[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

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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,382)

Search Parameters:
Keywords = oat

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 4283 KiB  
Article
Genome-Wide Identification of CAMTA Gene Family in Oat (Avena sativa) and Expression Analysis under Various Abiotic Stresses
by Yanjiao Yang, Jin Li, Mingjiu Yao and Shiyong Chen
Agronomy 2024, 14(9), 2053; https://doi.org/10.3390/agronomy14092053 - 8 Sep 2024
Viewed by 327
Abstract
Oat (Avena sativa) is one of the most important cereal crops and cool-season forage grasses in the world. The calmodulin-binding transcription activator (CAMTA) gene family is one of the largest families in plants, and it plays vital roles in [...] Read more.
Oat (Avena sativa) is one of the most important cereal crops and cool-season forage grasses in the world. The calmodulin-binding transcription activator (CAMTA) gene family is one of the largest families in plants, and it plays vital roles in multiple biological processes. However, the CAMTA genes in oats, especially those involved in abiotic stress, have not yet been elucidated. Herein, our findings reveal the presence of 20 distinct AsCAMTA genes, which were clustered into three subfamilies based on their gene structure and conserved motifs, indicating functional similarities within each subgroup. Chromosomal mapping indicated an uneven distribution across 10 chromosomes, suggesting a complex evolutionary history marked by potential gene duplication events. The results showed that most AsCAMTA genes contained stress-related cis-elements. The study further investigated the expression patterns of these genes under abiotic stress conditions utilizing RT-qPCR analysis. The results identified three AsCAMTA genes (AsCAMTA5, AsCAMTA7, and AsCAMTA19) that exhibited significant up-regulation under salt stress, with AsCAMTA7 also showing a marked increase in expression under drought stress. These findings suggest a pivotal role of AsCAMTA5, AsCAMTA7, and AsCAMTA19 genes in mediating the responses to various abiotic stresses by integrating multiple stress signals in oats. This investigation provides valuable insights into the potential functions of AsCAMTA genes in the stress response mechanisms of oats, laying a foundation for further functional studies aimed at enhancing abiotic stress tolerance in crops. Full article
Show Figures

Figure 1

Figure 1
<p>Phylogenetic tree, motif analysis, and gene structure of <span class="html-italic">AsCAMTA</span>: (<b>A</b>) The phylogenetic tree analysis of the AsCAMTA protein. (<b>B</b>) The motif composition of <span class="html-italic">AsCAMTA</span>. (<b>C</b>) The gene structure of the <span class="html-italic">AsCAMTA</span> genes in oats.</p>
Full article ">Figure 2
<p>Chromosomal localization of the members of the oat <span class="html-italic">CAMTA</span> gene family. Blue to yellow colors within the chromosomes indicate increased gene density. Chromosome numbers are shown at the right of the vertical bar; gene locations are shown at the left of the vertical bar.</p>
Full article ">Figure 3
<p>Synteny analysis of <span class="html-italic">CAMTAs</span> in <span class="html-italic">A. sativa</span> and three representative plants.</p>
Full article ">Figure 4
<p>Phylogenetic tree analysis of the CAMTA proteins from <span class="html-italic">A. sativa</span>, <span class="html-italic">A. thaliana</span>, <span class="html-italic">O. sativa</span>, <span class="html-italic">A. insularis</span>, and <span class="html-italic">A. longiglumis</span>. The <span class="html-italic">CAMTAs</span> were divided into three clades (Group I–III) based on the clustering of the protein sequence.</p>
Full article ">Figure 5
<p>Cis-acting elements in promoters of <span class="html-italic">AsCAMTA</span> family members.</p>
Full article ">Figure 6
<p>The expression patterns of 20 <span class="html-italic">AsCAMTA</span> genes were analyzed under salt stress by real-time quantitative RT-PCR. *, ** and *** represent significance levels of <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.001, respectively.</p>
Full article ">Figure 7
<p>The expression patterns of 20 <span class="html-italic">AsCAMTA</span> genes were analyzed under drought stress by real-time quantitative RT-PCR. * and ** represent significance levels of <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01, respectively.</p>
Full article ">
16 pages, 790 KiB  
Article
Dietary Effects of Black-Oat-Rich Polyphenols on Production Traits, Metabolic Profile, Antioxidative Status, and Carcass Quality of Fattening Lambs
by Zvonko Antunović, Željka Klir Šalavardić, Boro Mioč, Zvonimir Steiner, Mislav Đidara, Vinko Sičaja, Valentina Pavić, Lovro Mihajlović, Lidija Jakobek and Josip Novoselec
Agriculture 2024, 14(9), 1550; https://doi.org/10.3390/agriculture14091550 - 7 Sep 2024
Viewed by 278
Abstract
The study aimed to establish the dietary effects of black oat rich in polyphenols on the production traits, metabolic profile, antioxidant status, and carcass quality of fattening lambs, after weaning. In the BO group, in the feed mixture, common oats replaced the black [...] Read more.
The study aimed to establish the dietary effects of black oat rich in polyphenols on the production traits, metabolic profile, antioxidant status, and carcass quality of fattening lambs, after weaning. In the BO group, in the feed mixture, common oats replaced the black oat compared to the CO group. The research comprehensively investigated production indicators, blood metabolic profile, antioxidant status, and lamb carcass quality. No significant differences were found in the fattening or slaughter characteristics of lamb carcasses, except for lower pH1 values in BO lamb carcasses. Significant increases in RBC, HCT, and MCV levels as well as TP, ALB, and GLOB concentrations and GPx and SOD activities in the blood of BO lambs were found. The glucose and EOS content as well as the activity of the enzymes ALT and ALP were significantly lower in the blood of the BO group than in the CO group. In the liver, the DPPH activity was significantly higher in the BO lambs compared to the CO lambs. The observed changes in glucose, protein metabolism, and antioxidant status in the blood and tissues of lambs indicate that the use of polyphenol-rich black oats in the diet of lambs under stress conditions is justified. Full article
Show Figures

Figure 1

Figure 1
<p>Content of TBARS (reactive thiobarbituric acid substances) in muscle (musculus semimembranosus), liver, and kidney of lambs from the CO (control oat) and BO groups (black oat).</p>
Full article ">Figure 2
<p>DPPH (2,2-diphenyl-1-picrylhydrazyl radical) scavenging activity in muscle (musculus semimembranosus), liver, and kidney of lambs from CO (control oat) and BO groups (black oat); * <span class="html-italic">p</span>-values for DPPH-liver is 0.031.</p>
Full article ">
16 pages, 2288 KiB  
Article
Effect of Mono- and Polysaccharide on the Structure and Property of Soy Protein Isolate during Maillard Reaction
by Kun Wen, Qiyun Zhang, Jing Xie, Bin Xue, Xiaohui Li, Xiaojun Bian and Tao Sun
Foods 2024, 13(17), 2832; https://doi.org/10.3390/foods13172832 - 6 Sep 2024
Viewed by 307
Abstract
As a protein extracted from soybeans, soy protein isolate (SPI) may undergo the Maillard reaction (MR) with co-existing saccharides during the processing of soy-containing foods, potentially altering its structural and functional properties. This work aimed to investigate the effect of mono- and polysaccharides [...] Read more.
As a protein extracted from soybeans, soy protein isolate (SPI) may undergo the Maillard reaction (MR) with co-existing saccharides during the processing of soy-containing foods, potentially altering its structural and functional properties. This work aimed to investigate the effect of mono- and polysaccharides on the structure and functional properties of SPI during MR. The study found that compared to oat β-glucan, the reaction rate between SPI and D-galactose was faster, leading to a higher degree of glycosylation in the SPI–galactose conjugate. D-galactose and oat β-glucan showed different influences on the secondary structure of SPI and the microenvironment of its hydrophobic amino acids. These structural variations subsequently impact a variety of the properties of the SPI conjugates. The SPI–galactose conjugate exhibited superior solubility, surface hydrophobicity, and viscosity. Meanwhile, the SPI–galactose conjugate possessed better emulsifying stability, capability to produce foam, and stability of foam than the SPI–β-glucan conjugate. Interestingly, the SPI–β-glucan conjugate, despite its lower viscosity, showed stronger hypoglycemic activity, potentially due to the inherent activity of oat β-glucan. The SPI–galactose conjugate exhibited superior antioxidant properties due to its higher content of hydroxyl groups on its molecules. These results showed that the type of saccharides had significant influences on the SPI during MR. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>pH value, UV-Vis absorbance (<b>A</b>), and fluorescence value (<b>B</b>) of the Maillard reaction system. Different letters indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>FT-IR spectra (<b>A</b>) and secondary structure (<b>B</b>) of SPI conjugates.</p>
Full article ">Figure 3
<p>Fluorescence spectra of SPI conjugates.</p>
Full article ">Figure 4
<p>Solubility of SPI conjugates.</p>
Full article ">Figure 5
<p>The apparent viscosity of SPI conjugates.</p>
Full article ">Figure 6
<p>α-Glucosidase inhibitory activity of SPI conjugates. Different letters indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>DPPH scavenging activity of SPI conjugates. The dotted line is the concentration at which the DPPH scavenging rate reaches 50%.</p>
Full article ">Figure 8
<p>Reducing power of SPI conjugates. The dotted line is the reducing power of the samples at a concentration of 8 mg/mL.</p>
Full article ">Figure 9
<p>Ferrous ion-chelating capacity of SPI conjugates. The dotted line is the concentration at which the ferrous ion-chelating capacity reaches 50%.</p>
Full article ">
15 pages, 6781 KiB  
Article
Ultrasound-Assisted Process to Increase the Hydrophobicity of Cellulose from Oat Hulls by Surface Modification with Vegetable Oils
by Gina A. Gil-Giraldo, Janaina Mantovan, Beatriz M. Marim, João O. F. Kishima, Natália C. L. Beluci and Suzana Mali
Polysaccharides 2024, 5(3), 463-477; https://doi.org/10.3390/polysaccharides5030029 - 5 Sep 2024
Viewed by 253
Abstract
Cellulose obtained from oat hulls by bleaching with peracetic acid was modified, employing an ultrasound method that resulted in an esterification reaction with different vegetable oils (soybean, sunflower, and coconut) to produce modified cellulose (MC) with increased hydrophobicity. MC samples were characterized by [...] Read more.
Cellulose obtained from oat hulls by bleaching with peracetic acid was modified, employing an ultrasound method that resulted in an esterification reaction with different vegetable oils (soybean, sunflower, and coconut) to produce modified cellulose (MC) with increased hydrophobicity. MC samples were characterized by Fourier transform infrared spectroscopy (FTIR), X-ray diffraction, scanning electron microscopy, and their wettability and oil and water absorption capacities. FTIR indicated that the reaction occurred with all oils, which was observed by forming a new band associated with ester carbonyl groups at 1747 cm−1. The modification did not affect the crystalline structure or surface morphology of the cellulose. MC samples modified with all oil sources showed a 6 to 9-fold decrease in water absorption capacity, a 3-fold increase in oil absorption capacity, and a higher affinity for nonpolar solvents. The modified samples adsorbed lower amounts of water at a slower rate. Different oil sources did not affect the main properties of MC. The ultrasonication-assisted process was not only effective in modifying cellulose by esterification with vegetable oils but was also an eco-friendly and simple strategy that does not require toxic reagents, providing reassurance of its sustainability. Full article
(This article belongs to the Topic Polymers from Renewable Resources, 2nd Volume)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FTIR spectra of cellulose and MC samples modified with (<b>a</b>) soybean oil, (<b>b</b>) sunflower oil, and (<b>c</b>) coconut oil.</p>
Full article ">Figure 2
<p>X-ray diffractograms of cellulose and MC samples modified with (<b>a</b>) soybean oil, (<b>b</b>) sunflower oil, and (<b>c</b>) coconut oil.</p>
Full article ">Figure 2 Cont.
<p>X-ray diffractograms of cellulose and MC samples modified with (<b>a</b>) soybean oil, (<b>b</b>) sunflower oil, and (<b>c</b>) coconut oil.</p>
Full article ">Figure 3
<p>SEM images of cellulose and MC samples modified with soybean oil, sunflower oil, and coconut oil.</p>
Full article ">Figure 4
<p>Dispersions of cellulose and MC samples modified with soybean, sunflower, and coconut oils in a water/dichloromethane system.</p>
Full article ">Figure 5
<p>Moisture sorption isotherms and parameters of the GAB of cellulose and modified cellulose with (<b>a</b>) soybean, (<b>b</b>) sunflower, and (<b>c</b>) coconut oils. Lines are derived from the GAB model.</p>
Full article ">Figure 6
<p>Moisture adsorption kinetics of cellulose and MC samples modified with (<b>a</b>) soybean, (<b>b</b>) sunflower, and (<b>c</b>) coconut oils.</p>
Full article ">
15 pages, 545 KiB  
Article
Economic Efficiency versus Energy Efficiency of Selected Crops in EU Farms
by Paweł Boczar and Lucyna Błażejczyk-Majka
Resources 2024, 13(9), 123; https://doi.org/10.3390/resources13090123 - 4 Sep 2024
Viewed by 304
Abstract
The goal of farmers operating in a market economy is to maximize profit. In view of the changing political situation, the main social interest, in addition to food security, should be energy security. Here is a refined version of that sentence: This article [...] Read more.
The goal of farmers operating in a market economy is to maximize profit. In view of the changing political situation, the main social interest, in addition to food security, should be energy security. Here is a refined version of that sentence: This article examines the production efficiency of selected crops grown in the EU and how well their production can ensure both the economic security of the producers, i.e., the farmers, and Europe’s energy security. In addition, it aims to determine which costs incurred in the production process have the greatest impact on productivity. The paper uses data obtained from the Cash Crop agricultural benchmarking database, covering 19 crops and 39 cost categories for each crop. The data (averaged for 2019–2021) came from 30 farms located in 11 EU member states. The DEA method and stepwise multiple regression were used. Research has shown that crops are already being grown in Europe that provide high energy efficiency in production without compromising farm performance (including oats, peas, and winter rye). Moreover, improving the involvement of certain inputs results in improved production efficiency (e.g., through spending on agricultural consulting services). In addition, crop economic efficiency, as assessed by profit with and without subsidies, was found to be strongly correlated with production efficiency. This could indicate that subsidies do not play a key role in farm efficiency within the EU. Crop productivity remains a key factor in achieving economic and energy efficiency. The significance of the findings presented in connection with the recent COVID-19 pandemic and the escalation of the armed conflict in Ukraine has led to renewed interest in EU energy security, i.e., generating as much EU energy as possible for food and non-food production. Full article
(This article belongs to the Special Issue Assessment and Optimization of Energy Efficiency)
Show Figures

Figure 1

Figure 1
<p>Cost classification for crops adopted in the agri benchmark database. <a href="http://www.agribenchmark.org/cash-crop/publications-and-rojects0/methodology.html" target="_blank">http://www.agribenchmark.org/cash-crop/publications-and-rojects0/methodology.html</a> (accessed on 23 May 2024).</p>
Full article ">Figure 2
<p>Economic efficiency and energy efficiency in present-day Europe.</p>
Full article ">
15 pages, 8318 KiB  
Article
Exogenous Melatonin Alleviates Osmotic Stress by Enhancing Antioxidant Metabolism, Photosynthetic Maintenance, and Hormone Homeostasis in Forage Oat (Avena sativa) Seedlings
by Jingbo Yu, Xingyu Luo, Qingping Zhou, Zhou Li and Shiyong Chen
Grasses 2024, 3(3), 190-204; https://doi.org/10.3390/grasses3030014 - 3 Sep 2024
Viewed by 445
Abstract
Melatonin (MT) is a multifunctional hormone that enhances crop resilience against various abiotic stresses. However, its regulatory mechanism of osmotic tolerance in forage oats (Avena sativa) plants under water-limited scenarios is still unclear. This study aimed to delineate the impact of [...] Read more.
Melatonin (MT) is a multifunctional hormone that enhances crop resilience against various abiotic stresses. However, its regulatory mechanism of osmotic tolerance in forage oats (Avena sativa) plants under water-limited scenarios is still unclear. This study aimed to delineate the impact of MT pretreatment on the morphological, physiological, and biochemical functions of oat seedlings under osmotic stress. Our findings demonstrated that exogenous treatment of MT noticeably elevated leaf area while decreasing the root/shoot ratio of oat seedlings subjected to osmotic stress. Osmotic-induced 38.22% or 48.37% decrease in relative water content could be significantly alleviated by MT pretreatment on day 7 or day 14, respectively. MT treatment also significantly mitigated osmotic-induced decreases in photosynthetic parameters including net photosynthetic rate, stomatic conductance, and intercellular CO2 concentration as well as various chlorophyll fluorescence parameters, which could contribute to enhanced accumulations of free proline and soluble sugars in seedlings after being subjected to a prolonged duration of osmotic stress. Furthermore, MT markedly improved antioxidant enzyme activities including superoxide dismutase, ascorbate peroxidase, catalase, and peroxidase along with the accumulation of ascorbic acid contributing to a significant reduction in reactive oxygen species under osmotic stress. In addition, the MT application induced a 978.12%, 33.54%, or 30.59% increase in endogenous MT, indole acetic acid, or gibberellic acid content under osmotic stress but did not affect the accumulation of abscisic acid. These findings suggest that an optimal concentration of MT (100 μmol·L−1) could relieve osmotic stress via improvement in osmotic adjustment, the enzymatic antioxidant defense system, and endogenous hormonal balance, thereby contributing to enhanced photosynthetic functions and growth of oat seedlings under water-limited conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of exogenous melatonin (MT) on (<b>A</b>) relative water content, (<b>B</b>) free proline content, and (<b>C</b>) soluble sugar content in leaves of oat seedlings under osmotic stress. Different letters above columns show significant difference by LSD (<span class="html-italic">p</span> &lt; 0.05) on given day. Vertical bars indicate ± SE of mean (n = 4). C, control; D, osmotic stress; D + MT, osmotic-stressed seedlings were pretreated with 100 μmol·L<sup>−1</sup> of MT.</p>
Full article ">Figure 2
<p>Effects of exogenous melatonin (MT) on (<b>A</b>) seedling height, (<b>B</b>) leaf area, and (<b>C</b>) root/shoot ratio in oat seedlings under osmotic stress. Different letters above columns show significant difference by LSD (<span class="html-italic">p</span> &lt; 0.05) on given day. Vertical bars indicate ±SE of mean (n = 4). C, control; D, osmotic stress; D + MT, osmotic-stressed seedlings were pretreated with 100 μmol·L<sup>−1</sup> of MT.</p>
Full article ">Figure 3
<p>Effects of exogenous melatonin (MT) on (<b>A</b>) hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>), (<b>B</b>) superoxide anion (O<sub>2</sub><sup>−</sup>), (<b>C</b>) relative conductivity, and (<b>D</b>) malondialdehyde content in leaves of oat seedlings under osmotic stress. Different letters above columns show significant difference by LSD (<span class="html-italic">p</span> &lt; 0.05) on given day. Vertical bars indicate ±SE of mean (n = 4). C, control; D, osmotic stress; D + MT, osmotic-stressed seedlings were pretreated with 100 μmol·L<sup>−1</sup> of MT.</p>
Full article ">Figure 4
<p>Effects of exogenous melatonin (MT) on (<b>A</b>) superoxide dismutase (SOD), (<b>B</b>) catalase (CAT), (<b>C</b>) peroxidase (POD), and (<b>D</b>) ascorbate peroxidase (APX) activities in leaves of oat seedlings under osmotic stress. Different letters above columns show significant difference by LSD (<span class="html-italic">p</span> &lt; 0.05) on given day. Vertical bars indicate ±SE of mean (n = 4). C, control; D, osmotic stress; D + MT, osmotic-stressed seedlings were pretreated with 100 μmol·L<sup>−1</sup> of MT.</p>
Full article ">Figure 5
<p>Effects of exogenous melatonin (MT) on (<b>A</b>) ascorbic acid (ASA) and (<b>B</b>) glutathione (GSH) contents in leaves of oat seedlings under osmotic stress. Different letters above columns show significant difference by LSD (<span class="html-italic">p</span> &lt; 0.05) on given day. Vertical bars indicate ±SE of mean (n = 4). C, control; D, osmotic stress; D + MT, osmotic-stressed seedlings were pretreated with 100 μmol·L<sup>−1</sup> of MT.</p>
Full article ">Figure 6
<p>Effects of exogenous melatonin (MT) on (<b>A</b>) total chlorophyll content, (<b>B</b>) chlorophyll a, (<b>C</b>) chlorophyll b, and (<b>D</b>) carotenoid in leaves of oat seedlings under osmotic stress. Different letters above columns show significant difference by LSD (<span class="html-italic">p</span> &lt; 0.05) on given day. Vertical bars indicate ±SE of mean (n = 4). C, control; D, osmotic stress; D + MT, osmotic-stressed seedlings were pretreated with 100 μmol·L<sup>−1</sup> of MT.</p>
Full article ">Figure 7
<p>Effects of exogenous melatonin (MT) on (<b>A</b>) photosynthesis rate (Pn), (<b>B</b>) stomatal conductance (Gs), (<b>C</b>) intercellular CO<sub>2</sub> concentration (Ci), and (<b>D</b>) transpiration rate (Tr) in leaves of oat seedlings under osmotic stress. Vertical bars indicate ±SE of the mean (n = 4). C, control; D, osmotic stress; D + MT, osmotic-stressed seedlings were pretreated with 100 μmol·L<sup>−1</sup> of MT.</p>
Full article ">Figure 8
<p>Effects of exogenous melatonin (MT) on (<b>A</b>) optimal/maximal PSII efficiency (Fv/Fm), (<b>B</b>) electron transport rate (ETR), (<b>C</b>) photochemical quenching (qP), (<b>D</b>) non-photochemical quenching (qN), (<b>E</b>) photochemical efficiency of PSII (Fv’/Fm’), and (<b>F</b>) actual photochemical efficiency (ΦPSII) in leaves of oat seedlings under osmotic stress. Different letters above columns show significant difference by LSD (<span class="html-italic">p</span> &lt; 0.05) on given day. Vertical bars indicate ±SE of mean (n = 4). C, control; D, osmotic stress; D + MT, osmotic-stressed seedlings were pretreated with 100 μmol·L<sup>−1</sup> of MT.</p>
Full article ">Figure 9
<p>Effects of exogenous melatonin (MT) on (<b>A</b>) MT content, (<b>B</b>) indole-3-acetic acid (IAA) content, (<b>C</b>) abscisic acid (ABA) content, and (<b>D</b>) gibberellin (GA<sub>3</sub>) content in leaves of oat seedlings under osmotic stress. Different letters above columns show significant difference by LSD (<span class="html-italic">p</span> &lt; 0.05) on given day. Vertical bars indicate ±SE of mean (n = 4). C, control; D, osmotic stress; D + MT, osmotic-stressed seedlings were pretreated with 100 μmol·L<sup>−1</sup> of MT.</p>
Full article ">
15 pages, 2697 KiB  
Article
Nutrient Dynamics in Integrated Crop–Livestock Systems: Effects of Stocking Rates and Nitrogen System Fertilization on Litter Decomposition and Release
by Marcos Antonio de Bortolli, Tangriani Simioni Assmann, Betania Brum de Bortolli, Marcieli Maccari, Angela Bernardon, Jorge Jamhour, Alan J. Franzluebbers, Andre Brugnara Soares and Igor Kieling Severo
Agronomy 2024, 14(9), 2009; https://doi.org/10.3390/agronomy14092009 - 3 Sep 2024
Viewed by 374
Abstract
Current fertilizer recommendations often neglect nutrient cycling across crop rotations. This study aimed to assess the decay rate and nutrient (N, P, K) release patterns of sorghum, black oat, and corn residues (omitido) in an integrated crop–livestock system. The experiment used factorial treatments [...] Read more.
Current fertilizer recommendations often neglect nutrient cycling across crop rotations. This study aimed to assess the decay rate and nutrient (N, P, K) release patterns of sorghum, black oat, and corn residues (omitido) in an integrated crop–livestock system. The experiment used factorial treatments based on two sward heights (high and low) and two nitrogen fertilization levels (N-pasture at 200 kg N ha−1 and N-corn at 0 kg N ha−1). Litter bags were collected at various intervals from each crop to measure nutrient release patterns and decomposition rates. The results showed that pasture height and nitrogen fertilization significantly influenced decomposition and nutrient release, affecting the subsequent grain crop phase. Potassium was released rapidly and in high amounts. Nitrogen fertilization during the pasture phase prevented nitrogen and phosphorus immobilization in black oat residue and reduced immobilization in corn residue. These findings highlight the importance of accounting for nutrient cycling and decomposition rates in fertilization strategies to enhance the sustainability of integrated crop–livestock systems. Full article
(This article belongs to the Special Issue Integrated Nutrient Management for Farming Sustainability)
Show Figures

Figure 1

Figure 1
<p><b>Remaining</b> dry matter (% and Mg DM ha<sup>−1</sup>) of sorghum, black oat, and corn residue during litter-bag exposure in the field as affected by N-fertilization time (N-Fert time), sward height, and days after deposition. <b>NCHH:</b> N-corn fertilization–high sward height; NCLH: N-corn fertilization–low sward height; NPHH: N-pasture fertilization–high sward height; NPLH: N-pasture fertilization–low sward height.</p>
Full article ">Figure 2
<p>Remaining nitrogen (%) and released N (kg N ha<sup>−1</sup>) of sorghum, black oat, and corn residue during litter-bag exposure in the field as affected by N-fertilization time, sward height, and days after deposition. NCHH: N-corn fertilization–high sward height; NCLH: N-corn fertilization–low sward height; NPHH: N-pasture fertilization–high sward height; NPLH: N-pasture fertilization–low sward height.</p>
Full article ">Figure 3
<p>Remaining phosphorus (%) and released P (kg P ha<sup>−1</sup>) of sorghum, black oat, and corn residue during litter-bag exposure in the field as affected by N-fertilization time, sward height, and days after deposition. NCHH: N-corn fertilization–high sward height; NCLH: N-corn fertilization–low sward height; NPHH: N-pasture fertilization–high sward height; NPLH: N-pasture fertilization–low sward height.</p>
Full article ">Figure 4
<p>Remaining potassium (%) and released K (kg K ha<sup>−1</sup>) of sorghum, black oat, and corn residue during litter-bag exposure in the field as affected by N-fertilization time, sward height, and days after deposition. NCHH: N-corn fertilization–high sward height; NCLH: N-corn fertilization–low sward height; NPHH: N-pasture fertilization–high sward height; NPLH: N-pasture fertilization–low sward height.</p>
Full article ">Figure 5
<p>Single model of remaining potassium of sorghum, black oat, and corn residue during litter-bag exposure in the field. Each point is the average of three observations, <span class="html-italic">n</span> = 27.</p>
Full article ">
18 pages, 4735 KiB  
Article
How Do Off-Season Cover Crops Affect Soybean Weed Communities?
by Eduarda Grün, Alexandre Ferigolo Alves, Anelise Lencina da Silva, Alencar Junior Zanon, Arícia Ritter Corrêa, Eduard Mroginski Leichtweis, Roberto Costa Avila Neto and André da Rosa Ulguim
Agriculture 2024, 14(9), 1509; https://doi.org/10.3390/agriculture14091509 - 2 Sep 2024
Viewed by 393
Abstract
Weeds compete for environmental resources, leading to reduced soybean yield. In this context, integrated weed management strategies related to cultural control with the use of cover crops are necessary. Our aim was to evaluate weed occurrence in soybean systems with different cover crops. [...] Read more.
Weeds compete for environmental resources, leading to reduced soybean yield. In this context, integrated weed management strategies related to cultural control with the use of cover crops are necessary. Our aim was to evaluate weed occurrence in soybean systems with different cover crops. Field studies were conducted at Júlio de Castilhos, Santa Maria, Capão do Leão, Barra do Ribeiro, and Santo Ângelo, Rio Grande do Sul, Brazil. A randomized complete block design with four replications was used. Treatments consisted of black oats (Avena strigosa Schreb.), white oats (Avena sativa L.), rye (Secale cereale L.), hairy vetch (Vicia sativa L.), forage turnip (Raphanus sativus L.), and white clover (Trifolium repens L.) in pure stands or in mixtures. The analyzed variables were relative frequency, density, abundance, and importance value index, similarity index of weeds, dry shoot mass of cover crop, and soybean yield. Cover crops containing white or black oats reduced the relative importance value index of weeds, such as Lollium multiforum, Conyza spp., and Bidens pilosa. Forage turnip, hairy vetch, and white clover showed distinct responses. Black oats and forage turnip did not differ from cover crop mixtures in terms of dry shoot mass and grain yield, being superior to fallow, white clover, and hairy vetch. Full article
Show Figures

Figure 1

Figure 1
<p>Phytosociological analysis of weeds using the importance value index (IVIR) during cover crop development in Júlio de Castilhos in the 2020–2021 season. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 2
<p>Phytosociological analysis of weeds using the importance value index (IVIR) at the V4 stage of soybeans in Júlio de Castilhos in the 2020–2021 season. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 3
<p>Phytosociological analysis of weeds using the importance value index (IVIR) during cover crop development in Santa Maria in the 2020–2021 season. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>Phytosociological analysis of weeds using the importance value index (IVIR) at the V4 stage of soybeans in Santa Maria in the 2020–2021 season. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 5
<p>Phytosociological analysis of weeds using the importance value index (IVIR) during cover crop development in Capão do Leão in the 2020–2021 season. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 6
<p>Cluster analysis for similarity between weeds observed at different sites (JC—Júlio de Castilhos; CL—Capão do Leão; SM—Santa Maria) for treatments with cover crops used in the 2020–2021 harvest. Cophenetic coefficient of 0.93. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 7
<p>Phytosociological analysis of weeds using the importance value index (IVIR) during cover crop development in Barra in the 2021–2022 season. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 8
<p>Phytosociological analysis of weeds using the importance value index (IVIR) during cover crop development in Santa Maria in the 2021–2022 season. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 9
<p>Phytosociological analysis of weeds using the importance value index (IVIR) during cover crop development in Santo Ângelo in the 2021–2022 season. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 10
<p>Cluster analysis for similarity between weeds observed at different sites (BR—Barra do Ribeiro; SA—Santo Ângelo; SM—Santa Maria) for treatments with cover crops used in the 2021–2022 harvest. Cophenetic coefficient of 0.97. The composition of the treatments is provided in <a href="#agriculture-14-01509-t001" class="html-table">Table 1</a>.</p>
Full article ">
22 pages, 1748 KiB  
Article
Influence of Bacterial Fertilizers on the Structure of the Rhizospheric Fungal Community of Cereals South of Western Siberia
by Natalia Nikolaevna Shuliko, Olga Valentinovna Selitskaya, Elena Vasilyevna Tukmacheva, Alina Andreevna Kiselyova, Irina Anatolyevna Korchagina, Ekaterina Vladimirovna Kubasova and Artem Yuryevich Timokhin
Agronomy 2024, 14(9), 1989; https://doi.org/10.3390/agronomy14091989 - 2 Sep 2024
Viewed by 336
Abstract
The general lack of knowledge on the conditions of Western Siberia (Omsk region) and the taxonomic diversity of zonal soils determines the relevance of these studies. The research was carried out in order to study the effect of complex biologics on the taxonomic [...] Read more.
The general lack of knowledge on the conditions of Western Siberia (Omsk region) and the taxonomic diversity of zonal soils determines the relevance of these studies. The research was carried out in order to study the effect of complex biologics on the taxonomic diversity of the fungal component of the microbiome of the rhizosphere of cereals and the phytosanitary condition of crops in the southern forest-steppe (meadow-chernozem soil) and subtaiga (gray forest soil) zones of the Omsk Irtysh region (Western Siberia). This work was carried out in 2022–2023, using laboratory studies in combination with field experiments and metagenomic and statistical analyses. The objects of research were varieties of cereals and grain forage crops of Omsk selection: soil microorganisms. The scheme of the experiment involved the study of the following options: varieties of cereals (factor A): spring soft wheat—Omsk 42, Omsk 44, Tarskaya 12; durum wheat—Omsk coral; barley—Omsk 101; oats—Siberian hercules; bacterial preparation for seed inoculation (factor B) without the drug—Mizorin and Flavobacterin. The sampling of the plant rhizosphere for metagenomic analysis was carried out during the earing phase (July). For the first time, the taxonomic composition of the fungal community was determined based on the analysis of amplicon libraries of fragments of ribosomal operons of ITS2 fungi during colonization of crop roots by nitrogen-fixing bacteria in various soil and climatic zones of the Omsk region. The fungal component of the microbiome was analyzed in two zones of the Omsk region (southern forest-steppe and subtaiga). The five dominant phyla of soil fungi were located in the following decreasing series: Ascomycota (about 70%) > Mortierellomycota (about 7%) > Basidiomycota (about 5%) > Mucoromycota (3%) > Chytridiomycota (1%). The five main genera of fungi inhabiting the rhizosphere of cereals are located in a decreasing row: Giberella (6.9%) > Mortierella (6.6%) > Chaetomium (4.8%) > Cladosporium (3.8%) > Rhizopus (3.3%). The predominantly positive effect of biologics of associative nitrogen fixation on the fungal community of the soil (rhizosphere) of experimental sites located in different soil and climatic zones has been established. During seed bacterization, the growth of saprotrophic fungal genera was noted in relation to the control variants Pseudogymnoascus, Chloridium, Clonostachys, Trihoderma, etc., and the fungicidal properties of bacterial strains introduced into the soil were actively manifested relative to phytopathogenic fungi of the genera Alternaria, Blumeria, Fusarium, etc. According to the results of determining the number of infectious structures of Rhizoctonia solani, it was found that the population of the soil with viable cells of this pathogen was 1–3 pcs/g (below the threshold of harmfulness, PV 20 pcs/g of soil), which indicates a favorable phytosanitary situation with respect to the pathogen. The fungicidal effect of the applied bacterial fertilizers on Rhizoctonia solani could not be detected. The number of Bipolaris sorokiniana varied depending on the drug used. In the conditions of the southern forest-steppe zone of the Omsk region (meadow-chernozem soil), the greatest fungicidal effect was noted in Flavobacterin application variants on wheat of the Omsk 42 variety, durum wheat of the Omsk coral variety, and barley; the decrease in conidia relative to the control was 73, 35, and 29%, respectively. In the subtaiga zone of the Omsk Irtysh region (gray forest soil), as in the southern forest-steppe zone, pre-sowing bacterization of seeds with Flavobacterin led to a decrease in Bipolaris sorokiniana in the rhizosphere of wheat of the Omsk 42 variety by 18%, and oats by 27%, to control. The use of the drug Mizorin in some variants of the experiment led to an insignificant decrease in the harmful fungus or had no effect at all. Full article
(This article belongs to the Section Agricultural Biosystem and Biological Engineering)
Show Figures

Figure 1

Figure 1
<p>Composition and representatives of the dominant phyla in the microbiome of the rhizosphere of grain crops (meadow-chernozemic and gray forest soil) in 2023.</p>
Full article ">Figure 2
<p>Taxonomic structure (at the phylum level) of eukaryotic communities during inoculation in the conditions of the southern forest-steppe zone in 2023.</p>
Full article ">Figure 3
<p>Taxonomic structure (at the phylum level) of eukaryotic communities during inoculation in the conditions of the subtaiga zone in 2023.</p>
Full article ">Figure 4
<p>Composition and representatives of the dominant genera in the microbiome of the rhizosphere of grain crops (meadow-chernozemic and gray forest soil) in 2023.</p>
Full article ">Figure 5
<p>Taxonomic structure (at the genus level) of eukaryotic communities during inoculation in the conditions of the southern forest-steppe zone in 2023.</p>
Full article ">Figure 6
<p>Taxonomic structure (at the genus level) of eukaryotic communities during inoculation in the conditions of the subtaiga zone (%) in 2023.</p>
Full article ">Figure 7
<p>Soil population with <span class="html-italic">Rhizoctonia solani</span> micromycete (before sowing) (n = 3).</p>
Full article ">
17 pages, 5739 KiB  
Article
Oribatid Mites in a Humid Mediterranean Environment under Different Soil Uses and Fertilization Management
by Àngela D. Bosch-Serra, Jordi Orobitg, Martina Badia-Cardet, Jennifer L. Veenstra and Bernat Perramon
Diversity 2024, 16(9), 533; https://doi.org/10.3390/d16090533 - 1 Sep 2024
Viewed by 259
Abstract
Measuring soil quality and the use of indicators for its evaluation is a worldwide challenge. In Garrotxa Volcanic Zone Natural Park (northeastern Spain), different parameters related to oribatid mites as indicators of soil quality were evaluated under different land uses: forest, pasture, and [...] Read more.
Measuring soil quality and the use of indicators for its evaluation is a worldwide challenge. In Garrotxa Volcanic Zone Natural Park (northeastern Spain), different parameters related to oribatid mites as indicators of soil quality were evaluated under different land uses: forest, pasture, and a biennial double-crop rotation of forage crops. In forage crops, previous fertilization management (one based on mineral fertilizers, three on cattle manure, and one using both types) was also evaluated. Three samplings (April, June, and September) were performed over one season. Fifty-four oribatid species belonging to 28 families were identified. Abundance was the lowest in June for all land uses (average of 1184 individuals m−2). In the study period, abundance, diversity (Shannon index, H’), and dominance (Berger–Parker index, d) varied with different land uses, with the highest values of abundance and H’ in forests (9287 individuals m−2 and 2.19, respectively) and the lowest dominance in forests (d = 0.29) without differences between the other uses. Additionally, in the studied parameters, no differences were associated with previous fertilization management in forage crops. Hypochthoniella minutissima, Xenillus (X.) tegeocranus characterized the forest system, Epilohmannia cylindrica minima the forage crops, and Tectocepheus sarekensis the pasture. In pasture, the dominance of the parthenogenetic species Tectocepheus sarekensis raises concerns about potential management constraints. Full article
(This article belongs to the Special Issue Diversity and Ecology of the Acari)
Show Figures

Figure 1

Figure 1
<p>Monthly rainfall and mean monthly temperature at the experimental site during the agronomic experimental year.</p>
Full article ">Figure 2
<p>Location map of the studied areas devoted to the three different land uses (forest, pasture, and forage crops).</p>
Full article ">Figure 3
<p>Crop reference evapotranspiration (ET<sub>o</sub>) during the study period and soil moisture (%, <span class="html-italic">w</span>/<span class="html-italic">w</span>) for the three sampling dates in April, June, and September, and for the three land uses: forest, pasture, and forage crops. For each sampling, soil moisture values with different letters are significantly different according to Duncan’s multiple range test (α = 0.05).</p>
Full article ">Figure 4
<p>Mean abundance (<b>a</b>), Shannon index of diversity (<b>b</b>), and Berger–Parker index of dominance (<b>c</b>) for the three sampling months and different soil uses: forage crops (under the 250 CM fertilization management), pasture, and forest. Mean values for months and uses with different letters are significantly different (least squares means <span class="html-italic">p</span> &lt; 0.05): “A” or “B” for months, and “a” or “b” for uses.</p>
Full article ">Figure 5
<p>Ordination diagram from a principal component analysis for the oribatid mite community. The most relevant species are labeled as follows: <span class="html-italic">E. minima</span>: <span class="html-italic">Epilohmannia cylindrica minima</span> (Schuster, 1960); <span class="html-italic">H. minutissima</span>: <span class="html-italic">Hypochthoniella minutissima</span> (Berlese, 1904); <span class="html-italic">O. nova</span>: <span class="html-italic">Oppiella</span> (<span class="html-italic">O.</span>) <span class="html-italic">nova</span> (Oudemans, 1902); <span class="html-italic">R. insculpta</span>: <span class="html-italic">Ramusella (Insculptoppia) insculpta</span> (Paoli, 1908); <span class="html-italic">T. sarekensis</span>: <span class="html-italic">Tectocepheus sarekensis</span> (Trägårdh, 1910); <span class="html-italic">X. tegeocranus</span>: <span class="html-italic">Xenillus</span> (<span class="html-italic">X.</span>) <span class="html-italic">tegeocranus</span> (Hermann, 1804); <span class="html-italic">Z. acutirostris: Zetomimus (Protozetomimus) acutirostris</span> (Mihelčič, 1957).</p>
Full article ">Figure 6
<p>Hierarchical clustering analysis of the seven oribatid mite species, representing 98.06% of the cumulative variance from the principal component analysis.</p>
Full article ">
29 pages, 9700 KiB  
Article
Evaluation of Rheological Properties of Asphalt Binder Modified with Biochar from Oat Hulls
by Camila Martinez-Toledo, Gonzalo Valdes-Vidal, Alejandra Calabi-Floody, María Eugenia Gonzalez and Oscar Reyes-Ortiz
Materials 2024, 17(17), 4312; https://doi.org/10.3390/ma17174312 - 30 Aug 2024
Viewed by 497
Abstract
In this study, the effect of biochar from oat hulls (BO) on the rheological properties of a PG 64-22 asphalt binder was evaluated using a full factorial design, which included the following factors: pyrolysis temperature (PT) (300 °C and 500 °C), BO particle [...] Read more.
In this study, the effect of biochar from oat hulls (BO) on the rheological properties of a PG 64-22 asphalt binder was evaluated using a full factorial design, which included the following factors: pyrolysis temperature (PT) (300 °C and 500 °C), BO particle size (<20 µm and <75 µm), and the amount of BO (2.5%, 5%, and 7.5%). First, the morphological and physicochemical properties of BO were analyzed by comparing it with graphite powder (CFG) and commercial activated carbon (CAC). The physicochemical properties of the modified asphalt binder were then evaluated using confocal laser microscopy, scanning electron microscopy (SEM–EDX), and Fourier-transform infrared spectroscopy (FTIR). Its storage stability was also evaluated. Subsequently, the rutting parameter G*/sin(δ) and the Fraass breaking point were analyzed to select asphalt binders that extended their viscoelastic range. The asphalt binders selected were those with 2.5%, 5%, and 7.5% BO, produced at a PT of 300 °C with a particle size <20 µm (BO300S). Next, the rheological properties of the selected samples were evaluated by testing for rotational viscosity, rutting parameter G*/sin(δ), multiple stress creep recovery (MSCR), fatigue parameter G*·sin(δ), and creep stiffness by bending beam rheometry (BBR). The rheological aging index according to rutting parameter G*/sin(δ) (RAI) was also evaluated. These tests were conducted in different states of the asphalt binder: original, short-term aged, and long-term aged. According to the results, the application of BO300S significantly increased the resistance of the asphalt binder to rutting and rotational viscosity, proportional to the amount added to the asphalt binder. Moreover, low modifier percentages improved fatigue resistance, outperforming CFG and CAC. In addition, it performs well at low service temperatures, registering better resistance than the control asphalt binders. Full article
(This article belongs to the Special Issue Obtaining and Characterization of New Materials (5th Edition))
Show Figures

Figure 1

Figure 1
<p>BO production. “BO300S”: BO pyrolyzed at a PT of 300 °C with a particle size &lt;20 µm. “BO300L”: BO pyrolyzed at a PT of 300 °C with a particle size &lt;75 µm. “BO500S”: BO pyrolyzed at a PT of 500 °C with a particle size &lt;20 µm. “BO500L”: BO pyrolyzed at a PT of 500 °C with a particle size &lt;75 µm.</p>
Full article ">Figure 2
<p>Morphological characterization. (<b>A1</b>,<b>A2</b>) are SEM micrographs of BO300S. (<b>B1</b>,<b>B2</b>) are SEM micrographs of BO500S. (<b>C1</b>,<b>C2</b>) are SEM micrographs of BO300L. (<b>D1</b>,<b>D2</b>) are SEM micrographs of BO500L. (<b>E1</b>,<b>E2</b>) are SEM micrographs of the CFG. (<b>F1</b>,<b>F2</b>) are SEM micrographs of the CAC. Yellow-edged rectangles indicate micrographs at higher magnification, and yellow arrows indicate particle characteristics. (<b>G</b>) is the S<sub>BET</sub> surface area graph. (<b>H</b>) is the graph of D<sub>P</sub> versus V<sub>P</sub>. (<b>I</b>) is the graph of particle size distribution &lt;20 µm. (<b>J</b>) is the graph of particle size distribution &lt;75 µm.</p>
Full article ">Figure 3
<p>FTIR spectra. (<b>A</b>) compares the spectra of BO300S and BO300L to BO500S and BO500L. (<b>B</b>) compares the spectra of BO and the commercial modifier controls, CFG and CAC.</p>
Full article ">Figure 4
<p>Confocal laser microscopy images showing the distribution of different percentages of BO300S in the asphalt binder. Yellow arrows indicate elongated BO300S particles.</p>
Full article ">Figure 5
<p>FTIR spectra. (<b>A</b>) comparison between AB and AB-0.0. (<b>B</b>) comparison between AB, AB-0.0, and different percentages of AB-BO300L. (<b>C</b>) comparison between AB, AB-0.0, and various percentages of AB-BO500L.</p>
Full article ">Figure 6
<p>Storage stability of AB-BO300L and AB-BO500S with different modification percentages.</p>
Full article ">Figure 7
<p>The effect of BO factors on the rutting parameter G*/sin(δ) of the asphalt binder evaluated at 64 °C previously subjected to short-term aging in the RTFO. (<b>A1</b>,<b>A2</b>) correspond to the PT effect of 300 °C and 500 °C, respectively. (<b>B1</b>,<b>B2</b>) correspond to the effect of particle size &lt;20 µm and &lt;75 µm, respectively. (<b>C1</b>–<b>C3</b>) correspond to the effect of the amount of BO equivalent to 2.5%, 5%, and 7.5%, respectively.</p>
Full article ">Figure 8
<p>The effect of BO factors on the Fraass breaking point of the asphalt binder with long-term aging in a PAV. (<b>A1</b>,<b>A2</b>) correspond to the PT effect of 300 °C and 500 °C, respectively. (<b>B1</b>,<b>B2</b>) correspond to the effect of particle size &lt;20 µm and &lt;75 µm, respectively. (<b>C1</b>–<b>C3</b>) correspond to the effect of 2.5%, 5%, and 7.5% of BO, respectively.</p>
Full article ">Figure 9
<p>Viscoelastic range as a function of the rutting parameter G*/sin(δ) and the Fraas breaking point.</p>
Full article ">Figure 10
<p>Rotational viscosity at 120 °C, 135 °C, 150 °C, and 165 °C in the original condition of the samples.</p>
Full article ">Figure 11
<p>Rutting parameter G*/sin(δ) versus phase shift angle (δ) obtained between 58 °C and 76 °C in asphalt binders: (<b>A</b>) in the original state, and (<b>B</b>) short-term aged in the RTFO. The gray dashed lines indicate the allowable Superpave specification for the rutting parameter G*/sin(δ).</p>
Full article ">Figure 12
<p>Cumulative deformation obtained at 70 °C according to the MSCR test. Up to 200 s, a stress level of 0.1 kPa was applied, and between 200 s and 300 s, a stress level of 3.2 kPa was applied.</p>
Full article ">Figure 13
<p>Non-recoverable creep compliance (Jnr) versus recovery percentage (R%) under different stress levels: (<b>A</b>) 0.1 kPa and (<b>B</b>) 3.2 kPa.</p>
Full article ">Figure 14
<p>Fatigue parameter G*·sin(δ) evaluated between 16 °C and 31 °C in long-term aged samples by the PAV. The gray line represents the Superpave specification requirement for the fatigue parameter.</p>
Full article ">Figure 15
<p>Creep stiffness (S) and m-value results at −6 °C and −12 °C obtained using the BBR test. The gray lines represent the Superpave specification requirement for the S parameter and m-value.</p>
Full article ">Figure 16
<p>Rheological aging index (RAI) between 58 °C and 76 °C.</p>
Full article ">
8 pages, 2030 KiB  
Communication
CRISPR/Cas9-Mediated Resistance to Wheat Dwarf Virus in Hexaploid Wheat (Triticum aestivum L.)
by Xiaoyu Yuan, Keya Xu, Fang Yan, Zhiyuan Liu, Carl Spetz, Huanbin Zhou, Xiaojie Wang, Huaibing Jin, Xifeng Wang and Yan Liu
Viruses 2024, 16(9), 1382; https://doi.org/10.3390/v16091382 - 29 Aug 2024
Viewed by 414
Abstract
Wheat dwarf virus (WDV, genus Mastrevirus, family Geminiviridae) is one of the causal agents of wheat viral disease, which severely impacts wheat production in most wheat-growing regions in the world. Currently, there is little information about natural resistance against WDV in [...] Read more.
Wheat dwarf virus (WDV, genus Mastrevirus, family Geminiviridae) is one of the causal agents of wheat viral disease, which severely impacts wheat production in most wheat-growing regions in the world. Currently, there is little information about natural resistance against WDV in common wheat germplasms. CRISPR/Cas9 technology is being utilized to manufacture transgenic plants resistant to different diseases. In the present study, we used the CRISPR/Cas9 system targeting overlapping regions of coat protein (CP) and movement protein (MP) (referred to as CP/MP) or large intergenic region (LIR) in the wheat variety ‘Fielder’ to develop resistance against WDV. WDV-inoculated T1 progenies expressing Cas9 and sgRNA for CP/MP and LIR showed complete resistance against WDV and no accumulation of viral DNA compared with control plants. Mutation analysis revealed that the CP/MP and LIR targeting sites have small indels in the corresponding Cas9-positive plants. Additionally, virus inhibition and indel mutations occurred in T2 homozygous lines. Together, our work gives efficient results of the engineering of CRISPR/Cas9-mediated WDV resistance in common wheat plants, and the specific sgRNAs identified in this study can be extended to utilize the CRISPR/Cas9 system to confer resistance to WDV in other cereal crops such as barley, oats, and rye. Full article
(This article belongs to the Special Issue Plant Virus Interactions with Hosts: Mechanisms and Applications)
Show Figures

Figure 1

Figure 1
<p>CRISPR/Cas9-mediated wheat dwarf virus (WDV) interference in hexaploid wheat. (<b>A</b>) WDV genome structure and positions of the sgRNA target sites; (<b>B</b>) schematic diagram of the CRISPR/Cas9 system constructed in this study; (<b>C</b>,<b>D</b>) Hi-TOM sequence analysis of CP/MP-sgRNA and LIR-sgRNA-Cas9-induced mutations at the target sites and the representative Sanger sequencing chromatographs at 7 dpi. WT, Wild-type control. PAM is in red and gRNAs in blue. Green letters indicate inserted nucleotides. Blue dashes denote nucleotide deletions. −/+ indicates deletion/insertion of nucleotides.</p>
Full article ">Figure 2
<p>Molecular and phenotypic evidence of WDV resistance in T<sub>1</sub> generation of lines L69 and C41. (<b>A</b>,<b>B</b>) mRNA expression levels of the viral Rep gene in transgenic lines L69 and C41 were examined using semi-quantitative reverse transcription PCR (sqRT-PCR) assay at the indicated time points; (<b>C</b>) WDV accumulation in systemic leaves at 14 dpi were assessed using reverse transcription-quantitative PCR (RT-qPCR). Error bars represent SD, asterisks indicate significance, * <span class="html-italic">p</span> &lt; 0.05. Data are representative of three biological replicates; (<b>D</b>) Northern blot confirms the absence of WDV in systemic leaves at 30 dpi. Ethidium bromide staining of ribosomal RNAs (rRNA) was used as loading controls; (<b>E</b>) Western blot confirmation of the Cas9 protein at 30 dpi using an anti-FLAG primary antibody. Ponceau-S-stained Rubisco served as a loading control; (<b>F</b>,<b>G</b>) Expression level of sgRNAs in wheat systemic leaves at 30 dpi was assessed using RT-qPCR; (<b>H</b>,<b>I</b>) symptoms of WDV in the transgenic lines L69 and C41 in comparison with wild-type wheat plants at 80 dpi.</p>
Full article ">
25 pages, 6532 KiB  
Article
Exploring Evolutionary Pathways and Abiotic Stress Responses through Genome-Wide Identification and Analysis of the Alternative Oxidase (AOX) Gene Family in Common Oat (Avena sativa)
by Boyang Liu, Zecheng Zhang, Jinghan Peng, Haipeng Mou, Zhaoting Wang, Yixin Dao, Tianqi Liu, Dandan Kong, Siyu Liu, Yanli Xiong, Yi Xiong, Junming Zhao, Zhixiao Dong, Youjun Chen and Xiao Ma
Int. J. Mol. Sci. 2024, 25(17), 9383; https://doi.org/10.3390/ijms25179383 - 29 Aug 2024
Viewed by 356
Abstract
The alternative oxidase (AOX), a common terminal oxidase in the electron transfer chain (ETC) of plants, plays a crucial role in stress resilience and plant growth and development. Oat (Avena sativa), an important crop with high nutritional value, has not been [...] Read more.
The alternative oxidase (AOX), a common terminal oxidase in the electron transfer chain (ETC) of plants, plays a crucial role in stress resilience and plant growth and development. Oat (Avena sativa), an important crop with high nutritional value, has not been comprehensively studied regarding the AsAOX gene family. Therefore, this study explored the responses and potential functions of the AsAOX gene family to various abiotic stresses and their potential evolutionary pathways. Additionally, we conducted a genome-wide analysis to explore the evolutionary conservation and divergence of AOX gene families among three Avena species (Avena sativa, Avena insularis, Avena longiglumis) and four Poaceae species (Avena sativa, Oryza sativa, Triticum aestivum, and Brachypodium distachyon). We identified 12 AsAOX, 9 AiAOX, and 4 AlAOX gene family members. Phylogenetic, motif, domain, gene structure, and selective pressure analyses revealed that most AsAOXs, AiAOXs, and AlAOXs are evolutionarily conserved. We also identified 16 AsAOX segmental duplication pairs, suggesting that segmental duplication may have contributed to the expansion of the AsAOX gene family, potentially preserving these genes through subfunctionalization. Chromosome polyploidization, gene structural variations, and gene fragment recombination likely contributed to the evolution and expansion of the AsAOX gene family as well. Additionally, we hypothesize that AsAOX2 may have potential function in resisting wounding and heat stresses, while AsAOX4 could be specifically involved in mitigating wounding stress. AsAOX11 might contribute to resistance against chromium and waterlogging stresses. AsAOX8 may have potential fuction in mitigating ABA-mediated stress. AsAOX12 and AsAOX5 are most likely to have potential function in mitigating salt and drought stresses, respectively. This study elucidates the potential evolutionary pathways of the AsAOXs gene family, explores their responses and potential functions to various abiotic stresses, identifies potential candidate genes for future functional studies, and facilitates molecular breeding applications in A. sativa. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the mitochondrial respiratory electron transport chain in higher plants [<a href="#B13-ijms-25-09383" class="html-bibr">13</a>]. It illustrates the position of complexes I to V and AOX in the electron transport chain.</p>
Full article ">Figure 2
<p>The chromosomal localization of AsAOXs, AlAOXs, and AiAOXs was depicted using brown, green, and blue colors to represent <span class="html-italic">A. sativa</span>, <span class="html-italic">A. longiglumis</span>, and <span class="html-italic">A. insularis</span>, respectively. The shades of these colors on the chromosomes reflect varying gene densities.</p>
Full article ">Figure 3
<p>An ML phylogenetic tree of 81 AOX protein sequences from 11 species (<span class="html-italic">A. sativa</span>, <span class="html-italic">Oryza sativa</span>, <span class="html-italic">Arabidopsis thaliana</span>, <span class="html-italic">Triticum aestivum</span>, <span class="html-italic">Zea mays</span>, <span class="html-italic">Brachypodium distachyon</span>, <span class="html-italic">Hordeum vulgare</span>, <span class="html-italic">Glycine max</span>, <span class="html-italic">Medicago truncatula</span>, <span class="html-italic">Sorghum bicolor</span>, <span class="html-italic">Secale cereale</span>) was constructed using MEGA v7.0. Pink circles of different sizes represent various bootstrap values, with the circle size indicating bootstrap values ranging from 0 to 100. The number following each species’ Latin name indicates the quantity of AOX members from that species.</p>
Full article ">Figure 4
<p>(<b>A</b>) Phylogenetic analysis of 25 AOX protein sequences (12 AsAOX, 9 AiAOX, 4 AlAOX), with bootstrap values depicted on branches ranging from 0 to 100. (<b>B</b>) Conservation motif analysis. (<b>C</b>) Conservation domain analysis. (<b>D</b>) Gene structure composition, where black lines, yellow boxes, and green boxes represent introns, exons (CDS), and UTRs (Untranslated Regions), respectively. Different colored backgrounds indicate clustering of different AOX gene members.</p>
Full article ">Figure 5
<p>(<b>A</b>) Synteny analysis of AsAOXs with the AOX gene families of eight species. These include <span class="html-italic">A. sativa</span>, closely related Triticeae cereals <span class="html-italic">Triticum aestivum</span>, <span class="html-italic">Hordeum vulgare</span>, and <span class="html-italic">Secale cereale</span>, globally important crops Oryza sativa and <span class="html-italic">Sorghum bicolor</span>, as well as the model plant <span class="html-italic">Brachypodium distachyon</span>. (<b>B</b>) Syteny analysis of <span class="html-italic">A. sativa</span> with its possible two ancestral species <span class="html-italic">A. insularis</span> and <span class="html-italic">A. longiglumis</span>. Different colored lines represent synteny gene pairs between the AOX genes of the respective species and AsAOX genes, while grey lines indicate synteny blocks in other genes.</p>
Full article ">Figure 6
<p>Gene duplication analysis of AsAOXs, AiAOXs, and AlAOXs. (<b>A</b>) Gene duplication analysis of AsAOXs. Red lines represent paralogous gene pairs of the AsAOX gene family. Shades of grey indicate the density of genes. (<b>B</b>) The green and purple lines represent paralogous gene pairs of the AiAOX and AlAOX gene families, respectively.</p>
Full article ">Figure 7
<p>AsAOXs, AiAOXs, and AlAOXs gene promoter region CREs analysis. (<b>A</b>) An ML phylogenetic tree of AsAOXs, with bootstrap values indicated on branches ranging from 0 to 100. (<b>B</b>) Analysis of the number, distribution, and classification of Cis-regulatory elements in AsAOXs. (<b>C</b>) Comparative analysis of CREs in the promoter regions of AsAOXs, AlAOXs, and AiAOXs genes.</p>
Full article ">Figure 8
<p>GO enrichment analysis of 12 AsAOX genes. (<b>A</b>) GO terms enriched for each AsAOX gene. (<b>B</b>) Number of AsAOX genes contained within each GO term.</p>
Full article ">Figure 9
<p>RT-qPCR analysis reveals the relative expression levels of 12 AsAOX genes at different time points (0 h, 3 h, 6 h, 9 h, 24 h, 48 h) under five abiotic stress conditions in oat leaves and roots. If expression levels at 0 h are too low to detect, the subsequent time point is used as control (ck). Error bars represent the standard error of three biological replicates.</p>
Full article ">Figure 10
<p>(<b>A</b>,<b>B</b>) depict the expression profiles of 12 AsAOX genes under drought stress and salt stress from RNA-Seq data, respectively. Shades of red and green indicate the degree of downregulation and upregulation, respectively.</p>
Full article ">
11 pages, 704 KiB  
Article
A Randomized Controlled Trial to Assess the Feasibility and Practicability of an Oatmeal Intervention in Individuals with Type 2 Diabetes: A Pilot Study in the Outpatient Sector
by Michél Fiedler, Nicolle Müller, Christof Kloos, Guido Kramer, Christiane Kellner, Sebastian Schmidt, Gunter Wolf and Nadine Kuniß
J. Clin. Med. 2024, 13(17), 5126; https://doi.org/10.3390/jcm13175126 - 29 Aug 2024
Viewed by 409
Abstract
Background/Objectives: The aim of this study was to investigate the feasibility and practicability of repeated three-day sequences of a hypocaloric oat-based nutrition intervention (OI) in insulin-treated outpatients with type 2 diabetes and severe insulin resistance. Methods: A randomized, two-armed pilot study [...] Read more.
Background/Objectives: The aim of this study was to investigate the feasibility and practicability of repeated three-day sequences of a hypocaloric oat-based nutrition intervention (OI) in insulin-treated outpatients with type 2 diabetes and severe insulin resistance. Methods: A randomized, two-armed pilot study was conducted with three months of intervention and three months follow-up with 17 participants with insulin resistance (≥1 IU/kg body weight). Group A (n = 10) performed one sequence of OI; Group B (n = 7) performed two sequences monthly. A sequence was 3 consecutive days of oat consumption with approximately 800 kcal/d. The main objective was to assess feasibility (≥70% completers) and practicability regarding performance aspects. Biomedical parameters such as HbA1 c were observed. To evaluate the state of health, a standardized questionnaire was used (EQ-5 D). Results: OI was feasible (13/17 completer participants (76.5%): 70.0% Group A, 85.7% Group B). Individually perceived practicability was reported as good by 10/16 participants (62.5%). Total insulin dosage decreased from 138 ± 35 IU at baseline to 126 ± 42 IU after OI (p = 0.04) and 127 ± 42 IU after follow-up (p = 0.05). HbA1 c was lower after OI (−0.3 ± 0.1%; p = 0.01) in all participants. Participants in Group B tended to have greater reductions in insulin (Δ−19 IU vs. Δ−4 IU; p = 0.42) and weight loss (Δ−2.8 kg vs. Δ−0.2 kg; p = 0.65) after follow-up. Severe hypoglycemia was not observed. EQ-5 D increase not significantly after follow-up (57.2 ± 24.0% vs. 64.7 ± 21.5%; p = 0.21). Conclusions: The feasibility and practicability of OI in outpatients were demonstrated. OI frequency appears to correlate with insulin reduction and weight loss. Proper insulin dose adaptation during OI is necessary. Presumably, repeated OIs are required for substantial beneficial metabolic effects. Full article
(This article belongs to the Section Endocrinology & Metabolism)
Show Figures

Figure 1

Figure 1
<p>Flow chart of recruitment and study procedure. Sequence of OI = three consecutive days. * adjustment of insulin dosage, checking for hypoglycemia, checking for adverse events on every consultation.</p>
Full article ">Figure 2
<p>Course of total insulin per day at t(0): study entry, t(1): after intervention, t(2): after follow-up. * <span class="html-italic">p</span> &lt; 0.05 t(0) vs. t(1); ** <span class="html-italic">p</span> &lt; 0.05 t(0) vs. t(2).</p>
Full article ">
12 pages, 484 KiB  
Article
Role of Wilting Time on the Chemical Composition, Biological Profile, and Fermentative Quality of Cereal and Legume Intercropping Silage
by Cristiana Maduro Dias, Hélder Nunes, Mariana Aguiar, Arnaldo Pereira, João Madruga and Alfredo Borba
Fermentation 2024, 10(9), 448; https://doi.org/10.3390/fermentation10090448 - 28 Aug 2024
Viewed by 482
Abstract
Agricultural production in the Azores primarily focuses on the livestock sector, notably, dairy production, where cows graze year-round in a rotational system. To maintain pasture productivity, farmers often rely on synthetic nitrogen fertilizers, which have adverse environmental impacts like ammonia emissions and nitrate [...] Read more.
Agricultural production in the Azores primarily focuses on the livestock sector, notably, dairy production, where cows graze year-round in a rotational system. To maintain pasture productivity, farmers often rely on synthetic nitrogen fertilizers, which have adverse environmental impacts like ammonia emissions and nitrate leaching. Alternatively, nitrogen-fixing crops like legumes are explored as green manures to enhance soil quality and reduce dependence on chemical fertilizers. The traditional practice of using mixed forages of legumes and grasses, known as “outonos” or intercrops, has been crucial but is declining over time. These mixtures include plants such as lupins, Vicia faba, oats, and vetch, noted for their adaptability and nitrogen-fixing ability. Due to the high perishability of these crops, effective conservation strategies like ensiling are essential to preserve forage nutritional quality through controlled fermentation. This study evaluates the productivity and quality of intercrop forages in the Azores, focusing on fresh samples and silage prepared with wilting times of 0, 24, 48, and 96 h, followed by comprehensive chemical analyses. Results showed significant changes in fiber components (neutral detergent fiber, acid detergent fiber, and acid detergent lignin) with increased wilting time, leading to reduced digestibility. However, wilting improved dry matter content. Full article
(This article belongs to the Section Industrial Fermentation)
Show Figures

Figure 1

Figure 1
<p>Scheme of the use and partition of the samples.</p>
Full article ">
Back to TopTop