[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
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
remove_circle_outline

Search Results (322)

Search Parameters:
Keywords = nitrate leaching

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 906 KiB  
Article
Zootechnical and Municipal Solid Waste Digestates: Effects on Soil Nitrogen Mineralization and Kinetics
by Gabriella Rossi, Claudio Beni, Silvia Socciarelli and Ulderico Neri
Biomass 2025, 5(1), 5; https://doi.org/10.3390/biomass5010005 - 16 Jan 2025
Viewed by 314
Abstract
Soil fertilization with fertilizers derived from renewable sources is a topic of great interest in terms of the sustainable management of organic waste. To optimize the management of nitrogen supplied to the soil with digestates, it is necessary to deepen knowledge on the [...] Read more.
Soil fertilization with fertilizers derived from renewable sources is a topic of great interest in terms of the sustainable management of organic waste. To optimize the management of nitrogen supplied to the soil with digestates, it is necessary to deepen knowledge on the process of mineralization of organic nitrogen over time. In this research, a laboratory incubation system was utilized to study the impact of various digestate sources on nitrogen mineralization processes in soils and nitrogen mineralization kinetics. Six types of digestates of different origins and composition were administered to soil and the soil samples were placed under controlled conditions. The release of N was determined by measuring ammonium-N and nitrate-N concentrations in leachates during a 12-week period of incubation. The nonlinear regression technique was used to fit the cumulative leaching of total N to the Stanford and Smith first-order kinetic model during the incubation period. The results showed that the differences between digestates, nitrogen and organic carbon concentration, and C/N ratio influenced both ammonification and nitrification processes in the soil and the nitrogen mineralization kinetics. The processing of the statistical data highlighted that the potentially mineralizable nitrogen (MPN) followed first-order kinetics. Full article
Show Figures

Figure 1

Figure 1
<p>Cumulative concentrations (mg N kg<sup>−1</sup> of dry soil) of mineralized N (NH<sub>4</sub><sup>+</sup>-N + NO<sub>3</sub><sup>−</sup>-N). Bars with different letters are significant at <span class="html-italic">p</span> &lt; 0.05 according to Tukey’s test. Error bars describe the standard error of the mean (n = 3).</p>
Full article ">Figure 2
<p>Curve fit plots showing experimental data (as an average of three replicates) and predicted values according to the first-order kinetics model against incubation time (Week) of the net cumulative inorganic nitrogen (mg kg<sup>−1</sup> dry soil). Ammonium sulfate (AS), digested pig slurry as such (PS), digested pig slurry–solid fraction (PF), digested bovine manure (BM), digested OFMSW (DO), dried digested OFMSW (DD), and digested and composted OFMSW (DC).</p>
Full article ">
17 pages, 480 KiB  
Article
New Insights into Fertilisation with Animal Manure for Annual Double-Cropping Systems in Nitrate-Vulnerable Zones of Northeastern Spain
by Dolores Quilez, Monica Guillén, Marta Vallés, Arturo Daudén and Beatriz Moreno-García
Agronomy 2025, 15(1), 142; https://doi.org/10.3390/agronomy15010142 - 9 Jan 2025
Viewed by 393
Abstract
Maize double-cropping production systems in Mediterranean areas have a great nitrogen extraction capacity and high nitrogen (N) requirements. This study aims to assess whether in these farming systems, animal manure can be applied, using adequate management practices, at levels exceeding the maximum annual [...] Read more.
Maize double-cropping production systems in Mediterranean areas have a great nitrogen extraction capacity and high nitrogen (N) requirements. This study aims to assess whether in these farming systems, animal manure can be applied, using adequate management practices, at levels exceeding the maximum annual amount of livestock manure established in the European Nitrate Directive for vulnerable zones (170 kg N ha−1) without increasing the risk of water nitrate contamination. We compare the risk of nitrate leaching under two fertilisation strategies, one with synthetic fertilisers and the second with a maximised application of pig slurry, exceeding the limits of the EU Nitrate Directive, in two soil types. Crop yields, N extraction and nitrate concentrations below the crop root zone were not affected by the fertilisation strategies at each site. The results show that pig slurry can be applied above the limit of 170 kg N ha−1 under the conditions of the study, up to 360 kg N ha−1, without increasing the risk for nitrate leaching. Full article
Show Figures

Figure 1

Figure 1
<p>Average nitrate concentration in the soil solution at the different sampling times in the two treatments (P: pig slurry, M: synthetic fertiliser) in the Barluenga (1.2 m depth) and Torremira (0.45 m depth) fields.</p>
Full article ">
19 pages, 2360 KiB  
Article
Deep-Learning-Driven Insights into Nitrogen Leaching for Sustainable Land Use and Agricultural Practices
by Caixia Hu, Jie Li, Yaxu Pang, Lan Luo, Fang Liu, Wenhao Wu, Yan Xu, Houyu Li, Bingcang Tan and Guilong Zhang
Land 2025, 14(1), 69; https://doi.org/10.3390/land14010069 - 2 Jan 2025
Viewed by 364
Abstract
Nitrate leaching from soil presents a significant threat to soil health, as it can result in nutrient loss, soil acidification, and structural damage. It is crucial to quantify the spatial heterogeneity of nitrate leaching and its drivers. A total of 509 observational data [...] Read more.
Nitrate leaching from soil presents a significant threat to soil health, as it can result in nutrient loss, soil acidification, and structural damage. It is crucial to quantify the spatial heterogeneity of nitrate leaching and its drivers. A total of 509 observational data points regarding nitrate leaching in northern China were collected, capturing the spatial and temporal variations across crops such as winter wheat, maize, and greenhouse vegetables. A machine learning (ML) model for predicting nitrate leaching was then developed, with the random forest (RF) model outperforming the support vector machine (SVM), extreme gradient boosting (XGBoost), and convolutional neural network (CNN) models, achieving an R2 of 0.75. However, the performance improved significantly after integrating the four models with Bayesian optimization (all models had R2 > 0.56), which realized quantitative prediction capabilities for nitrate leaching loss concentrations. Moreover, the XGBoost model exhibited the highest fitting accuracy and the smallest error in estimating nitrate leaching losses, with an R2 value of 0.79 and an average absolute error (MAE) of 3.87 kg/ha. Analyses of the feature importance and SHAP values in the optimal XGBoost model identified soil organic matter, chemical nitrogen fertilizer input, and water input (including rainfall and irrigation) as the main indicators of nitrate leaching loss. The ML-based modeling method developed overcomes the difficulty of the determination of the functional relationship between nitrate loss intensity and its influencing factors, providing a data-driven solution for estimating nitrate–nitrogen loss in farmlands in North China and strengthening sustainable agricultural practices. Full article
Show Figures

Figure 1

Figure 1
<p>Raw data and processed data QQ plots of average annual temperature (<b>A</b>), average annual rainfall (<b>B</b>), soil type (<b>C</b>), chemical N fertilizer input (<b>D</b>), organic N fertilizer input (<b>E</b>), irrigation amount (<b>F</b>), irrigation methods (<b>G</b>), soil total N (<b>H</b>), soil organic matter (<b>I</b>), soil pH (<b>J</b>), soil bulk density (<b>K</b>) and soil clay (<b>L</b>).</p>
Full article ">Figure 2
<p>Pearson’s correlation matrix of independent variables.</p>
Full article ">Figure 3
<p>Comparison of R<sup>2</sup>, <span class="html-italic">RMSE</span>, and <span class="html-italic">MAE</span> using the SVM (<b>A</b>), RF (<b>B</b>), XGBoost (<b>C</b>), and CNN (<b>D</b>) models for nitrate–nitrogen loss rate prediction on training and test datasets. Abbreviations: SVM, support vector machine; RF, random forest; XGBoost, extreme gradient boosting; CNN, convolutional neural network. Moving average error (<span class="html-italic">MAE</span>), root mean square error (<span class="html-italic">RMSE</span>), and coefficient of determination (R<sup>2</sup>).</p>
Full article ">Figure 4
<p>Result of Bayesian-optimized hyperparameters in SVM (<b>A</b>), RF (<b>B</b>), XGBoost (<b>C</b>), and CNN (<b>D</b>) models for nitrate–nitrogen loss rate prediction. Abbreviations: SVM, support vector machine; RF, random forest; XGBoost, extreme gradient boosting; CNN, convolutional neural network. Moving average error (<span class="html-italic">MAE</span>), root mean square error (<span class="html-italic">RMSE</span>), and coefficient of determination (R<sup>2</sup>).</p>
Full article ">Figure 5
<p>Ranking of the importance of input features (<b>A</b>) and the SHAP value for a particular variable (<b>B</b>).</p>
Full article ">
18 pages, 1817 KiB  
Article
Model-Based Valuation of Ecosystem Services Using Bio-Economic Farm Models: Insights for Designing Green Tax Policies and Payment for Ecosystem Services
by Seyed-Ali Hosseini-Yekani, Stefan Tomaczewski and Peter Zander
Agriculture 2025, 15(1), 60; https://doi.org/10.3390/agriculture15010060 - 29 Dec 2024
Viewed by 463
Abstract
The integration of ecosystem services (ESs) valuation into agricultural policy frameworks is critical for fostering sustainable land management practices. This study leverages the redesigned version of the bio-economic farm model MODAM (Multi-Objective Decision Support Tool for Agro-Ecosystem Management) to estimate the shadow prices [...] Read more.
The integration of ecosystem services (ESs) valuation into agricultural policy frameworks is critical for fostering sustainable land management practices. This study leverages the redesigned version of the bio-economic farm model MODAM (Multi-Objective Decision Support Tool for Agro-Ecosystem Management) to estimate the shadow prices of ESs, enabling the derivation of demand and supply curves for nitrate leaching and soil erosion control, respectively. Two hypothetical farms in Brandenburg, Germany—a smaller, arable farm in Märkisch-Oderland and a larger, diversified farm with livestock in Oder-Spree—are analyzed to explore the heterogeneity in shadow prices and corresponding cropping patterns. The results reveal that larger farms exhibit greater elasticity in response to green taxes on nitrate use and lower costs for supplying erosion control compared to smaller farms. This study highlights the utility of shadow prices as proxies for setting green taxes and payments for ecosystem services (PESs), while emphasizing the need for differentiated policy designs to address disparities between farm types. This research underscores the potential of model-based ESs valuation to provide robust economic measures for policy design, fostering sustainable agricultural practices and ecosystem conservation. Full article
(This article belongs to the Special Issue Agricultural Policies toward Sustainable Farm Development)
Show Figures

Figure 1

Figure 1
<p>Farmer’s demand curve for ESs <span class="html-italic">s</span> in year <span class="html-italic">t</span>. Adapted from Kaiser and Messer (2012) [<a href="#B17-agriculture-15-00060" class="html-bibr">17</a>].</p>
Full article ">Figure 2
<p>Farmer’s supply curve for ESs <span class="html-italic">d</span> in year <span class="html-italic">t</span>. Adapted from Kaiser and Messer (2012) [<a href="#B17-agriculture-15-00060" class="html-bibr">17</a>].</p>
Full article ">Figure 3
<p>Nitrate leaching demand curve of the hypothetical farm type in Märkisch-Oderland (Brandenburg, Germany) and its optimal cropping patterns simulated under different levels of a green tax. Source: Own processing, results obtained using MODAM model implemented in GAMS software (Release 48.4.0).</p>
Full article ">Figure 4
<p>Nitrate leaching demand curve of the hypothetical farm type in Oder-Spree (Brandenburg, Germany) and its optimal cropping patterns simulated under different levels of a green tax. Source: Own processing, results obtained using MODAM model implemented in GAMS software (Release 48.4.0).</p>
Full article ">Figure 5
<p>Nitrate leaching demand curves of two hypothetical farm types in Märkisch-Oderland and Oder-Spree. Source: Own processing, results obtained using MODAM model implemented in GAMS software (Release 48.4.0).</p>
Full article ">Figure 6
<p>Erosion control supply curve of the hypothetical farm type in Märkisch-Oderland and its optimal cropping patterns simulated under different levels of a PES. Source: Own processing, results obtained using MODAM model implemented in GAMS software (Release 48.4.0).</p>
Full article ">Figure 7
<p>Erosion control supply curve of the hypothetical farm type in Oder-Spree and its optimal cropping patterns simulated under different levels of a PES. Source: Own processing, results obtained using MODAM model implemented in GAMS software (Release 48.4.0).</p>
Full article ">Figure 8
<p>Soil erosion control supply curves of two hypothetical farm types in Märkisch-Oderland and Oder-Spree. Source: Own processing, results obtained using MODAM model implemented in GAMS software (Release 48.4.0).</p>
Full article ">
23 pages, 3177 KiB  
Article
Blending of Slow-Release N Fertilizer and Urea Improve Rainfed Maize Yield and Nitrogen Use Efficiency While Reducing Apparent N Losses
by Jinjin Guo, Hanran Yang, Yong Yuan, Pengzhou Yin, Nv Zhang, Zhizhao Lin, Qichang Ma, Qiliang Yang, Xiaogang Liu, Haidong Wang and Fucang Zhang
Agronomy 2025, 15(1), 11; https://doi.org/10.3390/agronomy15010011 - 25 Dec 2024
Viewed by 307
Abstract
Effective nitrogen (N) management practices are essential for achieving efficient and sustainable agricultural production. The purpose of this study was to improve N use efficiency (NUE) and minimize N loss by optimizing the rate and type of N fertilizer application while maintaining a [...] Read more.
Effective nitrogen (N) management practices are essential for achieving efficient and sustainable agricultural production. The purpose of this study was to improve N use efficiency (NUE) and minimize N loss by optimizing the rate and type of N fertilizer application while maintaining a high yield of maize. A two-year field experiment with U (urea), S (slow-release N fertilizer), and SU (blending of S and U) under four N application levels (N1: 90 kg ha−1, N2: 120 kg ha−1, N3: 180 kg ha−1, N4: 240 kg ha−1) was conducted to investigate their effects on ammonia (NH3) volatilization, residual soil nitrate N (NO3-N), yield, NUE, apparent N losses of rainfed maize. NH3 volatilization in SU and S were 38.46% and 16.57% lower than that in U, respectively. SU and S were found to reduce the apparent N losses by 42.98% and 62.23%. SU decreased NO3-N leaching in deep soils and increased NO3-N content in topsoil. Compared with U and S, SU significantly increased yield, plant N accumulation, and NUE. SUN4 achieved the maximum maize yield and plant N accumulation, averaging 7968.36 kg ha−1 and 166.45 kg ha−1. In addition, the high yield and NUE were obtained when the mixing ratio of S and U was 53–58% and the N application rate was 150–220 kg ha−1. The findings highlight that SU effectively reduces N losses while ensuring high yield, which could be used as one of the optimal N fertilization strategies for rainfed maize in Northwest China. Full article
(This article belongs to the Section Soil and Plant Nutrition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Precipitation and temperature during the maize growing seasons in 2019 and 2020.</p>
Full article ">Figure 2
<p>Daily NH<sub>3</sub> volatilization flux of maize under different N application rates (N1 to N4) and fertilizer types (U, S, SU) in 2019 and 2020. (<b>a</b>–<b>h</b>) Daily NH<sub>3</sub> volatilization flux for each N rate and year. CK = no N application; U = urea; S = slow-release N fertilizer; SU = urea blended with slow-release N fertilizer at a ratio of 3:7. Error bars are ±1 standard deviation of the mean (n = 3). The Y-bars on each data point should indicate SD.</p>
Full article ">Figure 3
<p>Cumulative NH<sub>3</sub> volatilization of maize under different N application rates (N1 to N4) and fertilizer types (U, S, SU) in 2019 and 2020. (<b>a</b>,<b>b</b>) Cumulative NH<sub>3</sub> volatilization for each year. CK = no N application; U = urea; S = slow-release N fertilizer; SU = urea blended with slow-release N fertilizer at a ratio of 3:7. Different letters indicate significance at the 5% level for the same year by the LSD test. Error bars are ±1 standard deviation of the mean (n = 3). The Y-bars on each data point should indicate SD.</p>
Full article ">Figure 4
<p>Residual soil <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi mathvariant="normal">N</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>3</mn> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> <mo>-</mo> <mi mathvariant="normal">N</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi mathvariant="normal">N</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>3</mn> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> <mo>-</mo> <mi mathvariant="normal">N</mi> </mrow> </semantics></math> content in the 0–120 cm soil layer of maize under different N application rates (N1 to N4) and N fertilizer types (U, S, SU) in 2019 and 2020. (<b>a</b>,<b>b</b>) Residual soil <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi mathvariant="normal">N</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>3</mn> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> <mo>-</mo> <mi mathvariant="normal">N</mi> </mrow> </semantics></math> for each year. (<b>c</b>–<b>h</b>) <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi mathvariant="normal">N</mi> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>3</mn> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> <mo>-</mo> <mi mathvariant="normal">N</mi> </mrow> </semantics></math> content for each N fertilizer type and year. CK = no N application; U = urea; S = slow-release N fertilizer; SU = urea blended with slow-release N fertilizer at a ratio of 3:7. Different letters indicate significance at the 5% level for the same year by the LSD test. Error bars are ±1 standard deviation of the mean (n = 3). The Y-bars on each data point should indicate SD.</p>
Full article ">Figure 5
<p>Maize dry matter accumulation under different N application rates (N1 to N4) and fertilizer types (U, S, SU) in 2019 and 2020. (<b>a</b>–<b>h</b>) Maize dry matter accumulation for each N application rate and year. Maize dry matter accumulation at the seedling stage is shown in <a href="#app1-agronomy-15-00011" class="html-app">Supplementary Materials</a>. CK = no N application; U = urea; S = slow-release N fertilizer; SU = urea blended with slow-release N fertilizer at a ratio of 3:7. Different letters indicate significance at the 5% level for the same year by the LSD test. Error bars are ±1 standard deviation of the mean (n = 3). The Y-bars on each data point should indicate SD.</p>
Full article ">Figure 6
<p>Plant N accumulation of maize at the maturity stage under different N application rates (N1 to N4) and fertilizer types (U, S, SU) in 2019 and 2020. (<b>a</b>,<b>b</b>) Plant N accumulation for each year. CK = no N application; U = urea; S = slow-release N fertilizer; SU = urea blended with slow-release N fertilizer at a ratio of 3:7. Different letters indicate significance at the 5% level for the same year by the LSD test. Error bars are ±1 standard deviation of the mean (n = 3). The Y-bars on each data point should indicate SD.</p>
Full article ">Figure 7
<p>Grain yield and NUE of maize under different N application rates (N1 to N4) and fertilizer types (U, S, SU) in 2019 and 2020. (<b>a</b>,<b>b</b>) Grain yield for each year. (<b>c</b>,<b>d</b>) NUE for each year. CK = no N application; U = urea; S = slow-release N fertilizer; SU = urea blended with slow-release N fertilizer at a ratio of 3:7. Different letters indicate significance at the 5% level for the same year by the LSD test. Error bars are ±1 standard deviation of the mean (n = 3). Both X- and Y-bars on each data point should indicate SD.</p>
Full article ">Figure 8
<p>Relationships between N application rate (N1 to N4) and blending ratio of SU (0–1, 0 = urea, 1 = slow-release N fertilizer, 0.7 = urea blended with slow-release N fertilizer at a ratio of 3:7) with grain yield and NUE of maize in 2019 and 2020 (MATLAB software and RStudio software). (<b>a</b>,<b>b</b>) Grain yield for each year. (<b>c</b>,<b>d</b>) NUE for each year.</p>
Full article ">
26 pages, 9214 KiB  
Article
Evaluation of Agricultural Measures to Safeguard the Vulnerable Karst Groundwater Habitat of the Black Olm (Proteus anguinus parkelj) from Nitrate Pollution
by Matjaž Glavan and Rozalija Cvejić
Sustainability 2024, 16(24), 11309; https://doi.org/10.3390/su162411309 - 23 Dec 2024
Viewed by 536
Abstract
The black olm (Proteus anguinus parkelj Sket & Arntzen) is an endemic species found exclusively in the Dobličica River subterranean water systems of the Dinaric karst in southern Slovenia. These unique habitats are vulnerable to contamination due to rapid water flow, primarily [...] Read more.
The black olm (Proteus anguinus parkelj Sket & Arntzen) is an endemic species found exclusively in the Dobličica River subterranean water systems of the Dinaric karst in southern Slovenia. These unique habitats are vulnerable to contamination due to rapid water flow, primarily from nitrates from agricultural fertilisers and untreated urban wastewater. The safe limit of nitrate concentration for olms is 9.2 mg NO3/L, yet measurements in karst springs have shown levels ranging from 3 mg to over 20 mg NO3/L. The SWAT modelling tool assessed agri-environmental and land use scenarios for their impact on nitrate leaching. Using the model, we identified hotspots with high nitrogen leaching potential that require immediate attention and implementation of better agricultural practices for fertiliser use. For these hotspots, the most effective approach combines scenarios of cover crops (R2), reduced fertilisation (R3), crop rotation (R4), and conversion of cropland to grassland (E2, E4, E5), potentially decreasing nitrate leaching by up to 60%. Implementing the best scenarios is expected to reduce nitrogen levels below the limit value of 9.2 mg NO3/L, essential for maintaining the black olm habitat. Full article
(This article belongs to the Topic Human Impact on Groundwater Environment)
Show Figures

Figure 1

Figure 1
<p>Geographic characteristics of study area presenting (<b>a</b>) topography, river network, SWAT model subbasin numbers (1–22), main river gauging point and karst springs outflow points, (<b>b</b>) slope and karst springs (with SWAT model subbasin id number), (<b>c</b>) land use, and (<b>d</b>) soil type.</p>
Full article ">Figure 2
<p>Display of measured and simulated flow data at the Gradac gauging station for the period of (<b>a</b>,<b>b</b>) calibration (1998–2010) and (<b>c</b>,<b>d</b>) validation (2011–2022) on a daily time step.</p>
Full article ">Figure 3
<p>Comparison of observed and simulated data on (<b>a</b>) the nitrate nitrogen yield (kg N-NO<sub>3</sub><sup>−</sup>/day) and water flow at the Gradac measuring station (subbasin 22) and (<b>b</b>) the nitrate nitrogen concentration (mg N-NO<sub>3</sub><sup>−</sup>/day) and water flow at the Jelševnik karst spring (subbasin 9) on a daily time step.</p>
Full article ">Figure 4
<p>Comparison of the basic scenario with alternative scenarios of (<b>a</b>) agricultural crop rotation change (R) and (<b>b</b>) agricultural land use change (E) in relation to the average nitrate nitrogen transported (kg N-NO<sub>3</sub><sup>−</sup>/ha per year) from HRUs in the groundwater flow to the surface water flow in the study area of 11 subbasins.</p>
Full article ">
17 pages, 3153 KiB  
Article
Influence of Biochar Feedstocks on Nitrate Adsorption Capacity
by Riad Eissa, Lordwin Jeyakumar, David B. McKenzie and Jianghua Wu
Earth 2024, 5(4), 1080-1096; https://doi.org/10.3390/earth5040055 - 23 Dec 2024
Viewed by 378
Abstract
The demand for intensive agriculture to boost food and crop production has increased. High nitrogen (N) fertilizer use is crucial for increasing agricultural productivity but often leads to significant nitrate losses, posing risks to surface and groundwater quality. This study examines the role [...] Read more.
The demand for intensive agriculture to boost food and crop production has increased. High nitrogen (N) fertilizer use is crucial for increasing agricultural productivity but often leads to significant nitrate losses, posing risks to surface and groundwater quality. This study examines the role of biochar as a soil amendment to enhance nutrient retention and mitigate nitrate leaching. By improving nitrogen efficiency, biochar offers a sustainable strategy to reduce the environmental impacts of intensive agriculture while maintaining soil fertility. An incubation study investigated four biochar feedstocks: spruce bark biochar at 550 °C (SB550), hardwood biochar (75% sugar maple) at 500 °C (HW500), sawdust (fir/spruce) biochar at 427 °C (FS427), and softwood biochar at 500 °C (SW500), to identify the most effective nitrate adsorbent. Scanning electron microscopy (SEM) and Fourier transform infrared spectroscopy (FT-IR) were employed to analyze biochar morphology and surface functional groups. Adsorption isotherms were modeled using the Langmuir and Freundlich equations. The results indicated that surface functional groups, such as aromatic C=C stretching and bending, aromatic C–H bending, and phenolic O–H bending, play crucial roles in enhancing electrostatic attraction and, consequently, the nitrate adsorption capacity of biochar. The equilibrium adsorption data from this study fit well with both the Langmuir and Freundlich isotherm models. Among the four biochar types tested, SB550 exhibited the highest nitrate adsorption capacity, with a maximum of 184 mg/g. The adsorption data showed excellent conformity to the Langmuir and Freundlich models, with correlation coefficients (R2) exceeding 0.987 for all biochar types. These findings highlight the high accuracy of these models in predicting nitrate adsorption capacities. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of the biochars, including (<b>a</b>) SB550, (<b>b</b>) HW500, (<b>c</b>) FS427, and (<b>d</b>) SW500.</p>
Full article ">Figure 2
<p>FT-IR spectra of biochars before and after nitrate adsorption highlight key functional groups, such as aromatic C=C stretching and phenolic O–H bending, which contribute to nitrate adsorption efficiency. The spectra include the following biochars: (<b>a</b>) SB550, (<b>b</b>) HW500, (<b>c</b>) FS427, and (<b>d</b>) SW500.</p>
Full article ">Figure 3
<p>Langmuir and Freundlich isotherms models for nitrate adsorption onto biochar, including (<b>a</b>) SB550, (<b>b</b>) HW500, (<b>c</b>) S427, and (<b>d</b>) SW500.</p>
Full article ">Figure 4
<p>Nitrate removal rates from aqueous solutions by different biochar types, including (<b>a</b>) SB550, (<b>b</b>) HW500, (<b>c</b>) FS427, and (<b>d</b>) SW500.</p>
Full article ">Figure 5
<p>Effect of initial nitrate concentration on nitrate adsorption (mg/g) and nitrate removal rate (%) by biochar (<b>a</b>) SB550, (<b>b</b>) HW500, (<b>c</b>) FS427, and (<b>d</b>) SW500.</p>
Full article ">Figure 6
<p>Effect of initial solution pH and the equilibrium solution pH on nitrate removal rate, (<b>a</b>) SB550, (<b>b</b>) HW500, (<b>c</b>) FS427, and (<b>d</b>) SW500.</p>
Full article ">
26 pages, 7277 KiB  
Article
Field-Level Classification of Winter Catch Crops Using Sentinel-2 Time Series: Model Comparison and Transferability
by Kato Vanpoucke, Stien Heremans, Emily Buls and Ben Somers
Remote Sens. 2024, 16(24), 4620; https://doi.org/10.3390/rs16244620 - 10 Dec 2024
Viewed by 950
Abstract
Winter catch crops are promoted in the European Union under the Common Agricultural Policy to improve soil health and reduce nitrate leaching from agricultural fields. Currently, Member States often monitor farmers’ adoption through on-site inspections for a limited subset of parcels. Because of [...] Read more.
Winter catch crops are promoted in the European Union under the Common Agricultural Policy to improve soil health and reduce nitrate leaching from agricultural fields. Currently, Member States often monitor farmers’ adoption through on-site inspections for a limited subset of parcels. Because of its potential for region-wide coverage, this study investigates the potential of Sentinel-2 satellite time series to classify catch crops at the field level in Flanders (Belgium). The first objective was to classify catch crops and identify the optimal model and time-series input for this task. The second objective was to apply these findings in a real-world scenario, aiming to provide reliable early-season predictions in a separate target year, testing early-season performance and temporal transferability. The following three models were compared: Random Forest (RF), Time Series Forest (TSF), and a One-Dimensional Convolutional Neural Network (1D-CNN). The results showed that, with a limited field-based training dataset, RF produced the most robust results across different time-series inputs, achieving a median F1-score of >88% on the best dataset. Additionally, the early-season performance of the models was delayed in the target year, reaching the F1-score threshold of 85% at least one month later in the season compared to the training years, with large timing differences between the models. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Highlights of the applied workflow. Three experiments were conducted: The first and second experiments aimed to study the influence of temporal resolution and input features on the classification, respectively on the classification. The third experiment evaluated the models’ performance early in the winter season and their temporal transferability.</p>
Full article ">Figure 2
<p>The location of the study region in Europe; Flanders is colored in cyan. Administrative borders: European Environment Agency (<a href="http://www.eea.europa.eu/datahub" target="_blank">www.eea.europa.eu/datahub</a> (accessed on 22 November 2024)).</p>
Full article ">Figure 3
<p>Locations of the 6881 field observations in the study area: (<b>A</b>) colored by year of observation where the color indicates the year in which the winter year started (e.g., 2022 are observations in the winter year 2022–2023) and (<b>B</b>) the field observations located on an agricultural land use map. Land use: Department of Environment and Spatial Development (<a href="http://www.vlaanderen.be/datavindplaats" target="_blank">www.vlaanderen.be/datavindplaats</a> (accessed on 22 November 2024)).</p>
Full article ">Figure 4
<p>Overview of (<b>A</b>) the classification hierarchy and (<b>B</b>) the label distribution on both the first (left) and second level (right) of the hierarchy.</p>
Full article ">Figure 5
<p>Network architecture of the One-Dimensional Convolutional Neural Network (1D-CNN). The convolutional blocks each comprise a convolutional layer (number of filters, 8, 16, 32), a batch normalization, and a Rectified Linear Unit (ReLU) activation function. The fully connected blocks each consist of a linear layer, a batch normalization, and ReLU activation function.</p>
Full article ">Figure 6
<p>Average NDVI time series at (<b>A</b>) daily, (<b>B</b>) dekadal, and (<b>C</b>) monthly temporal resolution for each of the classes. Linear interpolation was used to create continuous time series.</p>
Full article ">Figure 7
<p>Violin plots for the F1-scores of the first experiment given for three temporal resolutions and three different models (Random Forest or RF, Time Series Forest or TSF, and a One-Dimensional Convolutional Neural Network or 1D-CNN). Within the violin plots, the boxplots are depicted in black.</p>
Full article ">Figure 8
<p>Violin plots for the F1-scores of the second experiment given for three temporal resolutions and three different models (Random Forest or RF, Time Series Forest or TSF, and One-Dimensional Convolutional Neural Network or 1D-CNN). (<b>A</b>) Part A of the second experiment and (<b>B</b>) part B of the second experiment. Within the violin plots, boxplots are depicted in black.</p>
Full article ">Figure 9
<p>Confusion matrices for the fine-label, end-of-season classification for three different models: (<b>A</b>) Random Forest (RF), (<b>B</b>) Time Series Forest (TSF), and (<b>C</b>) One-Dimensional Convolutional Neural Network (1D-CNN). The blue frames indicate the coarse-label classification groups.</p>
Full article ">Figure 10
<p>F1-scores for the third experiment, which was completed on daily time series with features NDVI-fAPAR-FCover for three models (RF, TSF, and 1D-CNN). (<b>A</b>) Results when trained and tested on the same years; (<b>B</b>) results when the models are applied on a target year. Month on the <span class="html-italic">x</span>-axis shows the end point of the time series: e.g., Dec represents a time series starting in August and ending in December. Shadowing indicates the standard deviation on the results from the diferent model runs. Note that the results for RF and TSF largely overlap in panel (<b>A</b>).</p>
Full article ">Figure 11
<p>Maps to illustrate the results of the Random Forest (RF) model on time series until April. Binary (<b>A</b>) ground truth and (<b>B</b>) classification results, fine-label (<b>C</b>) ground truth and (<b>D</b>) classification results. (<b>E</b>) Location of maps (<b>A</b>–<b>D</b>) in the study area. Land use: Department of Environment and Spatial Development (<a href="http://www.vlaanderen.be/datavindplaats" target="_blank">www.vlaanderen.be/datavindplaats</a> (accessed on 22 November 2024)).</p>
Full article ">Figure 11 Cont.
<p>Maps to illustrate the results of the Random Forest (RF) model on time series until April. Binary (<b>A</b>) ground truth and (<b>B</b>) classification results, fine-label (<b>C</b>) ground truth and (<b>D</b>) classification results. (<b>E</b>) Location of maps (<b>A</b>–<b>D</b>) in the study area. Land use: Department of Environment and Spatial Development (<a href="http://www.vlaanderen.be/datavindplaats" target="_blank">www.vlaanderen.be/datavindplaats</a> (accessed on 22 November 2024)).</p>
Full article ">Figure A1
<p>Comparison of the temporal profiles (NDVI, 1-day interval) of the different classification groups between the training data (2016–2021) and the validation data (2022). Shadowing indicates the standard deviation on each averaged profile. The group ‘Legumes’ is not visualized because of its absence in the validation dataset (2022).</p>
Full article ">Figure A1 Cont.
<p>Comparison of the temporal profiles (NDVI, 1-day interval) of the different classification groups between the training data (2016–2021) and the validation data (2022). Shadowing indicates the standard deviation on each averaged profile. The group ‘Legumes’ is not visualized because of its absence in the validation dataset (2022).</p>
Full article ">
19 pages, 10302 KiB  
Article
Investigation of Magnesium-Potassium Phosphates as Potential Nuclear Waste Form for the Immobilization of Minor Actinides
by Hans-Conrad zur Loye, Petr Vecernik, Monika Kiselova, Vlastislav Kašpar, Hana Korenkova, Vlastimil Miller, Petr Bezdicka, Jan Šubrt, Natalija Murafa, Volodymyr Shkuropatenko and Sergey Sayenko
Inorganics 2024, 12(12), 311; https://doi.org/10.3390/inorganics12120311 - 28 Nov 2024
Viewed by 707
Abstract
Several recent studies have evaluated technologies of spent nuclear fuel processing specifically for solidifying transuranic (TRU) waste as a by-product of fission. Of the TRU group, plutonium and the minor actinides will be responsible for the bulk of the radiotoxicity and heat generation [...] Read more.
Several recent studies have evaluated technologies of spent nuclear fuel processing specifically for solidifying transuranic (TRU) waste as a by-product of fission. Of the TRU group, plutonium and the minor actinides will be responsible for the bulk of the radiotoxicity and heat generation of spent nuclear fuel in the long term (300 to 20,000 years). In this study, we investigated magnesium potassium phosphate (MKP)-based compounds as host waste forms for the encapsulation of inactive trivalent Nd and Sm as analogues of the minor trivalent actinides, Am and Cm. Waste forms were fabricated under ambient atmospheric conditions by adding 5 wt.% of substances containing Nd or Sm via the following two routes: powder oxides and aqueous solutions of nitrate salts. Waste form performance was established using strength and aqueous medium leaching tests of MKP-based specimens. The MKP materials were analyzed by X-ray diffraction (XRD), scanning and transmission electron microscopy (SEM and TEM), energy-dispersive X-ray spectroscopy (EDS), and Raman spectroscopy. The waste forms exhibited a compressive strength of ≥30 MPa and were durable in an aqueous environment. The leachability indices for Nd and Sm, as per the ANS 16.1 procedure, were 19.55–19.78 and 19.74–19.89, respectively, which satisfy the acceptable criteria (>6). The results of the present room temperature leaching study suggest that MKPs can be effectively used as a host material to immobilize actinides (Am and Cm) contained in TRU waste. Full article
(This article belongs to the Section Inorganic Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Compressive strength of MKP specimens with different M/P ratios.</p>
Full article ">Figure 2
<p>XRD pattern of pure MKP specimen.</p>
Full article ">Figure 3
<p>XRD patterns of MKP specimen with Sm<sub>2</sub>O<sub>3</sub> and Nd<sub>2</sub>O<sub>3</sub> additives and pure MKP specimen.</p>
Full article ">Figure 4
<p>XRD patterns of MKP specimen with Sm(NO<sub>3</sub>)<sub>3</sub>·6H<sub>2</sub>O and Nd(NO<sub>3</sub>)<sub>3</sub>·6H<sub>2</sub>O additives and pure MKP specimen.</p>
Full article ">Figure 5
<p>Raman spectra of obtained MKP-based materials: MKP/Nd—with Nd<sub>2</sub>O<sub>3</sub> additive, MKP/Sm—with Sm<sub>2</sub>O<sub>3</sub> additive, MKP/SmN—with Sm(NO<sub>3</sub>)<sub>3</sub>·6H<sub>2</sub>O additive.</p>
Full article ">Figure 6
<p>SEM image of MKP specimen with addition of Nd<sub>2</sub>O<sub>3</sub> (<b>a</b>) and Sm<sub>2</sub>O<sub>3</sub> (<b>b</b>).</p>
Full article ">Figure 7
<p>SEM high magnification image and EDS data of MKP specimen with adding of Nd<sub>2</sub>O<sub>3</sub>: (<b>a</b>) grey matrix, (<b>b</b>) light-colored agglomerate.</p>
Full article ">Figure 8
<p>SEM elemental mapping for MKP specimen with Nd<sub>2</sub>O<sub>3</sub> additive.</p>
Full article ">Figure 9
<p>SEM image for MKP specimen prepared via the addition of an aqueous solution of Sm(NO<sub>3</sub>)<sub>3</sub>·6H<sub>2</sub>O, the yellow box is the area for EDS analysis.</p>
Full article ">Figure 10
<p>SEM elemental mapping for MKP specimen prepared via the addition of an aqueous solution of Sm(NO<sub>3</sub>)<sub>3</sub>·6H<sub>2</sub>O additive.</p>
Full article ">Figure 11
<p>TEM image of a particle of MKP specimen prepared via the addition of an aqueous solution of Nd(NO<sub>3</sub>)<sub>3</sub>·6H<sub>2</sub>O and EDS spectra for areas (<b>a</b>,<b>b</b>).</p>
Full article ">Figure 12
<p>The pH value of hardened (28 d) MKP specimens during leaching test.</p>
Full article ">Figure 13
<p>Cumulative concentrations for Nd and Sm.</p>
Full article ">
19 pages, 2136 KiB  
Article
The Application of Straw Return with Nitrogen Fertilizer Increases Rice Yield in Saline–Sodic Soils by Regulating Rice Organ Ion Concentrations and Soil Leaching Parameters
by Tianqi Bai, Cheng Ran, Qiyue Ma, Yue Miao, Shangze Li, Heng Lan, Xinru Li, Qinlian Chen, Qiang Zhang and Xiwen Shao
Agronomy 2024, 14(12), 2807; https://doi.org/10.3390/agronomy14122807 - 26 Nov 2024
Cited by 1 | Viewed by 677
Abstract
Soil salinization is a severe environmental problem that restricts crop productivity. Straw amendment could increase the fertility of saline–sodic soils by improving soil physical properties and carbon sequestration; however, the chemical mechanism of saline soil improvement via straw reclamation is not clear. This [...] Read more.
Soil salinization is a severe environmental problem that restricts crop productivity. Straw amendment could increase the fertility of saline–sodic soils by improving soil physical properties and carbon sequestration; however, the chemical mechanism of saline soil improvement via straw reclamation is not clear. This study aimed to investigate the effects of straw return with nitrogen fertilizer on soil leaching characteristics, rice organ ion concentrations, and yield. Therefore, a soil column leaching experiment was conducted in 2021 in Baicheng, Jilin Province, using two straw application rate treatments (0 and 8 t hm−2) and three nitrogen application rate treatments (0, 180, and 360 kg hm−2). The results revealed the following: 1. The combination of straw return and nitrogen fertilizer significantly increased the soil leachate volume, leachate pH, Na+ concentration, and Na+/K+ ratio, thereby reducing Na+ stress on rice; 2. The application of nitrogen fertilizer during straw return effectively minimized soil nitrogen loss by lowering the ammonium and nitrate nitrogen concentrations in the soil leachate; 3. This combination also reduced plant Na+ concentrations while increasing plant K+ concentrations, thus improving the Na+/K+ ratio in the plants; 4. Straw return with nitrogen fertilizer significantly enhanced rice yield, which increased with higher nitrogen application rates. In summary, the integration of straw return with nitrogen fertilizer not only regulates rice salinity tolerance but also boosts rice yield, presenting a novel approach for improving saline–sodic soils. Full article
(This article belongs to the Section Soil and Plant Nutrition)
Show Figures

Figure 1

Figure 1
<p>Monthly average temperature (°C) and monthly total precipitation (mm) at the test site. Note: MP: monthly total precipitation; MT: monthly average temperature.</p>
Full article ">Figure 2
<p>Percolation columns used for leaching experiments: illustration of the cross-section.</p>
Full article ">Figure 3
<p>Effect of straw return with nitrogen fertilizer application on the soil leachate volume. Note: The S0 treatment denotes the absence of straw return, while the S treatment indicates the presence of straw return. The designations N0, N180, and N360 correspond to nitrogen fertilizer application rates of 0, 180, and 360 kg hm<sup>−2</sup>, respectively. Different letters signify statistically significant differences among the nitrogen fertilizer application rates (<span class="html-italic">p</span> &lt; 0.05). The annotations ** represent significant at <span class="html-italic">p</span> &lt; 0.01 levels, respectively. Panel (<b>A</b>) depicts the volume dynamics of the leachate in 2021, while panel (<b>B</b>) presents the total volume of the leachate in 2021.</p>
Full article ">Figure 4
<p>Effect of straw return with nitrogen fertilizer application on the pH of soil leachate.</p>
Full article ">Figure 5
<p>Effect of straw return with nitrogen fertilizer application on the Na<sup>+</sup> and K<sup>+</sup> concentrations and Na<sup>+</sup>/K<sup>+</sup> ratio in the soil leachate. Note: panel (<b>A</b>) depicts the Na<sup>+</sup> concentration in the leachate, panel (<b>B</b>) depicts the K<sup>+</sup> concentration in the leachate, and panel (<b>C</b>) depicts the Na<sup>+</sup>/K<sup>+</sup> ratio in the leachate.</p>
Full article ">Figure 6
<p>Effect of straw return with nitrogen fertilizer application on the NO<sub>3</sub><sup>−</sup>-N and NH<sub>4</sub><sup>+</sup>-N concentrations in the soil leachate. Note: panel (<b>A</b>) depicts the NO<sub>3</sub><sup>−</sup>-N concentration dynamics in the leachate, while panel (<b>B</b>) depicts the NH<sub>4</sub><sup>+</sup>-N concentration dynamics in the leachate.</p>
Full article ">Figure 7
<p>Effect of straw return with nitrogen fertilizer application on Na<sup>+</sup> and K<sup>+</sup> contents, and Na<sup>+</sup>/K<sup>+</sup> ratio in different rice organs at the maturity stage. Note: panels (<b>A</b>,<b>B</b>) depict the Na<sup>+</sup> content in different organs, panels (<b>C</b>,<b>D</b>) depict the K<sup>+</sup> content in different organs, and panels (<b>E</b>,<b>F</b>) depict Na<sup>+</sup>/K<sup>+</sup> ratio in different organs. S0: straw removal; S: straw return. Different letters in the same column indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
57 pages, 3124 KiB  
Review
An Extensive Review of Leaching Models for the Forecasting and Integrated Management of Surface and Groundwater Quality
by Stephanos D. V. Giakoumatos, Christina Siontorou and Dimitrios Sidiras
Water 2024, 16(23), 3348; https://doi.org/10.3390/w16233348 - 21 Nov 2024
Viewed by 1327
Abstract
The present study reviews leachate models useful for proactive and rehab actions to safeguard surface and subsurface soft water, which have become even more scarce. Integrated management plans of water basins are of crucial importance since intensively cultivated areas are adding huge quantities [...] Read more.
The present study reviews leachate models useful for proactive and rehab actions to safeguard surface and subsurface soft water, which have become even more scarce. Integrated management plans of water basins are of crucial importance since intensively cultivated areas are adding huge quantities of fertilizers to the soil, affecting surface water basins and groundwater. Aquifers are progressively being nitrified on account of the nitrogen-based fertilizer surplus, rendering water for human consumption not potable. Well-tested solute leaching models, standalone or part of a model package, provide rapid site-specific estimates of the leaching potential of chemical agents, mostly nitrates, below the root zone of crops and the impact of leaching toward groundwater. Most of the models examined were process-based or conceptual approaches. Nonetheless, empirical prediction models, though rather simplistic and therefore not preferrable, demonstrate certain advantages, such as less demanding extensive calibration database information requirements, which in many cases are unavailable, not to mention a stochastic approach and the involvement of artificial intelligence (AI). Models were categorized according to the porous medium and agents to be monitored. Integrated packages of nutrient models are irreplaceable elements for extensive catchments to monitor the terrestrial nitrogen-balanced cycle and to contribute to policy making as regards soft water management. Full article
(This article belongs to the Special Issue Soil-Groundwater Pollution Investigations)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Water, carbon/nitrogen carbon cycle, ambient pathways including corresponding pools in soil, surface, and atmosphere and brief mechanisms of transformation.</p>
Full article ">Figure 2
<p>Schematic of best soil management practice by using HYDRUS 1-D soil-soluble water simulation adopted. HYDRUS obtains information from various submodules/processed data based on simulation mathematical modeling with proper feedback over interrelated aspects, i.e., dynamic ongoing phenomena (e.g., nitrification/denitrification/hydraulic conductivity, climatic data, physicochemical interactions), and calibration as self-correction of each individual step.</p>
Full article ">Figure 3
<p>Operational modules of DAISY, a nitrogen–water balance and transport model in croplands, with processed data interconnection/interrelation for complementary reasons.</p>
Full article ">Figure 4
<p>Operational modules of the APSIM crop growth model over individual fields of interest include soil–water interaction and nitrogen/crop/canopy dynamics.</p>
Full article ">Figure 5
<p>Dendritic variants of AI techniques over soil water leachate modeling in the manuscript.</p>
Full article ">Figure 6
<p>Model advantages, advantages/drawbacks, and drawbacks as given from the relative literature. The numbers in parentheses denote the countries where soil monitoring projects were implemented.</p>
Full article ">
16 pages, 3427 KiB  
Systematic Review
Slow-Release Fertilisers Control N Losses but Negatively Impact on Agronomic Performances of Pasture: Evidence from a Meta-Analysis
by Gunaratnam Abhiram
Nitrogen 2024, 5(4), 1058-1073; https://doi.org/10.3390/nitrogen5040068 - 17 Nov 2024
Viewed by 679
Abstract
High nitrogen (N) losses and low nitrogen utilisation efficiency (NUE) of conventional-nitrogen fertilisers (CNFs) are due to a mismatch between N-delivery and plant demand; thus, slow-release N fertilisers (SRNFs) are designed to improve the match. A quantitative synthesis is lacking to provide the [...] Read more.
High nitrogen (N) losses and low nitrogen utilisation efficiency (NUE) of conventional-nitrogen fertilisers (CNFs) are due to a mismatch between N-delivery and plant demand; thus, slow-release N fertilisers (SRNFs) are designed to improve the match. A quantitative synthesis is lacking to provide the overall assessment of SRNFs on pasture. This meta-analysis analyses application rate and type of SRNFs on N losses and agronomic performances with 65 data points from 14 studies in seven countries. Standardized mean difference of SRNFs for nitrate leaching losses and N2O emission were −0.87 and −0.69, respectively, indicating their effectiveness in controlling losses. Undesirably, SRNFs had a more negative impact on dry matter (DM) yield and NUE than CNFs. Subgroup analysis showed that SRNF type and application rate had an impact on all tested parameters. The biodegradable coating-type of SRNF outperformed other types in controlling N losses and improving agronomic performances. High application rates (>100 kg N ha−1) of SRNFs are more effective in controlling N losses. In conclusion, SRNFs are more conducive to controlling N losses, but they showed a negative impact on yield and NUE in pasture. Further studies are recommended to assess the efficacy of SRNFs developed using advanced technologies to understand their impact on pastoral agriculture. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram for inclusion criteria of articles for this systematic review and meta-analysis (PRISMA) [<a href="#B32-nitrogen-05-00068" class="html-bibr">32</a>].</p>
Full article ">Figure 2
<p>The summary of the reported parameters from each study included in this meta-analysis [<a href="#B10-nitrogen-05-00068" class="html-bibr">10</a>,<a href="#B24-nitrogen-05-00068" class="html-bibr">24</a>,<a href="#B25-nitrogen-05-00068" class="html-bibr">25</a>,<a href="#B27-nitrogen-05-00068" class="html-bibr">27</a>,<a href="#B35-nitrogen-05-00068" class="html-bibr">35</a>,<a href="#B36-nitrogen-05-00068" class="html-bibr">36</a>,<a href="#B37-nitrogen-05-00068" class="html-bibr">37</a>,<a href="#B38-nitrogen-05-00068" class="html-bibr">38</a>,<a href="#B39-nitrogen-05-00068" class="html-bibr">39</a>,<a href="#B40-nitrogen-05-00068" class="html-bibr">40</a>,<a href="#B41-nitrogen-05-00068" class="html-bibr">41</a>,<a href="#B42-nitrogen-05-00068" class="html-bibr">42</a>,<a href="#B43-nitrogen-05-00068" class="html-bibr">43</a>,<a href="#B44-nitrogen-05-00068" class="html-bibr">44</a>].</p>
Full article ">Figure 3
<p>The nitrate leaching losses of conventional nitrogen fertilisers (CNFs) and slow-release nitrogen fertilisers (SRNFs) based on (<b>a</b>) SRNF types (PC—polymer coating, BC—biodegradable coating, IC—inorganic coating, and PCH—polymer chain), (<b>b</b>) fertiliser application rates (kg N/ha) and (<b>c</b>) overall studies. SMD stands for standard mean difference. Numbers next to range graph indicate the number of studies included for analysis.</p>
Full article ">Figure 4
<p>The correlation between effect size (standardized mean difference: SMD) and application rate of SRNFs for (<b>a</b>) nitrate leaching losses, (<b>b</b>) ammonium leaching losses, (<b>c</b>) N<sub>2</sub>O emission, (<b>d</b>) dry matter yield, (<b>e</b>) nitrogen utilisation efficiency (NUE) and (<b>f</b>) herbage nitrogen. Dark shade and light shades indicate a 95% confidence interval and a 95% prediction level, respectively.</p>
Full article ">Figure 5
<p>The effect of SRNF on ammonium leaching losses. SRNF and CNF refer to slow-release nitrogen fertiliser and conventional nitrogen fertiliser, respectively [<a href="#B10-nitrogen-05-00068" class="html-bibr">10</a>,<a href="#B27-nitrogen-05-00068" class="html-bibr">27</a>,<a href="#B36-nitrogen-05-00068" class="html-bibr">36</a>].</p>
Full article ">Figure 6
<p>The effect of SRNF on N<sub>2</sub>O emission. SRNF and CNF refer to slow-release nitrogen fertiliser and conventional nitrogen fertiliser, respectively [<a href="#B10-nitrogen-05-00068" class="html-bibr">10</a>,<a href="#B27-nitrogen-05-00068" class="html-bibr">27</a>,<a href="#B42-nitrogen-05-00068" class="html-bibr">42</a>,<a href="#B44-nitrogen-05-00068" class="html-bibr">44</a>].</p>
Full article ">Figure 7
<p>Dry matter yield of conventional nitrogen fertilisers (CNFs) and slow-release nitrogen fertilisers (SRNFs) based on (<b>a</b>) SRNF types (PC—polymer coating, BC—biodegradable coating, IC—inorganic coating, and PCH—polymer chain), (<b>b</b>) fertiliser application rates (kg N/ha), and (<b>c</b>) overall studies. SMD stands for standard mean difference. Numbers next to range graph indicate the number of studies included for analysis.</p>
Full article ">Figure 8
<p>The plant nutrient demand (PND) and nutrient delivery by slow-release nitrogen fertiliser (NDSRNF) for (<b>a</b>) other crops and (<b>b</b>) pasture. The blue shaded area shows the PND of pasture during continuous grass grazing.</p>
Full article ">Figure 9
<p>Herbage nitrogen (HN) in a pasture of conventional nitrogen fertilisers (CNFs) and slow-release nitrogen fertilisers (SRNFs) based on (<b>a</b>) SRNF types (PC—polymer coating, BC—biodegradable coating, IC—inorganic coating, and PCH—polymer chain), (<b>b</b>) fertiliser application rates (kg N/ha), and (<b>c</b>) overall studies. SMD stands for standard mean difference. Numbers next to range graph indicate the number of studies included for analysis.</p>
Full article ">Figure 10
<p>Nitrogen utilisation efficiency (NUE) in a pasture of conventional nitrogen fertilisers (CNFs) and slow-release nitrogen fertilisers (SRNFs) based on (<b>a</b>) SRNF types (PC—polymer coating, BC—biodegradable coating, IC—inorganic coating, and PCH—polymer chain), (<b>b</b>) fertiliser application rates (kg N/ha), and (<b>c</b>) overall studies. SMD stands for standard mean difference. Numbers next to range graph indicate the number of studies included for analysis.</p>
Full article ">
16 pages, 3168 KiB  
Article
Impact of Subsurface Drainage System Design on Nitrate Loss and Crop Production
by Soonho Hwang, Shailendra Singh, Rabin Bhattarai, Hanseok Jeong and Richard A. Cooke
Appl. Sci. 2024, 14(22), 10180; https://doi.org/10.3390/app142210180 - 6 Nov 2024
Viewed by 728
Abstract
Subsurface (or tile) drainage offers a valuable solution for enhancing crop productivity in poorly drained soils. However, this practice is also associated with significant nutrient leaching, which can contribute to water quality problems at the regional scale. This research presents the findings from [...] Read more.
Subsurface (or tile) drainage offers a valuable solution for enhancing crop productivity in poorly drained soils. However, this practice is also associated with significant nutrient leaching, which can contribute to water quality problems at the regional scale. This research presents the findings from a 4-year tile depth and spacing study in central Illinois that included three drain spacings (12.2, 18.3, and 24.4 m) and two drain depths (0.8 and 1.1 m) implemented in six plots under the corn and soybean rotation system (plots CS-1 and CS-3: 12.2 m spacing and 1.1 m depth, plots CS-2 and CS-4: 24.4 m spacing and 1.1 m depth, and plots CS-5 and CS-6 18.3 m spacing and 0.8 m depth). Our observations indicate that drain flow and NO3-N losses were higher in plots with narrower drain spacings, while plots with wider drain spacing showed reduced drain flow and NO3-N losses. Specifically, plots set up with drain spacings of 18.3 m and 24.4 m showed significant reductions in drain flow compared to plots featuring a 12.2 m drain spacing. Likewise, plots characterized by 18.3 m and 24.4 m drain spacings (except CS-4) showed better NO3-N retention and lower leaching losses than those with 12.2 m spacing (CS-1 and CS-3). Crop yield results over a 3-year period indicated that CS-2 (wider spacing plot) showed the highest productivity, with up to 13.6% higher yield compared to other plots. Furthermore, when comparing plots with the same drainage designs, CS-2 and CS-4 showed 5.1% to 2.6% higher corn yield (3-year average) compared to CS-1 and CS-3, and CS-5 and CS-6, respectively. Overall, a wider drainage system showed the capacity to export lower nutrient levels while concurrently enhancing productivity. These findings represent that optimizing tile drainage systems can effectively reduce nitrate losses while increasing crop productivity. Full article
Show Figures

Figure 1

Figure 1
<p>Field layout and drainage system design of CS-field.</p>
Full article ">Figure 2
<p>Field boundaries and surface flow line of CS-field with Lidar DEM (<b>left</b>) and Topographic Position Index (TPI) map (<b>right</b>).</p>
Full article ">Figure 3
<p>Cumulative daily tile flow at each site from 2018 to 2022.</p>
Full article ">Figure 4
<p>Drainage flow differences between plots were analyzed using Tukey’s test (<span class="html-italic">p</span> &lt; 0.05). CS-1 and CS-4, and CS-2, CS-4, and CS-5 showed no significant differences.</p>
Full article ">Figure 5
<p>Cumulative daily nitrate-N loss at each site from 2018 to 2022.</p>
Full article ">Figure 6
<p>NO<sub>3</sub>-N load differences between plots were analyzed using Tukey’s test (<span class="html-italic">p</span> &lt; 0.05). CS-1, CS-3, and CS-4 and CS-1 and CS-5, CS-2 and CS-5 and CS-2 and CS-6 showed no significant differences.</p>
Full article ">Figure 7
<p>Nitrate concentration of soil below 0–15 cm and 15–30 cm. The nitrate concentration in the soil exhibits variation between depths and years across the plots.</p>
Full article ">
11 pages, 586 KiB  
Article
Newly Established Blueberry Plants: The Role of Inorganic Nitrogen Forms in Nitrogen and Calcium Absorption
by María Ignacia Arias, Adriana Nario, Krystel Rojas, Poulette Blanc and Claudia Bonomelli
Horticulturae 2024, 10(11), 1168; https://doi.org/10.3390/horticulturae10111168 - 4 Nov 2024
Viewed by 1025
Abstract
Efficient nitrogen (N) management is crucial for maximizing the growth of young blueberry plants (Vaccinium corymbosum). This study evaluates the effects of the N fertilization form (ammonium, NH4+; nitrate, or NO3) and application timing on [...] Read more.
Efficient nitrogen (N) management is crucial for maximizing the growth of young blueberry plants (Vaccinium corymbosum). This study evaluates the effects of the N fertilization form (ammonium, NH4+; nitrate, or NO3) and application timing on the blueberries’ establishment, N and Ca absorption, and N distribution. The experiment was conducted in the southern hemisphere, in Chile, from October 2023 to January 2024. Six-month-old blueberry cv. Blue Ribbon plants were cultivated in pots. NH4+ and NO3 were used as full or split-dose applications using the 15N isotopic dilution technique. Plant leaves, stems, root growth, and biomass, as well as their N and Ca contents, were measured. Our results showed that 90 days after nitrogen application, blueberry plants obtained the lowest biomass in their leaves, stems, and roots when NO3 was applied in T1 or T1T2. The same pattern was observed for N and Ca contents, hence for N recovery. During the first period (T1) of application, heavy rain (100 mm) was registered over the course of a few days and caused leaching. Therefore, applying nitrate to young blueberry plants cultivated in areas with spring rainfall and low temperatures would not be recommended because the leaching losses and lower growth conditions, such as low temperatures and high precipitation, led to reduced transpiration, resulting in lower calcium and nitrogen contents. These confirm that N fertilization management (form and timing) can ensure a better establishment for young blueberry plants, optimizing their growth and sustainable production by minimizing nitrogen losses. Full article
(This article belongs to the Special Issue The Effects of Fertilizers on Fruit Production)
Show Figures

Figure 1

Figure 1
<p><sup>15</sup>N recovery distribution (%) in plants of blueberry (<span class="html-italic">Vaccinium corymbosum</span> L.) and <sup>15</sup>N residual (%) in substrate under ammonium (NH<sub>4</sub><sup>+</sup>) and nitrate (NO<sub>3</sub><sup>−</sup>) fertilization across four application timings (T1, T2, T1*T2, and T1T2*).</p>
Full article ">
22 pages, 3961 KiB  
Article
Assessing Nitrogen Fertilization in Processing Pepper: Critical Nitrogen Curve, Yield Response, and Crop Development
by Jose Maria Vadillo, Carlos Campillo, Valme González and Henar Prieto
Horticulturae 2024, 10(11), 1141; https://doi.org/10.3390/horticulturae10111141 - 25 Oct 2024
Cited by 1 | Viewed by 763
Abstract
Groundwater pollution in intensive horticultural areas is becoming an increasingly important problem. Over-fertilization of these crops, combined with poor irrigation management, leads to groundwater contamination through leaching. Previous research on the effect of N on sweet peppers grown in greenhouses is abundant, but [...] Read more.
Groundwater pollution in intensive horticultural areas is becoming an increasingly important problem. Over-fertilization of these crops, combined with poor irrigation management, leads to groundwater contamination through leaching. Previous research on the effect of N on sweet peppers grown in greenhouses is abundant, but data on outdoor cultivation, especially considering variety and site influences, are lacking. Therefore, this study evaluates nitrogen (N) fertilization in open-field processing-pepper crop in Extremadura, Spain to mitigate this environmental impact. Field trials were conducted in 2020, 2021, and 2022 to determine the optimum N fertilizer rate for processing peppers, with the aim of reducing environmental impacts such as nitrate leaching while maintaining crop yields. The trial consisted of applying different N doses, 0, 60, 120, and 180 kg N/ha in 2020 and 2021 and 0, 100, and 300 kg N/ha in 2022. There were four replications of each treatment, arranged in randomized blocks. Measurements included crop yield, biomass, intercepted photosynthetically active radiation (PAR), and canopy cover. The study also developed a critical nitrogen curve (CNC) to determine the minimum N concentration required for optimal growth. The commercial yield results showed that there were no significant differences between the two treatments with higher N inputs in the three years; therefore, the application of more than 120 kg N/ha did not significantly increase yield. Nitrogen-free treatments resulted in earlier fruit maturity, concentrating the harvest and reducing waste. In addition, excessive N application led to environmental problems such as groundwater contamination due to nitrate leaching. The study concludes that outdoor pepper crops in this region can achieve optimal yields with lower N rates (around 120 kg N/ha) compared to current practices, taking into account that initial soil N values were higher than 100 kg N/ha, thereby reducing environmental risks and fertilizer costs. It also established relationships between biomass, canopy cover, and N uptake to improve fertilization strategies. These data support future crop modeling and sustainable fertilization practices. Full article
(This article belongs to the Special Issue Irrigation and Fertilization Management in Horticultural Production)
Show Figures

Figure 1

Figure 1
<p>Zenithal photo of the crop performance ‘reclassification method’.</p>
Full article ">Figure 2
<p>Daily ETo evolution of the 2020, 2021, and 2022 campaigns.</p>
Full article ">Figure 3
<p>Seasonal evolution of cumulative dry matter (DM) as a function of growing degree days (GDDs) under non-nitrogen limiting conditions for the years 2020, 2021, and 2022, (N3 for 2020 and 2021 and N2 for 2022). Each point is the average of 4 measurements in the elementary replicate plots.</p>
Full article ">Figure 4
<p>Relationship between the percentage of crop cover (f<sub>CC</sub>) and fraction of photosynthetically active radiation (PAR) intercepted by the vegetation cover (fi-PAR) obtained in the years 2020 and 2021. Each point corresponds to the measurements made on the same section of crop.</p>
Full article ">Figure 5
<p>Evolution of fi-PAR along the crop cycle expressed in cumulative growing degree days (GDDs) for 2020, 2021, and 2022 under non-nitrogen limiting conditions (N3 for 2020 and 2021 and N2 for 2022).</p>
Full article ">Figure 6
<p>Relationship between cumulative dry matter (DM) along the crop cycle and the fraction of photosynthetically active radiation (PAR) intercepted by the crop (fi-PAR) for the years 2020, 2021, and 2022. Each point is the average of 4 measurements in the elementary replicate plots.</p>
Full article ">Figure 7
<p>Critical N curve (CNC) for sweet pepper grown in open fields using total crop N content and total DM (continuous black line), CNC for C3 crops of Lemaire and Gastal (1997) [<a href="#B61-horticulturae-10-01141" class="html-bibr">61</a>] (dotted line), and CNC for greenhouse sweet pepper of Rodríguez et al. (2020) [<a href="#B6-horticulturae-10-01141" class="html-bibr">6</a>] (dashed line).</p>
Full article ">Figure 8
<p>Evolution of the fraction of PAR intercepted by the crop (fi-PAR) for the different N treatments in 2020 (<b>a</b>), 2021 (<b>b</b>), and 2022 (<b>c</b>). Each point is the average of 4 measurements in the elementary replicate plots in each treatment. Letters indicate whether there are significant differences between treatments in each sample for <span class="html-italic">p</span> ≤ 0.05. In samples where no letters appear, there were no significant differences between treatments.</p>
Full article ">Figure 9
<p>Evolution of flower production expressed in dry matter (DM) for the different N treatments in the (<b>a</b>) 2020 and (<b>b</b>) 2021 seasons. Each point is the average of 4 measurements in the elementary replicate plots in each treatment. Letters indicate whether there are significant differences between treatments in each sample with <span class="html-italic">p</span> ≤ 0.05. In samples where no letters appear, there were no significant differences between treatments for <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 10
<p>Total dry matter and fruit dry matter evolution in cumulative growing degree days (GDDs) for the different treatments in (<b>a</b>) 2020, (<b>b</b>) 2021, and (<b>c</b>) 2022. Each point is the average of 4 measurements in the elementary replicate plots in each treatment. Letters indicate whether there are significant differences in total dry matter between treatments in each sample with <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 11
<p>Commercial fruit production of seasons (<b>a</b>) 2020, (<b>b</b>) 2021, and (<b>c</b>) 2022 per harvest for treatments. Black letters indicate whether there are significant differences in each individual harvest, and red letters indicate significant differences in total yield.</p>
Full article ">Figure 12
<p>Evolution of the N content in relation to cumulative growing degree days (GDDs) for the different treatments and parts of the plant. (<b>a</b>,<b>b</b>) correspond to leaves in 2020 and 2022, respectively, (<b>c</b>,<b>d</b>) to stems in 2020 and 2022, respectively, and (<b>e</b>,<b>f</b>) to fruits in 2020 and 2022, respectively. Each point is the average of 4 measurements in the elementary replicate plots in each treatment. Letters indicate whether there are significant differences between treatments in each sample with <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 13
<p>Data on N content (%) in dry matter of aerial parts versus total dry matter (DM) weight over the crop cycle for each treatment in 2020 (<b>a</b>), 2021 (<b>b</b>), and 2022 (<b>c</b>), contrasted with the CNC generated in this work.</p>
Full article ">Figure 14
<p>Seasonal evolution of nitrogen (N) uptake in cumulative growing degree days (GDDs) for the different treatments in the years (<b>a</b>) 2020, (<b>b</b>) 2021, and (<b>c</b>) 2022. Each point is the average of 4 measurements in the elementary replicate plots in each treatment. Letters indicate whether there are significant differences between treatments in each sample with <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 15
<p>Relationship between total N uptake and total fresh-fruit yield for the years 2020, 2021, and 2022. Each point is the average of 4 measurements in the elementary replicate plots of each treatment.</p>
Full article ">Figure 16
<p>Increased mineralization throughout the year in response to crop N demand. Source: Tei et al. (2020) [<a href="#B75-horticulturae-10-01141" class="html-bibr">75</a>].</p>
Full article ">
Back to TopTop