduction of waste did not affect the quality of the grain. The use of slag increased the lead content in the soil, which is probably due to the sorption properties of calcium compounds in the slag, since lead was not found in the analyzed waste. Presumably, lead is sorbed by slag from the lower soil horizons, concentrating and immobilizing it in the upper layer. This version is supported by the absence of lead accumulation in straw and oat grain. The zinc-containing sludge increased the content of this element by 33% in the soil, as well as by 6% in straw and by 14% in grain. Thus, we found that the studied metallurgical wastes can be used as nutrients for agriculture, both individually and jointly. Overall, the proposed approach will contribute both to reducing the amount of accumulated waste and to improving the efficiency and sustainability of agricultural production and CO2 sequestration. However, the features of the accumulation of heavy metals in soil and plants under the influence of the analyzed types of waste require more in-depth study, including within the framework of long-term field experiments. Full article
(This article belongs to the Section Plant-Crop Biology and Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Experimental field.</p>
Full article ">Figure 2
<p>Waste characterization: (<b>a</b>) SEM micrograph with elemental composition and diffraction pattern of blast-furnace sludge sample; (<b>b</b>) SEM micrograph with elemental composition and diffraction pattern of converter slag.</p>
Full article ">Figure 3
<p>Morphophysiological parameters of oats: (<b>a</b>) germination; (<b>b</b>) average stem length; (<b>c</b>) average root length; (<b>d</b>) experimental plants. The * symbol marks significant differences with the control at a significance level of <span class="html-italic">p</span> &lt; 0.05. *—differences from the untreated variant, **—differences from the slag-treated variant.</p>
Full article ">Figure 4
<p>Weight of spring oat: (<b>a</b>) raw stem weight; (<b>b</b>) dry stem weight; (<b>c</b>) raw root weight; (<b>d</b>) dry root weight. The * symbol marks significant differences with the control at a significance level of <span class="html-italic">p</span> &lt; 0.05. *—differences from the untreated variant, **—differences from the slag-treated variant.</p>
Full article ">Figure 5
<p>Elemental analysis of plants: (<b>a</b>) root, control; (<b>b</b>) stem, control; (<b>c</b>) root, sludge; (<b>d</b>) stem, sludge; (<b>e</b>) root, slag + sludge; (<b>f</b>) stem, slag + sludge.</p>
Full article ">Figure 6
<p>Productivity and crop quality indicators of spring oats: (<b>a</b>) yield; (<b>b</b>) height of plants; (<b>c</b>) net photosynthesis productivity; (<b>d</b>) protein content in grain. The * symbol marks significant differences with the control at a significance level of <span class="html-italic">p</span> &lt; 0.05. *—differences from the untreated variant, **—differences from the slag-treated variant.</p>
Full article ">
19 pages, 4092 KiB  
Article
Effect of Integrated Crop–Livestock Systems on Soil Properties and Microbial Diversity in Soybean Production
by Namita Sinha, Brett R. Rushing, Aniruddha Acharya and Shankar Ganapathi Shanmugam
Appl. Biosci. 2024, 3(4), 484-502; https://doi.org/10.3390/applbiosci3040031 - 8 Nov 2024
Viewed by 611
Abstract
Integrated crop and livestock systems (ICLSs) have been considered an important management-based decision to improve soil health by carbon sequestration. A two-year study (2019–2021) at CPBES in Newton, MS, was conducted to evaluate the effect of an ICLS on soil microbial diversity in [...] Read more.
Integrated crop and livestock systems (ICLSs) have been considered an important management-based decision to improve soil health by carbon sequestration. A two-year study (2019–2021) at CPBES in Newton, MS, was conducted to evaluate the effect of an ICLS on soil microbial diversity in the southeastern region of the USA, representing agroclimatic conditions that are warm and humid. Amplicons targeting bacterial 16S rRNA genes and fungal ITS2 regions were sequenced. Taxonomic assignment and characterization of microbial diversity were performed using QIIME2®. Soil fungal diversity pattern showed significant difference (alpha diversity, p = 0.031 in 2020 and beta diversity, p = 0.037 in 2021). In contrast, no significant differences were observed in bacterial diversity. However, there were several beneficial bacterial phyla, such as Proteobacteria and Actinobacteria, and fungal phyla such as Ascomycota, which were dominant in both years and did not show significant differences due to cover crop treatments. Canonical Correspondence Analysis (CCA) and Mantel test showed significant influence on fungal diversity due to carbon (rm = 0.2581, p = 0.022), nitrogen (rm = 0.2921, p = 0.0165), and electrical conductivity (rm = 0.1836, p = 0.0583) in 2021, and on bacterial diversity due to EE-GRSP (rm = 0.22, p = 0.02) in 2020. However, the results showed that there were no significant differences between the cover crop treatments that were consistent over a two-year study period. However, the mix of different cover crops such as oats (Avena sativa L.), crimson clover (Trifolium incarnatum L.), and tillage radish (Raphanus sativus L.) demonstrated higher positive correlation and lower negative correlation with different bacterial and fungal phyla. Long term study of ICLS is suggested to understand the shift in microbiome that would help in understanding the role of cover crops and grazing in improving crop production sustainably. Full article
(This article belongs to the Special Issue Feature Papers in Applied Biosciences 2024)
Show Figures

Figure 1

Figure 1
<p>The study site and experimental design to study the effect of an ICLS on microbial diversity; (<b>a</b>) map of Newton in Mississippi; (<b>b</b>) paddock with different cover crop treatments, where O = Oats; OC = mix of oats and crimson clover; OCR = mix of oats, crimson clover, and tillage radish, Orange boxes represents different replications, and blue boxes represents the cover crop mixes in each replication; (<b>c</b>) cattle grazing on cover crops.</p>
Full article ">Figure 2
<p>Box plots showing bacterial community richness index, Chao1 in (<b>a</b>) 2020 (<span class="html-italic">p</span>-value = 0.33), (<b>b</b>) 2021 (<span class="html-italic">p</span>-value = 0.52), and Shannon diversity index in (<b>c</b>) 2020 (<span class="html-italic">p</span>-value = 0.19) and (<b>d</b>) 2021 (<span class="html-italic">p</span>-value = 0.08) across cover crop treatments. O = oats, OC = mix of oats and crimson clover, and OCR = mix of oats, crimson clover, and tillage radish. Significant differences were observed by ANOVA at the level of 0.05. The black diamond shows mean, and different colored circles are individual OTUs for cover crop treatments. Red boxplot denotes O, green boxplot represents OC, and blue boxplot represents OCR.</p>
Full article ">Figure 3
<p>Box plots showing fungal community richness index, Chao1 in (<b>a</b>) 2020 (<span class="html-italic">p</span>-value = 0.031), (<b>b</b>) 2021 (<span class="html-italic">p</span>-value = 0.43), and Shannon diversity index in (<b>c</b>) 2020 (<span class="html-italic">p</span>-value = 0.27) and (<b>d</b>) 2021 (<span class="html-italic">p</span>-value = 0.10) across cover crop treatments. O = oats, OC = mix of oats and crimson clover, and OCR = mix of oats, crimson clover, and tillage radish. Significant differences were observed by ANOVA at the level of 0.05. The black diamond shows mean, and different colored circles are individual OTUs of different cover crop treatments. Red boxplot denotes O, green boxplot represents OC, and blue boxplot represents OCR.</p>
Full article ">Figure 4
<p>Principal coordinate analysis (PCoA) showing bacterial beta diversity based on the Bray Curtis dissimilarity matrix for cover crop treatments (O = oats, OC = mix of oats and crimson clover, and OCR = mix of oats, crimson clover, and tillage radish) in (<b>a</b>) 2020, R-squared: 0.080084; <span class="html-italic">p</span>-value &lt; 0.366, and (<b>b</b>) 2021, R-squared: 0.082693; <span class="html-italic">p</span>-value &lt; 0.301. PERMANOVA was conducted at significance level of 0.05. Each dot represents a sample point with several OTUs, where purple dot is cover crop treatment, O, green dot represents OC, and yellow dot represents OCR. Axis 1 on x-axis is major axis and axis-2 on y-axis is minor axis.</p>
Full article ">Figure 5
<p>Principal coordinate analysis (PCoA) showing fungal beta diversity based on the Bray Curtis dissimilarity matrix for cover crop treatments (O = oats, OC = mix of oats and crimson clover, and OCR = mix of oats, crimson clover, and tillage radish) in (<b>a</b>) 2020, R-squared: 0.08645; <span class="html-italic">p</span>-value &lt; 0.139, and (<b>b</b>) 2021 (R-squared: 0.0986; <span class="html-italic">p</span>-value &lt; 0.037). PERMANOVA was conducted at significance level of 005. Each dot represents a sample point with several OTUs, where purple dot is cover crop treatment, O, green dot represents OC, and yellow dot represents OCR. Axis 1 on x-axis is major axis and axis-2 on y-axis is minor axis.</p>
Full article ">Figure 6
<p>Impact of cover crop treatments (O—oats, OC—mix of oats and crimson clover, and OCR—mix of oats, crimson clover, and tillage radish) on different bacterial phyla based on Pearson r correlation coefficient shown by pattern search and heat map in (<b>a</b>) 2020 and (<b>b</b>) 2021. X-axes show correlation coefficients from −1 to +1, and Y-axes show different bacterial phyla. The pink bar shows positive correlation as it is higher than 0, and the blue bar shows negative correlation as it is lower than 0. The mini heat map on right shows high and low correlation of cover crop treatments, where blue is low, yellow is medium, and red is high correlation.</p>
Full article ">Figure 7
<p>Impact of cover crop treatments (O = oats, OC = mix of oats and crimson clover, and OCR = mix of oats, crimson clover, and tillage radish) on different fungal phyla based on Pearson r correlation coefficient in (<b>a</b>) 2020 and (<b>b</b>) 2021. X-axes show correlation coefficients from −1 to +1, and Y-axes show different fungal phyla. The pink bar shows positive correlation as it is higher than 0, and the blue bar shows negative correlation as it is lower than 0. The mini heat map on the right shows high and low correlation of cover crop treatments, where blue is low, yellow is medium, and red is high correlation.</p>
Full article ">Figure 8
<p>Canonical correspondence analysis showing the correlation between different soil properties and bacterial communities across different cover crop treatments in (<b>a</b>) 2020 and (<b>b</b>) 2021. O is represented by blue circles, red triangles for OC, and dark blue squares for OCR. Green line represents different physicochemical properties such as pH, total C, total N, C:N ratio, EC, aggregate stability, and glomalin. The length of the line shows the strength of the relationship. Abbreviations: O = oats; OC = mix of oats and crimson clover; OCR = mix of oats, crimson clover, and tillage radish; EC = electrical conductivity; C = carbon; N = nitrogen; C:N = carbon–nitrogen ratio, glomalin = easily extractable glomalin-related soil protein.</p>
Full article ">Figure 9
<p>Canonical correspondence analysis shows the correlation between soil properties and bacterial abundance in (<b>a</b>) 2020 and (<b>b</b>) 2021. Points represents different bacterial phyla. Green lines represent different physicochemical properties such as pH, total C, total N, C:N ratio, EC, aggregate stability, glomalin The line length indicated the strength of variable responsible for relationship. Abbreviations: EC = electrical conductivity; C = carbon; N = nitrogen; C:N = carbon–nitrogen ratio, glomalin = easily extractable glomalin-related soil protein.</p>
Full article ">Figure 10
<p>Canonical correspondence analysis shows the correlation between soil properties and fungal communities over different cover crop treatments in (<b>a</b>) 2020 and (<b>b</b>) 2021. O is represented by a blue circle, a red triangle for OC, and a blue square for OCR. The arrow length indicates the strength of the variable responsible for the relationship. The bottom x- and y-axes are scales for sample points. The top x- and y-axes represent plotted arrows. Lines represent different physicochemical properties. Abbreviations: O = oats; OC = mix of oats and crimson clover; OCR = mix of oats, crimson clover, and tillage radish; EC = electrical conductivity; C = carbon; N = nitrogen; C:N = carbon–nitrogen ratio, EE-GRSP = easily extractable glomalin-related soil protein.</p>
Full article ">Figure 11
<p>Canonical correspondence analysis shows the correlation between different soil properties and fungal communities in (<b>a</b>) 2020 and (<b>b</b>) 2021. Blue points represent different fungal phyla. Green lines represent different physicochemical properties such as pH, total C, total N, C:N ratio, EC, aggregate stability, and glomalin. The line length indicates the strength of the variable responsible for the relationship. Blue point represents different fungal phyla. Green line represents different physicochemical. Abbreviations: EC = electrical conductivity; C = carbon; N = nitrogen; C: N = carbon–nitrogen ratio, glomalin = easily extractable glomalin-related soil protein.</p>
Full article ">
16 pages, 3711 KiB  
Article
The Optimum Mixed Cropping Ratio of Oat and Alfalfa Enhanced Plant Growth, Forage Yield, and Forage Quality in Saline Soil
by Guanglong Zhu, Jiao Liu, Hao Wu, Yiming Zhu, Nimir Eltyb Ahmed Nimir and Guisheng Zhou
Plants 2024, 13(21), 3103; https://doi.org/10.3390/plants13213103 - 4 Nov 2024
Viewed by 804
Abstract
The forage shortage is more aggravating than ever before, with husbandry development accelerating and meat and dairy product demand increasing. Salinized soils are important reserve land encouraged to be used for forage production in China. However, the salt-tolerant cultivation techniques for forage crops [...] Read more.
The forage shortage is more aggravating than ever before, with husbandry development accelerating and meat and dairy product demand increasing. Salinized soils are important reserve land encouraged to be used for forage production in China. However, the salt-tolerant cultivation techniques for forage crops are still inadequate. Therefore, a field experiment was conducted to study the effects of the mixed cropping ratio of oat and alfalfa on plant growth and physiological traits, forage yield, and forage quality in saline soils. Oat (Avena sativa L.) variety of Canadian Monopoly and alfalfa variety of WL525HQ were used, and five mixed cropping ratios (T1 = 100% oat + 0% alfalfa, CK, T2 = 75% oat + 25% alfalfa, T3 = 50% oat + 50% alfalfa, T4 = 25% oat + 75% alfalfa, and T5 = 0% oat + 100% alfalfa) were evaluated. The results showed that plant height, chlorophyll, soluble sugar, starch, antioxidant enzymes, and crude fat were increased firstly and then decreased prominently with decreased oats and increased alfalfa sowing rate; the maximum values showed under T2 but the minimum value under T5 at evaluated growth periods. On the contrary, malondialdehyde and acid detergent fiber were significantly decreased and then increased; the lowest contents were recorded under T2 and highest under T5. Furthermore, the relative growth rate, forage yield, neutral detergent fiber, and crude ash were decreased prominently with decreased oats and increased alfalfa sowing rate, and the highest and lowest values showed under T1 and T5, respectively. Oppositely, the contents of sucrose, proline, N, P, K, relative feeding value, and crude protein were all increased, with the highest contents generated under T2 and the lowest under T1. On the whole, the mixed cropping treatment of T2 showed the best performance in improving both biomass yield and forage quality by enhanced antioxidant enzyme activity, osmotic regulatory substances, and nutrient uptake and utilization. Therefore, this study indicates that 75% oat mixed cropping with 25% alfalfa can be recommended as a salt-tolerant cultivation technique for forage high-yield and high-quality production in moderately saline soil. Full article
Show Figures

Figure 1

Figure 1
<p>The effect of mixed cropping ratios of oat and alfalfa on plant height in saline soil. Different letters indicate significant differences between different treatments at the same growth stage at the <span class="html-italic">p</span> &lt; 0.05 level. (<b>A</b>), oat plant height; (<b>B</b>), alfalfa plant height.</p>
Full article ">Figure 2
<p>The effect of mixed cropping ratios of oat and alfalfa on relative growth rate in saline soil. Different letters indicate significant differences between different treatments at the same growth stage at the <span class="html-italic">p</span> &lt; 0.05 level.</p>
Full article ">Figure 3
<p>The effect of mixed cropping ratios of oat and alfalfa on fresh weight and dry weight. Different letters indicate significant differences between different treatments at the same growth stage at the <span class="html-italic">p</span> &lt; 0.05 level. (<b>A</b>), fresh forage yield; (<b>B</b>), dry forage yield.</p>
Full article ">Figure 4
<p>The effect of mixed cropping ratios of oat and alfalfa on chlorophyll content in saline soil. Different letters indicate significant differences between different treatments at the same growth stage at the <span class="html-italic">p</span> &lt; 0.05 level. ((<b>A</b>), 60 DAS; (<b>B</b>), 89 DAS; (<b>C</b>), 121 DAS).</p>
Full article ">Figure 5
<p>Effects of mixed cropping ratios of oat and alfalfa on soluble sugar, sucrose, and starch contents in saline soil. Different letters indicate significant differences between different treatments at the same growth stage at the <span class="html-italic">p</span> &lt; 0.05 level. ((<b>A</b>), soluble sugar content; (<b>B</b>), sucrose content; (<b>C</b>), starch content).</p>
Full article ">Figure 6
<p>Effects of mixed cropping ratios of oat and alfalfa on MDA content in saline soil. Different letters indicate significant differences between different treatments at the same growth stage at the <span class="html-italic">p</span> &lt; 0.05 level.</p>
Full article ">Figure 7
<p>Effects of mixed cropping ratios of oat and alfalfa on Proline (Pro) content in saline soil. Different letters indicate significant differences between different treatments at the same growth stage at the <span class="html-italic">p</span> &lt; 0.05 level.</p>
Full article ">Figure 8
<p>Effects of mixed cropping ratios of oat and alfalfa on SOD, POD, and CAT activity in saline soil. Different letters indicate significant differences between different treatments at the same growth stage at the <span class="html-italic">p</span> &lt; 0.05 level. (<b>A</b>), SOD; (<b>B</b>), POD; (<b>C</b>), CAT.</p>
Full article ">Figure 9
<p>Effect of mixed cropping ratios of oat and alfalfa on nitrogen, phosphorus, and potassium content in saline soil. A, 60 DAS; B, 89 DAS; C, 121 DAS; D, 189 DAS.</p>
Full article ">Figure 10
<p>Effect of mixed cropping ratios of oat and alfalfa on forage quality in saline soil. (<b>A</b>): RFV, relative feeding value; ADF, acid detergent fiber; aNDF, neutral detergent fiber. (<b>B</b>): CP, crude protein; ASH, crude ash; FAT, crude fat.</p>
Full article ">
16 pages, 3059 KiB  
Article
Effects of High-Density Mixed Planting in Artificial Grassland on Microbial Community
by Ziwei Tao, Jinjuan Li, Hui Li and Guozhen Du
Sustainability 2024, 16(21), 9212; https://doi.org/10.3390/su16219212 - 24 Oct 2024
Viewed by 922
Abstract
The construction level of artificial grassland is an important index of the development degree of grassland animal husbandry. Therefore, improving the productivity level of artificial grassland and promoting the sustainable utilization of artificial grassland have become important tasks that need to be urgently [...] Read more.
The construction level of artificial grassland is an important index of the development degree of grassland animal husbandry. Therefore, improving the productivity level of artificial grassland and promoting the sustainable utilization of artificial grassland have become important tasks that need to be urgently addressed. There have been numerous studies on the effects of monoculture on the soil microbial community structure in artificial grassland, but there is limited research on the effects of mixed sowing on the soil microbial community structure and the related patterns. In this study, Elymus nutans (En), Festuca sinensis (Fs), Avena sativa (As), and Poa pratensis (Pp) were used as common herbage materials in an alpine grassland pastoral area of the eastern Tibetan Plateau. Multi-density monoculture and mixed seeding were employed to establish artificial grassland communities with varying structures. By comparing the soil microbial community structure of the differently treated artificial planting grass, degraded grassland with bald spots, and natural grassland, it was confirmed that plant community diversity significantly influences the microbial community structure. The high-density planting treatment of multiple forage grasses had a more pronounced impact on the soil microbial community structure compared to that of the high-density planting treatment of a single variety of forage grass. The soil microbial community diversity index of the four mixed-planting treatments was higher than those of the other artificial grassland treatments and the natural grassland treatments, and the soil microbial community structure was most similar to that of the natural grassland. Avena sativa planting increased the abundance of Actinobacteria and Basidiomycota and decreased the number of Acidobacteria by increasing the soil pH value. The AFP (As+Fs+Pp) treatment reduced the proportion of Mortierellomycota in the soil by decreasing the content of available phosphorus. The AEFP (As+En+Fs+Pp) treatment increased the number of Proteobacteria by raising the soil total phosphorus content and reduced the abundance of Acidobacteria by lowering the soil pH value. Additionally, a machine learning method was used to evaluate the comprehensive performance of 21 artificial grassland treatments on nine soil physical and chemical properties. It was found that the AEFP mixed-planting and high-density planting treatments had the greatest improvement effect on the nine soil physical and chemical properties, which was conducive to sustainable land use. Full article
Show Figures

Figure 1

Figure 1
<p>Alpha diversity of soil bacterial community in different treatments. Note: (<b>A</b>): dilution curve; (<b>B</b>): Shannon index (diversity); (<b>C</b>): Simpson even index (evenness); (<b>D</b>): T detection index difference between groups. Different lowercase letters represent significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Alpha diversity of soil fungi community in different treatments. Note: (<b>A</b>): dilution curve; (<b>B</b>): Shannon index (diversity); (<b>C</b>): Simpson even index (evenness); (<b>D</b>): T detection index difference between groups. Different lowercase letters represent significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Relationship between soil microbial Circos samples and species. Note: This figure not only reflects composition proportion of dominant species in each sample, but also reflects distribution proportion of each dominant species in different samples. (<b>A</b>): bacterial community structure; (<b>B</b>): fungal community structure.</p>
Full article ">Figure 4
<p>An interaction network and a modular network of the soil microbial community. Note: Each connection represents a strong correlation (Spearman’s |<span class="html-italic">p</span>| &gt; 0.8). The size of each node in this figure (<b>A</b>) is a proportional to the number of connections (degrees), and the lines between the nodes are red for positive correlations and green for negative correlations. Figure (<b>B</b>) is colored by modular classification. In the picture, the text “OTUB” represents the bacterial group, and “OTUF” represents the fungal group.</p>
Full article ">Figure 5
<p>Comprehensive evaluation of different treatments.</p>
Full article ">Figure 6
<p>Correlations between the environmental factors and the dominant microbial taxa based on the Spearman correlation coefficient. Note: The color represents the value of the Spearman correlation coefficient. “B_” represents the bacterial phylum, and “F_” represents the fungal phylum. SOM: soil organic matter; AN: ammonium nitrogen; AP: available phosphorus; Apo: available potassium; TN: total nitrogen; YPh: total phosphorus; TPo: total potassium.</p>
Full article ">
Back to TopTop