[go: up one dir, main page]

 
 

Topic Editors

Department of Meteorology and Climatology, School of Geology, Aristotle University of Thessaloniki, Thessaloniki, Greece
Laboratory of Remote Sensing, Spectroscopy and GIS, School of Agriculture, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
Soil and Water Resources Insitute, Hellenic Agricultural Organization—Demeter, 57001 Thessaloniki, Greece

Advances in Crop Simulation Modelling

Abstract submission deadline
closed (15 April 2024)
Manuscript submission deadline
closed (15 June 2024)
Viewed by
15909

Topic Information

Dear Colleagues,

In contrast to statistical models, process-based crop simulation models consider dynamic interactions between environment, genotype, and management (including but not limited to agricultural water) factors, something that justifies their application in decision making, in the assessment of the impacts of climate change/variability, and management practices on productivity and environmental performance of alternative cropping systems, to promote better and sustainable agriculture. However, application of these models is often hindered by limited input data availability (such as climate, cultivar and soil characteristics, and management practices) for model calibration and testing and extensive computing time.

Although it is well recognized that the choice of model calibration strategy and incomplete/ poor in quality/not easily accessible input data have implications on the overall reliability of the crop model simulations, only few attempts have been made to quantify errors in crop simulation results related to the above-mentioned issues on a variety of spatial scales (from field to large area applications).

In this context, this Topic aims to highlight the challenges of producing locally relevant and climate informed results from crop simulation models, promoting this way their effective use, across various time frames (from seasonal to future climate change) for agriculture, under historical and future climate conditions.

Dr. Mavromatis Theodoros
Dr. Thomas Alexandridis
Dr. Vassilis Aschonitis
Topic Editors

Keywords

  • crop simulation models
  • calibration strategies
  • input availability
  • gridded data
  • climate models
  • cultivar and soil characteristics
  • uncertainty assessment
  • climate change scenarios
  • remote sensing data assimilation

Participating Journals

Journal Name Impact Factor CiteScore Launched Year First Decision (median) APC
Agronomy
agronomy
3.3 6.2 2011 15.5 Days CHF 2600
Climate
climate
3.0 5.5 2013 21.9 Days CHF 1800
Earth
earth
2.1 3.3 2020 21.7 Days CHF 1200
Remote Sensing
remotesensing
4.2 8.3 2009 24.7 Days CHF 2700
Water
water
3.0 5.8 2009 16.5 Days CHF 2600

Preprints.org is a multidiscipline platform providing preprint service that is dedicated to sharing your research from the start and empowering your research journey.

MDPI Topics is cooperating with Preprints.org and has built a direct connection between MDPI journals and Preprints.org. Authors are encouraged to enjoy the benefits by posting a preprint at Preprints.org prior to publication:

  1. Immediately share your ideas ahead of publication and establish your research priority;
  2. Protect your idea from being stolen with this time-stamped preprint article;
  3. Enhance the exposure and impact of your research;
  4. Receive feedback from your peers in advance;
  5. Have it indexed in Web of Science (Preprint Citation Index), Google Scholar, Crossref, SHARE, PrePubMed, Scilit and Europe PMC.

Published Papers (15 papers)

Order results
Result details
Journals
Select all
Export citation of selected articles as:
22 pages, 8325 KiB  
Article
Improving the Gross Primary Productivity Estimation by Simulating the Maximum Carboxylation Rate of Maize Using Leaf Age
by Xin Zhang, Shuai Wang, Weishu Wang, Yao Rong, Chenglong Zhang, Chaozi Wang and Zailin Huo
Remote Sens. 2024, 16(15), 2747; https://doi.org/10.3390/rs16152747 - 27 Jul 2024
Viewed by 335
Abstract
Although the maximum carboxylation rate (Vcmax) is an important parameter to calculate the photosynthesis rate for the terrestrial biosphere models (TBMs), current models could not satisfactorily estimate the Vcmax of a crop because the Vcmax is always changing during [...] Read more.
Although the maximum carboxylation rate (Vcmax) is an important parameter to calculate the photosynthesis rate for the terrestrial biosphere models (TBMs), current models could not satisfactorily estimate the Vcmax of a crop because the Vcmax is always changing during crop growth period. In this study, the Breathing Earth System Simulator (BESS) and light response curve (LRC) were combined to invert the time-continuous Vm25 (Vcmax normalized to 25 °C) using eddy covariance measurements and remote sensing data in five maize sites. Based on the inversion results, we propose a Two-stage linear model using leaf age to estimate crop Vm25. The leaf age can be readily calculated from the date of emergence, which is usually recorded or can be readily calculated from the leaf area index (LAI), which can be readily obtained from high spatiotemporal resolution remote sensing images. The Vm25 used to calibrate and validate our model was inversely solved by combining the BESS and LRC and using eddy covariance measurements and remote sensing data in five maize sites. Our Two-stage linear model (R2 = 0.71–0.88, RMSE = 5.40–7.54 μmol m−2 s−1) performed better than the original BESS (R2 = 0.01–0.67, RMSE = 13.25–18.93 μmol m−2 s−1) at capturing the seasonal variation in the Vm25 of all of the five maize sites. Our Two-stage linear model can also significantly improve the accuracy of maize gross primary productivity (GPP) at all of the five sites. The GPP estimated using our Two-stage linear model (underestimated by 0.85% on average) is significantly better than that estimated by the original BESS model (underestimated by 12.60% on average). Overall, our main contributions are as follows: (1) by using the BESS model instead of the BEPS model coupled with the LRC, the inversion of Vm25 took into account the photosynthesis process of C4 plants; (2) the maximum value of Vm25 (i.e., PeakVm25) during the growth and development of maize was calibrated; and (3) by using leaf age as a predictor of Vm25, we proposed a Two-stage linear model to calculate Vm25, which improved the estimation accuracy of GPP. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>The process of inverting V<sub>m25</sub> through coupling BESS and LRC with eddy covariance measurements and remotely sensed data, and the validation of the proposed Two-stage linear model [<a href="#B12-remotesensing-16-02747" class="html-bibr">12</a>].</p>
Full article ">Figure 2
<p>R<sup>2</sup> between V<sub>m25</sub> and LAI, DOY during (<b>a</b>) the ascending stage and (<b>b</b>) the descending stage. QR indicate quartiles. The top and bottom boundaries of boxes indicate the values of the 25% percentile (QR1) and 75% percentile (QR3) values, respectively. The lines at both ends represent the maximum range (QR3 + 1.5 IQR) and the minimum range (QR1–1.5 IQR). The lines in the middle of the box represent the median, and the black diamonds represent the mean.</p>
Full article ">Figure 3
<p>Schematic diagram of the Two-stage linear model.</p>
Full article ">Figure 4
<p>Comparing the V<sub>m25</sub> simulated by the Two-stage linear model (Simulated_V<sub>m25</sub>) and the V<sub>m25</sub> inversely solved by coupling the BESS and LRC using EC data (Inverted_V<sub>m25</sub>) at (<b>a</b>) US-Ne1, (<b>b</b>) US-Ne2, (<b>c</b>) US-Ne3, (<b>d</b>) Daman, (<b>e</b>) Fenzidi. Black dots represent values for the calibration data, red dots for the validation data.</p>
Full article ">Figure 5
<p>Comparison of V<sub>m25</sub> calculated by BESS (blue lines), BESS_P (green lines) and our Two-stage linear model (red lines) in simulating the time series of V<sub>m25</sub> inversely solved by coupling the BESS and LRC using EC data at the five flux sites: (<b>a</b>–<b>l</b>) show the verification results for US-Ne1; (<b>m</b>–<b>r</b>) show the verification results for US-Ne2; (<b>s</b>–<b>x</b>) show the verification results for US-Ne3; (<b>y</b>–<b>aa</b>) show the verification results for Daman; (<b>ab</b>–<b>ad</b>) show the verification results for Fenzidi.</p>
Full article ">Figure 6
<p>The GPP calculated by BESS (blue dots and trend lines), BESS_P (green dots and trend lines), and BESS_TL (red dots and trend lines) vs. flux site observed GPP (GPP_EC) at (<b>a</b>) US-Ne1, (<b>b</b>) US-Ne2, (<b>c</b>) US-Ne3, (<b>d</b>) Daman, (<b>e</b>) Fenzidi.</p>
Full article ">Figure 7
<p>The annual GPP estimation discrepancy of the BESS (blue lines), BESS_P (green lines) and BESS_TL (red lines) at (<b>a</b>) US-Ne1, (<b>b</b>) US-Ne2, (<b>c</b>) US-Ne3.</p>
Full article ">Figure 8
<p>The annual GPP estimation discrepancy of the BESS (blue lines), BESS_P (green lines) and BESS_TL (red lines) at (<b>a</b>) Daman, (<b>b</b>) Fenzidi.</p>
Full article ">Figure 9
<p>The estimation of daily GPP provided by BESS (blue lines), BESS_P (green lines) and BESS_TL (red lines) compared to observation by flux sites (black lines).</p>
Full article ">Figure 9 Cont.
<p>The estimation of daily GPP provided by BESS (blue lines), BESS_P (green lines) and BESS_TL (red lines) compared to observation by flux sites (black lines).</p>
Full article ">Figure 10
<p>The daily GPP estimation discrepancy of BESS (blue lines), BESS_P (green lines) and BESS_TL (red lines). The center lines are the multi-year averages, and the edges of the shadow area are mean ± standard deviation.</p>
Full article ">Figure A1
<p>Machine learning training result for LAI using ExtraTreesRegressor.</p>
Full article ">
19 pages, 7787 KiB  
Article
Agent-Based Spatial Dynamic Modeling of Diatraea saccharalis and the Natural Parasites Cotesia flavipes and Trichogramma galloi in Sugarcane Crops
by Rayanna Barroso de Oliveira Alves, Diego Bogado Tomasiello, Cláudia Maria de Almeida, David Luciano Rosalen, Luiz Henrique Pereira, Hernande Pereira da Silva and Cesar Leandro Rodrigues
Remote Sens. 2024, 16(15), 2693; https://doi.org/10.3390/rs16152693 - 23 Jul 2024
Viewed by 303
Abstract
In recent years, sugarcane production areas in Brazil have experienced a slower evolution in productivity, and one of the reasons for this is related to the increase in phytosanitary problems, such as the presence of pests. Nevertheless, limited attention has been paid to [...] Read more.
In recent years, sugarcane production areas in Brazil have experienced a slower evolution in productivity, and one of the reasons for this is related to the increase in phytosanitary problems, such as the presence of pests. Nevertheless, limited attention has been paid to the development of tools for simulating the spatial dynamics of pests, including their impact on production. This study aims to simulate the potential population growth and dispersal of Diatraea saccharalis (sugarcane borer) in sugarcane crop fields and its estimated impacts on this crop production and to simulate biological control strategies. We developed an agent-based model to simulate the pest population and its dispersal in a one-hectare sugarcane crop field in Pederneiras, São Paulo, Brazil, delimited with the aid of satellite imagery, considering two scenarios: the first without biological control and the second with biological control using the parasites Trichogramma galloi and Cotesia flavipes. The model was developed using the NetLogo 6.3.0 software. The results indicate that the model accurately reproduced the infestation rates reported in the literature. Additionally, it provided insights into expected pest dispersal, potential production losses, and how the use of T. galloi in association with C. flavipes could mitigate production losses. We believe that the model can be used to simulate different biological control strategies and the implementation of integrated pest management (IPM) to achieve adequate control levels and greater productivity in sugarcane production. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>Evolutionary phases of <span class="html-italic">D. saccharalis</span>. (<b>A</b>) Insect eggs; (<b>B</b>) larva; (<b>C</b>) pupa; (<b>D</b>) pest in the adult stage, male on the left side and female on the right side [<a href="#B5-remotesensing-16-02693" class="html-bibr">5</a>].</p>
Full article ">Figure 2
<p>(<b>A</b>) Hole caused by <span class="html-italic">D. saccharalis</span> larva on sugarcane; (<b>B</b>) larva inside a gallery; (<b>C</b>) culm with a larva; (<b>D</b>) damaged stems. Adapted from [<a href="#B9-remotesensing-16-02693" class="html-bibr">9</a>].</p>
Full article ">Figure 3
<p>Study area location map.</p>
Full article ">Figure 4
<p>Flowchart of <span class="html-italic">D. saccharalis</span> dispersal without biological control.</p>
Full article ">Figure 5
<p>Second scenario simulation flowchart.</p>
Full article ">Figure 6
<p>Flowchart of the life cycle of the parasite <span class="html-italic">T. galloi</span>. Note: * Parasitism percentage relative to distance: up to 5 m—51%; between 5 m and 10 m—41%; and between 10 m and 15 m—21%.</p>
Full article ">Figure 7
<p>Flowchart of the <span class="html-italic">C. flavipes</span> life cycle.</p>
Full article ">Figure 8
<p><span class="html-italic">D. saccharalis</span> simulated population distribution without biological control.</p>
Full article ">Figure 9
<p>Simulated sugarcane crop damage caused by <span class="html-italic">D. saccharalis</span> without biological control.</p>
Full article ">Figure 10
<p><span class="html-italic">D. saccharalis</span> simulated population distribution under biological control.</p>
Full article ">Figure 11
<p>The parasitized population of <span class="html-italic">D. saccharalis</span> eggs.</p>
Full article ">Figure 12
<p>The parasitized population of <span class="html-italic">D. saccharalis</span> larvae.</p>
Full article ">Figure 13
<p>Simulated sugarcane crop damage caused by <span class="html-italic">D. saccharalis</span> with biological control.</p>
Full article ">
14 pages, 5642 KiB  
Article
From Marginal Lands to Biofuel Bounty: Predicting the Distribution of Oilseed Crop Idesia polycarpa in Southern China’s Karst Ecosystem
by Yangyang Wu, Panli Yuan, Siliang Li, Chunzi Guo, Fujun Yue, Guangjie Luo, Xiaodong Yang, Zhonghua Zhang, Ying Zhang, Jinli Yang, Haobiao Wu and Guanghong Zhou
Agronomy 2024, 14(7), 1563; https://doi.org/10.3390/agronomy14071563 - 18 Jul 2024
Viewed by 439
Abstract
With the global energy crisis and the decline of fossil fuel resources, biofuels are gaining attention as alternative energy sources. China, as a major developing country, has long depended on coal and is now looking to biofuels to diversify its energy structure and [...] Read more.
With the global energy crisis and the decline of fossil fuel resources, biofuels are gaining attention as alternative energy sources. China, as a major developing country, has long depended on coal and is now looking to biofuels to diversify its energy structure and ensure sustainable development. However, due to its large population and limited arable land, it cannot widely use corn or sugarcane as raw materials for bioenergy. Instead, the Chinese government encourages the planting of non-food crops on marginal lands to safeguard food security and support the biofuel sector. The Southern China Karst Region, with its typical karst landscape and fragile ecological environment, offers a wealth of potential marginal land resources that are suitable for planting non-food energy crops. This area is also one of the most impoverished rural regions in China, confronting a variety of challenges, such as harsh natural conditions, scarcity of land, and ecological deterioration. Idesia polycarpa, as a fast-growing tree species that is drought-tolerant and can thrive in poor soil, is well adapted to the karst region and has important value for ecological restoration and biodiesel production. By integrating 19 bioclimatic variables and karst landform data, our analysis reveals that the Maximum Entropy (MaxEnt) model surpasses the Random Forest (RF) model in predictive accuracy for Idesia polycarpa’s distribution. The karst areas of Sichuan, Chongqing, Hubei, Hunan, and Guizhou provinces are identified as highly suitable for the species, aligning with regions of ecological vulnerability and poverty. This research provides critical insights into the strategic cultivation of Idesia polycarpa, contributing to ecological restoration, local economic development, and the advancement of China’s biofuel industry. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>The study area of the Southern Karst Region of China. YGGKDR (Yunnan–Guizhou–Guangxi Karst Desertification Region), WYBMR (Western Yunnan Border Mountain Region), FPTR (Four Provinces Tibetan Region), LXMR (Luoxiao Mountain Region), WLMR (Wuling Mountain Region), DBMR (Dabie Mountain Region), WMMR (Wumeng Mountain Region), and QBMR (Qinba Mountain Region).</p>
Full article ">Figure 2
<p>Heatmap of the correlation analysis between environmental variables.</p>
Full article ">Figure 3
<p>Accuracy of species distribution model forecasts: (<b>a</b>) Maximum Entropy model; (<b>b</b>) Random Forest model.</p>
Full article ">Figure 4
<p>Spatial distribution and provincial proportion of suitable areas for <span class="html-italic">Idesia polycarpa</span> in the Southern China Karst Region: (<b>a</b>) spatial distribution of suitable areas for <span class="html-italic">Idesia polycarpa</span> in the Southern China Karst Region; (<b>b</b>) proportion of suitable areas for <span class="html-italic">Idesia polycarpa</span> in each province of the Southern China Karst Region.</p>
Full article ">Figure 5
<p>The proportion and area of suitable areas of karst and non-karst in karst areas of Southern China: (<b>a</b>) proportion of suitable areas of karst and non-karst at all levels; (<b>b</b>) karst and non-karst suitable area of each grade.</p>
Full article ">Figure 6
<p>The distribution and proportion of suitable areas of each grade in contiguous poverty-stricken areas: (<b>a</b>) distribution of suitable areas of each grade in contiguous poverty-stricken areas; (<b>b</b>) proportion of suitable areas of each grade in contiguous poverty-stricken areas. YGGKDR (Yunnan–Guizhou–Guangxi Karst Desertification Region), WYBMR (Western Yunnan Border Mountain Region), FPTR (Four Provinces Tibetan Region), LXMR (Luoxiao Mountain Region), WLMR (Wuling Mountain Region), DBMR (Dabie Mountain Region), WMMR (Wumeng Mountain Region), and QBMR (Qinba Mountain Region).</p>
Full article ">
27 pages, 3747 KiB  
Article
Employing an Artificial Neural Network Model to Predict Citrus Yield Based on Climate Factors
by Saad S. Almady, Mahmoud Abdel-Sattar, Saleh M. Al-Sager, Saad A. Al-Hamed and Abdulwahed M. Aboukarima
Agronomy 2024, 14(7), 1548; https://doi.org/10.3390/agronomy14071548 - 16 Jul 2024
Viewed by 721
Abstract
Agricultural sustainability is dependent on the ability to predict crop yield, which is vital for farmers, consumers, and researchers. Most of the works used the amount of rainfall, average monthly temperature, relative humidity, etc. as inputs. In this paper, an attempt was made [...] Read more.
Agricultural sustainability is dependent on the ability to predict crop yield, which is vital for farmers, consumers, and researchers. Most of the works used the amount of rainfall, average monthly temperature, relative humidity, etc. as inputs. In this paper, an attempt was made to predict the yield of the citrus crop (Washington Navel orange, Valencia orange, Murcott mandarin, Fremont mandarin, and Bearss Seedless lime) using weather factors and the accumulated heat units. These variables were used as input parameters in an artificial neural network (ANN) model. The necessary information was gathered during the growing seasons between 2010/2011 and 2021/2022 under Egyptian conditions. Weather factors were daily precipitation, yearly average air temperature, and yearly average of air relative humidity. A base air temperature of 13.0 °C was used to determine the accumulated heat units. The heat use efficiency (HUE) for cultivars was determined. The Bearss Seedless lime had the lowest HUE of 9.5 kg/ha °C day, while the Washington Navel orange had the highest HUE of 20.2 kg/ha °C day. The predictive performance of the ANN model with a structure of 9-20-1 with the backpropagation was evaluated using standard statistical measures. The actual and estimated yields from the ANN model were compared using a testing dataset, resulting in a value of RMSE, MAE, and MAPE of 2.80 t/ha, 2.58 t/ha, and 5.41%, respectively. The performance of the ANN model in the training phase was compared to multiple linear regression (MLR) models using values of R2; for MLR models for all cultivars, R2 ranged between 0.151 and 0.844, while the R2 value for the ANN was 0.87. Moreover, the ANN model gave the best performance criteria for evaluation of citrus yield prediction with a high R2, low root mean squared error, and low mean absolute error compared to the performance criteria of data mining algorithms such as K-nearest neighbor (KNN), KStar, and support vector regression. These encouraging outcomes show how the current ANN model can be used to estimate fruit yields, including citrus fruits and other types of fruit. The novelty of the proposed ANN model lies in the combination of weather parameters and accumulated heat units for accurate citrus yield prediction, specifically tailored for Egyptian regional citrus crops. Furthermore, especially in low- to middle-income countries such as Egypt, the findings of this study can greatly enhance the reliance on statistics when making decisions regarding agriculture and climate change. The citrus industry can benefit greatly from these discoveries, which can help with optimization, harvest planning, and postharvest logistics. We recommended furthering proving the robustness and generalization ability of the results in this study by adding more data points. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>The average numbers of days required from flowering to harvesting for the investigated citrus cultivars during the period of seasons 2010/2011 to 2021/2022.</p>
Full article ">Figure 2
<p>Data collection steps and methodology for creating an ANN model to estimate citrus yield.</p>
Full article ">Figure 3
<p>Citrus yield prediction using a multi-layer ANN with a 9-20-1 architecture.</p>
Full article ">Figure 4
<p>Values of yearly precipitation (mm/year) in the experimental area.</p>
Full article ">Figure 5
<p>Average accumulated heat units and fruit yields of the investigated citrus cultivars from season 2010/2011 to season 2020/2021.</p>
Full article ">Figure 6
<p>The fruit yields of all the citrus cultivars in the different seasons for minimum, maximum, and average ± SD (SD means standard deviation).</p>
Full article ">Figure 7
<p>Variation in maximum, minimum, and average ± SD of HUE according to citrus cultivars (SD means standard deviation).</p>
Full article ">Figure 8
<p>Relationship between average numbers of days required for flowering to harvesting and heat use efficiency (HUE) of citrus cultivars.</p>
Full article ">Figure 9
<p>Actual vs. predicted yields using the developed ANN model using the training dataset for citrus prediction.</p>
Full article ">Figure 10
<p>Actual vs. predicted yields using the developed ANN model using the testing dataset for citrus prediction.</p>
Full article ">Figure 11
<p>The contribution percentages of the input variables to citrus yield using an ANN structure of 9-20-1.</p>
Full article ">
19 pages, 7807 KiB  
Article
Calibration and Verification of Discrete Element Parameters of Surface Soil in Camellia Oleifera Forest
by Xueting Ma, Yong You, Deqiu Yang, Decheng Wang, Yunting Hui, Daoyi Li and Haihua Wu
Agronomy 2024, 14(5), 1011; https://doi.org/10.3390/agronomy14051011 - 10 May 2024
Viewed by 616
Abstract
To analyze the interaction between the surface soil and the soil-contacting component (65 Mn) in the camellia oleifera forest planting area in Changsha City, Hunan, China, in this study, we conducted discrete element calibration using physical and simulation tests. The chosen contact model [...] Read more.
To analyze the interaction between the surface soil and the soil-contacting component (65 Mn) in the camellia oleifera forest planting area in Changsha City, Hunan, China, in this study, we conducted discrete element calibration using physical and simulation tests. The chosen contact model was Hertz–Mindlin with JKR cohesion, with the soil repose angle as the response variable. The repose angle of the soil was determined to be 36.03° based on the physical tests. The significant influencing factors of the repose angle determined based on the Plackett–Burman test were the soil–soil recovery coefficient, soil–soil rolling friction coefficient, soil-65 Mn static friction coefficient, and surface energy of soil for the JKR model. A regression model for the repose angle was developed using the Box–Behnken response surface optimization method to identify the best parameter combination. The optimal parameter combination for the JKR model was determined as follows: surface energy of soil: 0.400, soil–soil rolling friction coefficient: 0.040, soil-65 Mn static friction coefficient: 0.404, and soil–soil recovery coefficient: 0.522. The calibrated discrete element parameters were validated through experiments on the repose angle and steel rod insertion. The results indicated that the relative errors obtained from the two verification methods were 2.44% and 1.71%, respectively. This research offers fundamental insights for understanding the interaction between soil and soil-contacting components and optimizing their design. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>Process of parameter calibration.</p>
Full article ">Figure 2
<p>Soil screening.</p>
Full article ">Figure 3
<p>Size classification and mass fraction of soil.</p>
Full article ">Figure 4
<p>Principle of determination of the soil static friction coefficient. <span class="html-italic">F</span><sub>1</sub> is the traction force, N; <span class="html-italic">F</span><sub>2</sub> is the pressure, N; <span class="html-italic">G</span> is the gravity, N; <span class="html-italic">f</span> is the friction force, N; <span class="html-italic">N</span> is the support force, N; <span class="html-italic">θ</span> is the angle between the base and slope, °; <span class="html-italic">b</span> is the base length, mm; <span class="html-italic">h</span> is the height of the slope, mm.</p>
Full article ">Figure 5
<p>Measurement of the soil-65 Mn steel static friction angle. (<b>a</b>) Static friction measurement device. (<b>b</b>) Soil-65 Mn static friction angle measurement results.</p>
Full article ">Figure 6
<p>Soil repose angle determination test.</p>
Full article ">Figure 7
<p>Process of obtaining the soil repose angle. (<b>a</b>) Soil accumulation; (<b>b</b>) image cropping; (<b>c</b>) contrast enhancement and target extraction; (<b>d</b>) binarization; (<b>e</b>) contour shape extraction; (<b>f</b>) linear fitting of the contour shape and calculation of the repose angle.</p>
Full article ">Figure 8
<p>Fitting results of the soil accumulation test.</p>
Full article ">Figure 9
<p>Soil particle model and device model. (<b>a</b>) Soil particle; (<b>b</b>) device model.</p>
Full article ">Figure 10
<p>Schematic diagram of the contact model. Note: <span class="html-italic">O<sub>i</sub></span> and <span class="html-italic">O<sub>j</sub></span> are the spherical center positions of the two particles, respectively; <span class="html-italic">R<sub>i</sub></span> and <span class="html-italic">R<sub>j</sub></span> are the radii of the two particles, respectively; <span class="html-italic">δ<sub>n</sub></span> is the normal overlap amount when the particles collide (m); <span class="html-italic">k<sub>n</sub></span> and <span class="html-italic">k<sub>τ</sub></span> are the normal and tangential Huco coefficients of the particles, respectively; <span class="html-italic">β<sub>n</sub></span> and <span class="html-italic">β<sub>τ</sub></span> are the normal and tangential damping coefficients of the particles, respectively; and <span class="html-italic">μ</span> is the static friction coefficient between the particles.</p>
Full article ">Figure 11
<p>Repose angle in the simulation test.</p>
Full article ">Figure 12
<p>Ranking of the influencing factors.</p>
Full article ">Figure 13
<p>The interactive effects of various influencing factors on the repose angle. (<b>a</b>) CG response surface; (<b>b</b>) EG response surface; (<b>c</b>) AG response surface; (<b>d</b>) EC response surface; (<b>e</b>) AC response surface; (<b>f</b>) AE response surface.</p>
Full article ">Figure 13 Cont.
<p>The interactive effects of various influencing factors on the repose angle. (<b>a</b>) CG response surface; (<b>b</b>) EG response surface; (<b>c</b>) AG response surface; (<b>d</b>) EC response surface; (<b>e</b>) AC response surface; (<b>f</b>) AE response surface.</p>
Full article ">Figure 14
<p>Comparison between the simulation test and physical test. (<b>a</b>) Simulation test; (<b>b</b>) physical test.</p>
Full article ">Figure 15
<p>Physical test and simulation test. (<b>a</b>) Physical test. (<b>b</b>) Simulation test.</p>
Full article ">Figure 16
<p>Simulation test and physical test results of steel rod insertion.</p>
Full article ">
18 pages, 2252 KiB  
Article
Simulation Model for Assessing High-Temperature Stress on Rice
by Haoyang Zhou, Xianguan Chen, Minglu Li, Chunlin Shi and Min Jiang
Agronomy 2024, 14(5), 900; https://doi.org/10.3390/agronomy14050900 - 25 Apr 2024
Viewed by 699
Abstract
Rice is a staple grain crop extensively cultivated in Fujian Province, China. This study examined the impact of high-temperature stress on rice yield and its components, focusing on four representative rice varieties, including early and middle rice grown in Fujian Province. Results indicate [...] Read more.
Rice is a staple grain crop extensively cultivated in Fujian Province, China. This study examined the impact of high-temperature stress on rice yield and its components, focusing on four representative rice varieties, including early and middle rice grown in Fujian Province. Results indicate significant yield losses, with the most severe reduction of 60.8% observed during the flowering stage of early rice and over 40% during the meiosis and flowering stages of middle rice. High-temperature stress primarily affects early rice yield more at the flowering stage than at the grain-filling stage, whereas in middle rice, it is more severe at the meiosis stage than at the flowering stage. Leveraging historical climatic data spanning the past 20 years, a simulation model for high-temperature stress on rice yield was developed to assess disaster-induced yield loss rates, aiming to enhance prevention and disaster damage assessment for rice under high-temperature stress. Application of the model to four rice planting sites in Fujian Province revealed contrasting temporal changes between loss rates and meteorological yield, with middle rice experiencing more severe damage than early rice. The model’s effectiveness is validated by the strong correspondence between yield loss rate and meteorological yield across different regions, highlighting its robust simulation capabilities. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>Fluctuations in daily maximum and average temperatures in the natural environment during the high-temperature treatment period.</p>
Full article ">Figure 2
<p>Changes in yield components of early rice and middle rice at the different stages with ADHT.</p>
Full article ">Figure 3
<p>Comparison between simulated and observed values of rice growth period.</p>
Full article ">Figure 4
<p>Trends of rice yield at four representative sites from 2001 to 2020.</p>
Full article ">Figure 5
<p>High-temperature yield loss rate and meteorological yield in four representative sites from 2001 to 2020.</p>
Full article ">
15 pages, 2645 KiB  
Article
Synchronous Retrieval of Wheat Cab and LAI from UAV Remote Sensing: Application of the Optimized Estimation Inversion Framework
by Jiangtao Ji, Xiaofei Wang, Hao Ma, Fengxun Zheng, Yi Shi, Hongwei Cui and Shaoshuai Zhao
Agronomy 2024, 14(2), 359; https://doi.org/10.3390/agronomy14020359 - 10 Feb 2024
Cited by 1 | Viewed by 854
Abstract
Chlorophyll a and b content (Cab) and leaf area index (LAI) are two key parameters of crops, and their quantitative inversions are important for growth monitoring and the field management of wheat. However, due to the close correlation between the spectral signals of [...] Read more.
Chlorophyll a and b content (Cab) and leaf area index (LAI) are two key parameters of crops, and their quantitative inversions are important for growth monitoring and the field management of wheat. However, due to the close correlation between the spectral signals of these two parameters and the effects of soil and atmospheric conditions, as well as modeling errors, synchronous retrieval of LAI and Cab from remote sensing data is still a challenging task. In a previous study, we introduced the optimal estimation theory and established the inversion framework by coupling the PROSAIL (PROSPECT + SAIL) model with the unified linearized vector radiative transfer model (UNL-VRTM). The framework fully utilizes the simulated radiance spectra for synchronous retrieval of Cab and LAI at the UAV observation scale and has good convergence and self-consistency. In this study, based on this inversion framework, synchronized retrieval of Cab and LAI was carried out by real wheat UAV observation data and validated with the ground-measured data. By comparing with the empirical statistical model constructed by the PROSAIL model and coupled model, least squares support vector machine (LSSVM), and random forest (RF), the proposed method has the highest accuracy of Cab and LAI estimated from UAV multispectral data (for Cab, R2 = 0.835, RMSE = 14.357; for LAI, R2 = 0.892, RMSE = 0.564). Our proposed method enables the fast and efficient estimation of Cab and LAI in multispectral data without prior measurements and training. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>The study area (red points in the right diagram are the ground data validation collection points).</p>
Full article ">Figure 2
<p>Distribution of Cab and LAI measure data in different periods.</p>
Full article ">Figure 3
<p>Comparison of simulation data with UAV observation data. (<b>a</b>) The comparison of a set of simulated data with the UAV observations; (<b>b</b>) The comparison of a total of 63 sets of simulated data with the UAV observations.</p>
Full article ">Figure 4
<p>Results of Cab estimation based on four regression models: CIgreen (<b>a</b>), CIre (<b>b</b>), MTCI (<b>c</b>) and NDRE (<b>d</b>).</p>
Full article ">Figure 5
<p>Results of LAI estimation based on four regression models: NDVI (<b>a</b>), GNDVI (<b>b</b>), BNDVI (<b>c</b>) and RVI (<b>d</b>).</p>
Full article ">Figure 6
<p>Results of Cab and LAI estimation based on the optimal estimation method and the simulated statistical regression method.</p>
Full article ">
18 pages, 4156 KiB  
Article
Deep Learning Model Effectiveness in Forecasting Limited-Size Aboveground Vegetation Biomass Time Series: Kenyan Grasslands Case Study
by Efrain Noa-Yarasca, Javier M. Osorio Leyton and Jay P. Angerer
Agronomy 2024, 14(2), 349; https://doi.org/10.3390/agronomy14020349 - 8 Feb 2024
Cited by 2 | Viewed by 1312
Abstract
Timely forecasting of aboveground vegetation biomass is crucial for effective management and ensuring food security. However, research on predicting aboveground biomass remains scarce. Artificial intelligence (AI) methods could bridge this research gap and provide early warning to planners and stakeholders. This study evaluates [...] Read more.
Timely forecasting of aboveground vegetation biomass is crucial for effective management and ensuring food security. However, research on predicting aboveground biomass remains scarce. Artificial intelligence (AI) methods could bridge this research gap and provide early warning to planners and stakeholders. This study evaluates the effectiveness of deep learning (DL) algorithms in predicting aboveground vegetation biomass with limited-size data. It employs an iterative forecasting procedure for four target horizons, comparing the performance of DL models—multi-layer perceptron (MLP), long short-term memory (LSTM), gated recurrent unit (GRU), convolutional neural network (CNN), and CNN-LSTM—against the traditional seasonal autoregressive integrated moving average (SARIMA) model, serving as a benchmark. Five limited-size vegetation biomass time series from Kenyan grasslands with values at 15-day intervals over a 20-year period were chosen for this purpose. Comparing the outcomes of these models revealed significant differences (p < 0.05); however, none of the models proved superior among the five time series and the four horizons evaluated. The SARIMA, CNN, and CNN-LSTM models performed best, with the statistical model slightly outperforming the other two. Additionally, the accuracy of all five models varied significantly according to the prediction horizon (p < 0.05). As expected, the accuracy of the models decreased as the prediction horizon increased, although this relationship was not strictly monotonic. Finally, this study indicated that, in limited-size aboveground vegetation biomass time series, there is no guarantee that deep learning methods will outperform traditional statistical methods. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>Aboveground vegetation biomass time series generated from calibrated PHYGROW model simulations across five representative rangeland sites in Kenya. (<b>a</b>) Time series 1 (TS1), (<b>b</b>) Time series 2 (TS2), (<b>c</b>) Time series 3 (TS3), (<b>d</b>) Time series 4 (TS4), and (<b>e</b>) Time series 5 (TS5).</p>
Full article ">Figure 2
<p>Using a sliding window to configure time series as a supervised data set.</p>
Full article ">Figure 3
<p>Long short-term memory network architecture.</p>
Full article ">Figure 4
<p>Gated Recurrent Unit network architecture.</p>
Full article ">Figure 5
<p>Model architecture overview: CNN with MLP hidden layer, CNN-LSTM with LSTM hidden layer.</p>
Full article ">Figure 6
<p>Performance of DL and statistical models in aboveground vegetation biomass time series forecasting over four horizons. (<b>a</b>–<b>e</b>) correspond to time series TS1 to TS5, respectively.</p>
Full article ">
18 pages, 8310 KiB  
Article
Estimating the Light Interception and Photosynthesis of Greenhouse-Cultivated Tomato Crops under Different Canopy Configurations
by Yue Zhang, Michael Henke, Yiming Li, Zhouping Sun, Weijia Li, Xingan Liu and Tianlai Li
Agronomy 2024, 14(2), 249; https://doi.org/10.3390/agronomy14020249 - 24 Jan 2024
Cited by 3 | Viewed by 1176
Abstract
Understanding the spatial heterogeneity of light and photosynthesis distribution within a canopy is crucial for optimizing plant growth and yield, especially in the context of greenhouse structures. In previous studies, we developed a 3D functional-structural plant model (FSPM) of the Chinese solar greenhouse [...] Read more.
Understanding the spatial heterogeneity of light and photosynthesis distribution within a canopy is crucial for optimizing plant growth and yield, especially in the context of greenhouse structures. In previous studies, we developed a 3D functional-structural plant model (FSPM) of the Chinese solar greenhouse (CSG) and tomato plants, in which the greenhouse was reconstructed as a 3D mockup and implemented in the virtual scene. This model, which accounts for various environmental factors, allows for precise calculations of radiation, temperature, and photosynthesis at the organ level. This study focuses on elucidating optimal canopy configurations for mechanized planting in greenhouses, building upon the commonly used north–south (N–S) orientation by exploring the east–west (E–W) orientation. Investigating sixteen scenarios with varying furrow distance (1 m, 1.2 m, 1.4 m, 1.6 m) and row spacing (0.3 m, 0.4 m, 0.5 m, 0.6 m), corresponding to 16 treatments of plant spacing, four planting patterns (homogeneous row, double row, staggered row, incremental row) and two orientations were investigated. The results show that in Shenyang city, an E–W orientation with the path width = 0.5 (furrow distance + row distance) = 0.8 m (homogeneous row), and a plant distance of 0.32 m, is the optimal solution for mechanized planting at a density of 39,000 plants/ha. Our findings reveal a nuanced understanding of how altering planting configurations impacts the light environment and photosynthesis rate within solar greenhouses. Looking forward, these insights not only contribute to the field of CSG mechanized planting, but also provide a basis for enhanced CSG planting management. Future research could further explore the broader implications of these optimized configurations in diverse geographic and climatic conditions. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>Photograph of a real CSG (Liaoshen type) (<b>a</b>). Three-dimensional model of a 15 m long Chinese Liaoshen solar greenhouse including the tomato canopy with E–W row orientation (<b>b</b>). The 15 m long version of the greenhouse was used to simulate all scenarios, which can significantly reduce the calculation time and has no impact on the overall simulated results (the CSG has 60 m length).</p>
Full article ">Figure 2
<p>Schematic representation of the E–W row orientation with four trellis patterns, namely, homogeneous, double row, staggered, and incremental row. E.g., F: furrow distance, R: row distance, S: plant spacing, P: (furrow distance + row distance)/2.</p>
Full article ">Figure 3
<p>The daily average plant incident solar radiation under four planting patterns of E–W orientation as the furrow distance (<b>a</b>), row distance (<b>b</b>), and plant distance (<b>c</b>) increases.</p>
Full article ">Figure 4
<p>The midday plant average leaflet temperature under four planting patterns of E–W orientation as the furrow distance (<b>a</b>), row distance (<b>b</b>), and plant distance (<b>c</b>) increases.</p>
Full article ">Figure 5
<p>The midday plant average leaflet photosynthesis under four planting patterns of E–W orientation as the furrow distance (<b>a</b>), row distance (<b>b</b>), and plant distance (<b>c</b>) increases. 1.5 IQR represents 1.5 times interquartile range; any data point exceeding 1.5IQR should be considered as an outlier.</p>
Full article ">Figure 6
<p>The plant daily incident solar radiation under the four planting patterns of N–S orientation (<b>a</b>) and E–W orientation (<b>b</b>) at F = 1.2 m, R = 0.4 m, S = 0.32 m. Each color circle and height represent the incident radiation intensity of a whole tomato plant that is the sum of all leaves.</p>
Full article ">Figure 7
<p>The daily averaged incident solar radiation per leaf rank under the four planting patterns in N–S orientation (<b>a</b>) and E–W orientation (<b>b</b>).</p>
Full article ">Figure 8
<p>The daily averaged incident solar radiation per leaf rank of the planting pattern 1 (P = 0.8 m, S = 0.32 m) of N–S and E–W orientations.</p>
Full article ">Figure 9
<p>Histogram analysis of the number of plants corresponding with plant daily total incident solar radiation of N–S and E–W orientation of planting pattern 1 (homogeneous row, P = 0.8 m, S = 0.32 m) (<b>a</b>). The plant daily accumulated solar radiation of N–S and E–W orientations under planting pattern 1 (<b>b</b>). The plant average solar radiation intensity of N–S and E–W orientations under planting pattern 1 (<b>c</b>).</p>
Full article ">Figure 10
<p>Heat map of leaflet photosynthesis of tomato canopy at upper (leaf rank = 18), middle (leaf rank = 11), and lower layers (leaf rank = 4) at 9:00 a.m. (<b>a</b>), 12:00 p.m. (<b>b</b>), and 4:00 p.m. (<b>c</b>) of N–S (left side) and E–W orientations (right side) under scenario 6 (P = 0.8 m, PS = 0.32 m).</p>
Full article ">Figure 11
<p>The daily averaged photosynthesis rate per leaf rank of planting pattern 1 (homogeneous row, P = 0.8 m, PS = 0.32 m) under N–S orientation and E–W orientation (<b>a</b>). Average plant photosynthesis of middle lane tomato crop with increasing distance from south corner (1–6.5 m) simulated by day (<b>b</b>).</p>
Full article ">Figure 12
<p>Partial least squares path modeling (PLS-PM) of all collected physiological and morphological data. The finalized version of the PLS-PM using SMARTPLS3.0 is shown. Simulated variables are represented in a rectangle form, while traits within the big circles and polygons are latent variables. Indicated are the loadings (the correlations between a latent variable and its simulated variables) and the path coefficients calculated after 1000 bootstraps.</p>
Full article ">
22 pages, 1766 KiB  
Article
ARMOSA Model Parametrization for Winter Durum Wheat Cultivation under Diverse Cropping Management Practices in a Mediterranean Environment
by Pasquale Garofalo, Marco Parlavecchia, Luisa Giglio, Ivana Campobasso, Alessandro Vittorio Vonella, Marco Botta, Tommaso Tadiello, Vincenzo Tucci, Francesco Fornaro, Rita Leogrande, Carolina Vitti, Alessia Perego, Marco Acutis and Domenico Ventrella
Agronomy 2024, 14(1), 164; https://doi.org/10.3390/agronomy14010164 - 11 Jan 2024
Viewed by 991
Abstract
In anticipation of climate changes, strategic soil management, encompassing reduced tillage and optimized crop residue utilization, emerges as a pivotal strategy for climate impact mitigation. Evaluating the transition from conventional to conservative cropping systems, especially the equilibrium shift in the medium to long [...] Read more.
In anticipation of climate changes, strategic soil management, encompassing reduced tillage and optimized crop residue utilization, emerges as a pivotal strategy for climate impact mitigation. Evaluating the transition from conventional to conservative cropping systems, especially the equilibrium shift in the medium to long term, is essential. ARMOSA, a robust crop simulation model, adeptly responds to varied soil management practices such as no tillage, minimum tillage, and specific straw management options such as chopping and incorporating crop residue into the soil (with or without prior nitrogen and water addition before ploughing). It effectively captures dynamic fluctuations in total organic carbon over an extended period. While challenges persist in precisely predicting grain yield due to climatic intricacies, ARMOSA stands out as a valuable and versatile tool. The model excels in comprehending and simulating wheat cultivar responses in dynamic agricultural ecosystems, shedding light on phenological patterns, biomass accumulation, and soil organic carbon dynamics. This research significantly advances our understanding of the intricate complexities associated with past wheat cultivation in diverse environmental conditions. ARMOSA’s ability to inform decisions on conservation practices positions it as a valuable asset for researchers, agronomists, and policymakers navigating the challenges of sustainable agriculture amidst climate change. Its real-world significance lies in its potential to guide informed decisions, contributing to global efforts in sustainable agriculture and climate resilience. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>Trend of the total dry biomass (<b>a</b>) and grain yield (<b>b</b>) at the harvest of durum wheat following one another in the growing years for P_30. <span class="html-italic">Va</span> stands for Valgerardo, <span class="html-italic">Ap</span> for Appulo, <span class="html-italic">La</span> for Latino, <span class="html-italic">Ai</span> for Appio, <span class="html-italic">Si</span> for Simeto, <span class="html-italic">Cl</span> from Claudio, and <span class="html-italic">Sa</span> for Saragolla.</p>
Full article ">Figure 2
<p>Comparison between simulated and observed data on TOC (0–40 cm) for P_30. Bars indicate the standard deviation.</p>
Full article ">Figure 3
<p>Linear regression (thin line) between observed grain yield (Obs_yield) and simulated grain yield (Sim_yield) of P_30. Empty circles indicate the yield averaged for each cultivar. Thin black line indicates 1:1 fit.</p>
Full article ">Figure 4
<p>Linear regression between observed TOC (Obs_TOC) and simulated yield (Sim_TOC) achieved by <span class="html-italic">NT</span> (grey line), <span class="html-italic">MT</span> (thin black line), P_32 (dashed line) in the validation step (<b>a</b>); thin black line indicates 1:1 fit. TOC (0–40 cm) dynamics of observed (obs) and simulated (sim) <span class="html-italic">NT</span> and <span class="html-italic">MT</span> verified across experimental years of LTE (<b>b</b>).</p>
Full article ">
15 pages, 4203 KiB  
Article
Estimation of Spring Maize Planting Dates in China Using the Environmental Similarity Method
by Meiling Sheng, A-Xing Zhu, Tianwu Ma, Xufeng Fei, Zhouqiao Ren and Xunfei Deng
Agronomy 2024, 14(1), 97; https://doi.org/10.3390/agronomy14010097 - 30 Dec 2023
Viewed by 946
Abstract
Global climate change is a serious threat to food and energy security. Crop growth modelling is an important tool for simulating crop food production and assisting in decision making. Planting date is one of the important model parameters. Larger-scale spatial distribution with high [...] Read more.
Global climate change is a serious threat to food and energy security. Crop growth modelling is an important tool for simulating crop food production and assisting in decision making. Planting date is one of the important model parameters. Larger-scale spatial distribution with high accuracy for planting dates is essential for the widespread application of crop growth models. In this study, a planting date prediction method based on environmental similarity was developed in accordance with the third law of geography. Spring maize planting date observations from 124 agricultural meteorological experiment stations in China over the years 1992–2010 were used as the data source. Samples spanning from 1992 to 2009 were allocated as training data, while samples from 2010 constituted the independent validation set. The results indicated that the root mean square error (RMSE) for spring maize planting date based on environmental similarity was 10 days, which is better than that of multiple regression analysis (RMSE = 13 days) in 2010. Additionally, when applied at varying scales, the accuracy of national-scale prediction was better than that of regional-scale prediction in areas with large differences in planting dates. Consequently, the method based on environmental similarity can effectively and accurately estimate planting date parameters at multiple scales and provide reasonable parameter support for large-scale crop growth modelling. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>Spring maize cultivation zones of China. The sample points are the locations of agricultural meteorological stations in the cultivation zones.</p>
Full article ">Figure 2
<p>Flowchart of spatially predicting PD based on environmental similarity.</p>
Full article ">Figure 3
<p>Scatterplot of actual versus predicted planting dates using independent validation sample points for spring maize of China: (<b>a</b>) environmental similarity method; (<b>b</b>) multiple line regression method. The dashed line in the plot represents the 1:1 line.</p>
Full article ">Figure 4
<p>The scatter plots of the residual-predicted values of the validation sample points obtained from the two prediction methods: (<b>a</b>) environmental similarity method; (<b>b</b>) multiple line regression method.</p>
Full article ">Figure 5
<p>Predicted spatial distribution of planting dates (day of year) for spring maize in 2010: (<b>a</b>) using environmental similarity method; (<b>b</b>) multiple linear regression method.</p>
Full article ">Figure 6
<p>The spatial distribution of predicted planting dates uncertainty using the environmental similarity method.</p>
Full article ">Figure 7
<p>The distribution of the absolute value of the residuals of the predicted spring maize sowing dates for the 2010 validation sample points: (<b>a</b>) regional-scale prediction; (<b>b</b>) national-scale prediction.</p>
Full article ">
21 pages, 1949 KiB  
Article
Optimal Planting Density and Nutrient Application of Soybeans: A Case Study in Northeastern China
by Huicheng Hao, Shixin Lv and Fulin Wang
Agronomy 2023, 13(12), 2902; https://doi.org/10.3390/agronomy13122902 - 26 Nov 2023
Viewed by 1220
Abstract
In the context of the Chinese government’s policy guidance, there is black soil protection and ecological environment protection. The purpose of this paper is to solve the problem that the soil ecology of the black soil in Northeast China is changing year by [...] Read more.
In the context of the Chinese government’s policy guidance, there is black soil protection and ecological environment protection. The purpose of this paper is to solve the problem that the soil ecology of the black soil in Northeast China is changing year by year, and it is necessary to explore the sowing and fertilization strategy under the new situation; most Chinese growers rely excessively on their personal experience in the process of soybean sowing and fertilization. In this study, we used “Heihe 43” soybeans and used regression experimental design methods to analyze the effects of planting density, nitrogen, phosphorus, and potassium fertilizer application on soybean yield and to determine the optimal planting density and fertilizer ratios. The study reveals that the optimal soybean planting density in Northeast China is 45.37 × 104 plants/ha, with nitrogen at 98.4 kg/ha, phosphorus at 218.96 kg/ha, and potash at 47.62 kg/ha. Under these conditions, soybean yields can reach 3816.67 kg/ha. This study can provide a theoretical method for decision-making to obtain the optimal planting density and fertilizer ratio for different regions of the farming system. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>Experimental Field Plot Dimensions Chart.</p>
Full article ">Figure 2
<p>Comparison of two sampling methods. (<b>a</b>) non-block sampling, (<b>b</b>) grid-based systematic sampling.</p>
Full article ">Figure 3
<p>Annual temperature and total precipitation data for Nenjiang County in 2020.</p>
Full article ">Figure 4
<p>Illustrative graph depicting the relationship between planting density, fertilizer application, and soybean yield.</p>
Full article ">
20 pages, 7141 KiB  
Article
Comprehensive Growth Index (CGI): A Comprehensive Indicator from UAV-Observed Data for Winter Wheat Growth Status Monitoring
by Yuanyuan Tang, Yuzhuang Zhou, Minghan Cheng and Chengming Sun
Agronomy 2023, 13(12), 2883; https://doi.org/10.3390/agronomy13122883 - 24 Nov 2023
Cited by 1 | Viewed by 1149
Abstract
Crop growth monitoring plays an important role in estimating the scale of food production and providing a decision-making basis for agricultural policies. Moreover, it can allow understanding of the growth status of crops, seedling conditions, and changes in a timely manner, overcoming the [...] Read more.
Crop growth monitoring plays an important role in estimating the scale of food production and providing a decision-making basis for agricultural policies. Moreover, it can allow understanding of the growth status of crops, seedling conditions, and changes in a timely manner, overcoming the disadvantages of traditional monitoring methods such as low efficiency and inaccuracy. In order to realize rapid and non-destructive monitoring of winter wheat growth status, this study introduced an equal weight method and coefficient of variation method to construct new comprehensive growth indicators based on drone images and measured data obtained from field experiments. The accuracy of the indicators in evaluating the growth of winter wheat can be judged by the construction, and the effects of different machine learning methods on the construction of indicators can be compared. Correlation analysis and variable screening were carried out on the constructed comprehensive growth indicators and the characteristic parameters extracted by the drone, and the comprehensive growth index estimation model was constructed using the selected parameter combination. Among them, when estimating the comprehensive growth index (CGIavg), the optimal model at the jointing stage is the support vector regression (SVR) model: R2 is 0.77, RMSE is 0.095; at the booting stage, the optimal model is the Gaussian process regression (GPR) model: R2 is 0.71, RMSE is 0.098; at the flowering stage, the optimal model is the SVR model: R2 is 0.78, RMSE is 0.087. When estimating the comprehensive growth index based on the coefficient of variation method (CGIcv), the optimal model at the jointing stage is the multi-scale retinex (MSR) model: R2 is 0.73, RMSE is 0.084; at the booting stage, the optimal model is the GPR model: R2 is 0.74, RMSE is 0.092; at the flowering stage, the optimal model is the SVR model, R2 is 0.78: RMSE is 0.085. The conclusion shows that the method of constructing the comprehensive growth index is superior to the function of a single parameter to some extent, providing a new way for wheat growth monitoring and process management. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>Orthophoto image of the test field taken with a drone.</p>
Full article ">Figure 2
<p>The flowchart of methods.</p>
Full article ">Figure 3
<p>Correlation analysis between the characteristic parameters at the jointing stage and CGIavg. Note: the solid orange line indicates the significance level at <span class="html-italic">p</span> &gt; 0.01; the red dotted line indicates the significance level at <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 4
<p>Correlation analysis between characteristic parameters at the booting stage and CGIavg. Note: the solid orange line indicates the significance level at <span class="html-italic">p</span> &gt; 0.01; the red dotted line indicates the significance level at <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 5
<p>Correlation analysis between characteristic parameters at the flowering stage and CGIavg. Note: the solid orange line indicates the significance level at <span class="html-italic">p</span> &gt; 0.01; the red dotted line indicates the significance level at <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 6
<p>The process of filtering the characteristic parameters of each period using the SPA algorithm. Note: The black curves represent the variables selected at elongation stage, the red curve represents the variables selected booting stage, the blue curves represent the variables selected for flowering stage.</p>
Full article ">Figure 7
<p>Correlation analysis between characteristic parameters at the jointing stage and comprehensive index CGIcv.</p>
Full article ">Figure 8
<p>Correlation analysis between characteristic parameters t the booting stage and comprehensive index CGIcv.</p>
Full article ">Figure 9
<p>Correlation analysis between characteristic parameters at the flowering period and comprehensive index CGIcv.</p>
Full article ">Figure 10
<p>The process of filtering the characteristic parameters of each period using the SPA algorithm. Note: The black curves represent the variables selected at elongation stage, the red curve represents the variables selected booting stage, the blue curves represent the variables selected for flowering stage.</p>
Full article ">Figure 11
<p>Scatterplot with a 1:1 reference line of CGIavg measured and predicted values for the SVR model validation set at the jointing stage.</p>
Full article ">Figure 12
<p>Scatterplot with a 1:1 reference line of CGIavg measured and predicted values in the booting stage GPR model validation set.</p>
Full article ">Figure 13
<p>Scatterplot with a 1:1 reference line of CGIavg measured and predicted values for the SVR model validation set at the flowering stage.</p>
Full article ">Figure 14
<p>Scatterplot with a 1:1 reference line of the measured and predicted values of the MSR model validation set CGIcv at the jointing stage.</p>
Full article ">Figure 15
<p>Scatterplot with a 1:1 reference line graph of the measured and predicted values of the GPR model validation set CGIcv at the booting stage.</p>
Full article ">Figure 16
<p>Scatterplot with a 1:1 reference line of the measured and predicted values of the SVR model validation set CGIcv at the flowering stage.</p>
Full article ">
26 pages, 5989 KiB  
Article
STICS Soil–Crop Model Performance for Predicting Biomass and Nitrogen Status of Spring Barley Cropped for 31 Years in a Gleysolic Soil from Northeastern Quebec (Canada)
by Nomena Ravelojaona, Guillaume Jégo, Noura Ziadi, Alain Mollier, Jean Lafond, Antoine Karam and Christian Morel
Agronomy 2023, 13(10), 2540; https://doi.org/10.3390/agronomy13102540 - 30 Sep 2023
Cited by 4 | Viewed by 2059
Abstract
Spring barley (Hordeum vulgare L.) is an increasingly important cash crop in the province of Quebec (Canada). Soil–crop models are powerful tools for analyzing and supporting sustainable crop production. STICS model has not yet been tested for spring barley grown over several [...] Read more.
Spring barley (Hordeum vulgare L.) is an increasingly important cash crop in the province of Quebec (Canada). Soil–crop models are powerful tools for analyzing and supporting sustainable crop production. STICS model has not yet been tested for spring barley grown over several decades. This study was conducted to calibrate and evaluate the STICS model, without annual reinitialization, for predicting aboveground biomass and N nutrition attributes at harvest during 31 years of successive cropping of spring barley grown in soil (silty clay, Humic Gleysol) from the Saguenay–Lac-Saint-Jean region (northeastern Quebec, Canada). There is a good agreement between observed and predicted variables during the 31 successive barley cropping years. STICS predicted well biomass accumulation and plant N content with a low relative bias (|normalized mean error| = 0–13%) and small prediction error (normalized root mean square error = 6–25%). Overall, the STICS outputs reproduced the same trends as the field-observed data with various tillage systems and N sources. Predictions of crop attributes were more accurate in years with rainfall close to the long-term average. These ‘newly calibrated’ parameters in STICS for spring barley cropped under continental cold and humid climates require validation using independent observation datasets from other sites. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>STICS soil–crop model: inputs, outputs, the different modules/processes, and their respective influences.</p>
Full article ">Figure 2
<p>Predicted aboveground biomass (AGB) of cultivar Scarlett (blue dashed line) and the newly calibrated cultivar Chapais (black solid line) with calibration dataset (7 years and 4 treatments per year making 28 AGB curves).</p>
Full article ">Figure 3
<p>Annual mean of observed (dots) and predicted (grey bars) (<b>a</b>) spring barley aboveground biomass [AGB] and (<b>b</b>) grain yield [GY] from 1990 to 2020 for soil tillage and N fertilization source treatments. Bars on dots are standard deviations (<span class="html-italic">n</span> = 4). LDM: liquid dairy manure; MIN: ammonium nitrate; MP: moldboard plow; CP: chisel plow.</p>
Full article ">Figure 4
<p>Predicted versus observed spring barley aboveground biomass (AGB) and grain yield (GY) for ‘calibration’ (<b>a</b>,<b>c</b>) and ‘evaluation’ (<b>b</b>,<b>d</b>) dataset. Each point is the mean of four replicates. Mean Obs: mean of observed values; Mean Pred: mean of predicted values; n: number of simulation units; MAE: mean absolute error; NME: normalized mean error; NRMSE: normalized root mean square error; EF: model efficiency; R<sup>2</sup>: coefficient of determination; PLP: percentage lack of precision; PLA: percentage lack of accuracy; LDM: liquid dairy manure; MIN: ammonium nitrate; MP: moldboard plow; CP: chisel plow.</p>
Full article ">Figure 5
<p>Annual mean of observed (dots) and predicted (grey bars) (<b>a</b>) N concentration in spring barley aboveground biomass and (<b>b</b>) grain from 1997 to 2020 for soil tillage and N fertilization source treatments. Bars on dots are standard deviations (<span class="html-italic">n</span> = 4). LDM: liquid dairy manure; MIN: ammonium nitrate; MP: moldboard plow; CP: chisel plow.</p>
Full article ">Figure 6
<p>Predicted versus observed N concentration in aboveground biomass (NCAGB) and in grain (NCG) for the ‘calibration’ (<b>a</b>,<b>c</b>) and ‘evaluation’ (<b>b</b>,<b>d</b>) dataset. Each point is the mean of four replicates. Mean Obs: mean of observed values; Mean Pred: mean of predicted values; n: number of simulation units; MAE: mean absolute error; NME: normalized mean error; NRMSE: normalized root mean square error; EF: model efficiency; R<sup>2</sup>: coefficient of determination; PLP: percentage lack of precision; PLA: percentage lack of accuracy; LDM: liquid dairy manure; MIN: ammonium nitrate; MP: moldboard plow; CP: chisel plow.</p>
Full article ">Figure 7
<p>Annual mean of observed (dots) and predicted (grey bars) (<b>a</b>) N uptake by spring barley shoot (NU) and (<b>b</b>) N amount in grain (NAG) from 1997 to 2020 for soil tillage and N fertilization source treatments. Bars on dots are standard deviations (<span class="html-italic">n</span> = 4). LDM: liquid dairy manure; MIN: ammonium nitrate; MP: moldboard plow; CP: chisel plow.</p>
Full article ">Figure 8
<p>Predicted versus observed shoot N uptake (NU) and N amount in grain (NAG) for the ‘calibration’ (<b>a</b>,<b>c</b>) and ‘evaluation’ (<b>b</b>,<b>d</b>) datasets. Each point is the mean of four replicates. Mean Obs: mean of observed values; Mean Pred: mean of predicted values; n: number of simulation units; MAE: mean absolute error; NME: normalized mean error; NRMSE: normalized root mean square error; EF: model efficiency; R<sup>2</sup>: coefficient of determination; PLP: percentage lack of precision; PLA: percentage lack of accuracy; LDM: liquid dairy manure; MIN: ammonium nitrate; MP: moldboard plow; CP: chisel plow.</p>
Full article ">
10 pages, 1049 KiB  
Article
GGE Biplot-Based Transcriptional Analysis of 7 Genes Involved in Steroidal Glycoalkaloid Biosynthesis in Potato (Solanum tuberosum L.)
by Feng Zhao, Yajie Li, Tongxia Cui and Jiangping Bai
Agronomy 2023, 13(8), 2127; https://doi.org/10.3390/agronomy13082127 - 14 Aug 2023
Cited by 1 | Viewed by 1349
Abstract
Steroidal glycoalkaloids (SGAs) are secondary metabolites that are closely associated with the sensory and processing qualities of potato tubers. GGE biplots are a widely used tool for analyzing crop breeding analysis. This study aimed to investigate the effect of light on SGA biosynthesis [...] Read more.
Steroidal glycoalkaloids (SGAs) are secondary metabolites that are closely associated with the sensory and processing qualities of potato tubers. GGE biplots are a widely used tool for analyzing crop breeding analysis. This study aimed to investigate the effect of light on SGA biosynthesis by employing GGE biplots to analyze the transcriptional gene expression of seven genes involved in the SGA biosynthesis pathway. Tubers of five different potato genotypes were incubated for 6, 12, and 24 h under red light. The expression levels of the seven genes were measured using qRT-PCR for analysis. Further analysis of the data was performed using GGE biplots. Our results indicated significantly higher expression levels for Pvs1, Sgt1, and Sgt3 genes than those of the remaining tested genes. Across the three red light illumination durations, Sgt3 showed high and stable expression, although it showed less stability across the different genotypes. Interestingly, the expression patterns of the seven genes were extremely similar for the 12 h and 24 h treatments. It was found that at least 6 h of red light illumination was required for optimal gene expression in all five genotypes, particularly in the genotype Zhuangshu-3 (DXY) after 24 h of treatment. Additionally, significant expression of the seven genes was observed in the L-6 genotype after 12 and 6 h of red light illumination. These results highlight that GGE biplots are an appropriate tool for analyzing and illustrating the differential expression profiles of the seven key genes involved in SGA biosynthesis in potato tubers. This study provides valuable insights into the biosynthesis and metabolism of SGAs in potatoes. Moreover, it demonstrates the potential application of GGE biplots in crop breeding and other research fields. Full article
(This article belongs to the Topic Advances in Crop Simulation Modelling)
Show Figures

Figure 1

Figure 1
<p>The GGE Biplot analysis showing relevance of gene expression and treatments (red light illumination duration and genotypes) and the polygon view of gene expression—treatment interaction for seven genes in five genotypes under three red light illumination durations. (<b>a</b>) Relevance of gene expression and red light illumination duration. (<b>b</b>) Relevance of gene expression and genotypes. (<b>c</b>) The polygon view of the interaction between gene expression and red light illumination durations interaction. (<b>d</b>) The polygon view of the interaction between gene expression and genotypes. The numbers 6, 12, and 24 represent the three illumination durations; D-3, D-6, DXY, JZ-12, and HA represent genotypes involved in the experiment. <span class="html-italic">Pvs1</span>, <span class="html-italic">Sgt1</span>, <span class="html-italic">Sgt2</span>, <span class="html-italic">Sgt3</span>, <span class="html-italic">Pss1</span>, <span class="html-italic">Hmg1</span>, and <span class="html-italic">Hmg2</span> represent seven target enzymes involved in the SGA biosynthetic pathway. The plot is based on gene centering (center = 2) with SD scaling and without transformation of data (scaling = 1, transform = 0), and it is gene metric-preserving (SVP = 2).</p>
Full article ">Figure 2
<p>The GGE Biplot analysis showing the ranking of seven genes for the transcriptional expression across red light induction durations and genotypes. (<b>a</b>) Ranking of average gene expression across light illumination durations. (<b>b</b>) Ranking of average gene expression across genotypes. The numbers 6, 12, and 24 represent the three illumination durations; D-3, D-6, DXY, JZ-12 and HA represent genotypes involved in the experiment. <span class="html-italic">Pvs1</span>, <span class="html-italic">Sgt1</span>, <span class="html-italic">Sgt2</span>, <span class="html-italic">Sgt3</span>, <span class="html-italic">Pss1</span>, <span class="html-italic">Hmg1</span>, and <span class="html-italic">Hmg2</span> represent seven target enzymes involved in the SGA biosynthetic pathway. The plot is based on gene centering (center = 2) with SD scaling and no transformation of the data (scaling = 1, transformation = 0) and is genotype metric-preserving (SVP = 1).</p>
Full article ">Figure 3
<p>The GGE Biplot analysis showing the relationships between red light induction durations and genotypes. (<b>a</b>) Relationship between light illumination durations; (<b>b</b>) relationship among genotypes. The numbers 6, 12, and 24 represent the three illumination durations; D-3, D-6, DXY, JZ-12 and HA represent the genotypes involved in the experiment. <span class="html-italic">Pvs1</span>, <span class="html-italic">Sgt1</span>, <span class="html-italic">Sgt2</span>, <span class="html-italic">Sgt3</span>, <span class="html-italic">Pss1</span>, <span class="html-italic">Hmg1</span>, and <span class="html-italic">Hmg2</span> represent seven target enzymes involved in the SGA biosynthetic pathway. ‘g’ is used to show the distribution of genotypes on the GGE Biplot graph. The plot is based on gene centering (center = 2) with SD scaling and without transformation of data (scaling = 1, transform = 0), and it is gene metric-preserving (SVP = 2).</p>
Full article ">Figure 4
<p>The GGE Biplot analysis showing genotype vs. illumination duration based on the mean value of red light-induced expression of seven genes. (<b>b</b>) The polygon view of the interaction between genotypes and red light illumination duration. (<b>d</b>) Representative comparison of different red light exposure times on gene expression. The numbers 6, 12, and 24 represent the three illumination durations; D-3, D-6, DXY, JZ-12 and L-6 represent genotypes involved in the experiment. <span class="html-italic">Pvs1</span>, <span class="html-italic">Sgt1</span>, <span class="html-italic">Sgt2</span>, <span class="html-italic">Sgt3</span>, <span class="html-italic">Pss1</span>, <span class="html-italic">Hmg1</span>, and <span class="html-italic">Hmg2</span> represent seven target enzymes involved in the SGA biosynthetic pathway. In (<b>a</b>,<b>b</b>,<b>d</b>), the plot is based on gene centering (center = 2) with SD scaling and without transformation of data (scaling = 1, transform = 0), and it is gene metric-preserving (SVP = 2). In <a href="#agronomy-13-02127-f004" class="html-fig">Figure 4</a>c, the plot is based on gene centering (center = 2) with SD scaling and without transformation of data (scaling = 1, transformation = 0), and it is genotype metric-preserving (SVP = 1).</p>
Full article ">
Back to TopTop