Next Article in Journal
Remote-Sensed Determination of Spatiotemporal Properties of Drought and Assessment of Influencing Factors in Ordos, China
Previous Article in Journal
Progress in Research on Deep Learning-Based Crop Yield Prediction
Previous Article in Special Issue
Improving the Spatiotemporal Transferability of Hyperspectral Remote Sensing for Estimating Soil Organic Matter by Minimizing the Coupling Effect of Soil Physical Properties on the Spectrum: A Case Study in Northeast China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Winter Wheat Yield Prediction Based on the ASTGNN Model Coupled with Multi-Source Data

1
College of Resources and Environment, Anhui Agricultural University, Hefei 230031, China
2
Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences, Wuhan 430077, China
3
Anhui Vocational College of Grain Engineering, Hefei 230011, China
*
Author to whom correspondence should be addressed.
Agronomy 2024, 14(10), 2262; https://doi.org/10.3390/agronomy14102262
Submission received: 4 September 2024 / Revised: 27 September 2024 / Accepted: 27 September 2024 / Published: 1 October 2024

Abstract

:
Timely and accurate prediction of winter wheat yields, which is crucial for optimizing production management, maintaining supply–demand balance, and ensuring food security, depends on interactions among numerous factors, such as climate, surface characteristics, and soil quality. Despite the extensive application of deep learning models in this field, few studies have analyzed the effect of the large-scale geospatial characteristics of neighboring regions on crop yields. Therefore, we present an attention-based spatio-temporal Graph Neural Network (ASTGNN) model coupled with geospatial characteristics and multi-source data for improved accuracy of winter wheat yield estimation. The datasets used in this study included multiple types of remote sensing, meteorological, soil, crop yield, and planting area data for Anhui, China, from 2005 to 2020. The results showed that multi-source data led to higher prediction performance than single-source data, and enabled accurate prediction of winter wheat yields three months prior to harvest. Furthermore, the ASTGNN model provided better prediction performance than two traditional crop yield prediction models (R2 = 0.70, RMSE = 0.21 t/ha, MAE = 0.17 t/ha). Therefore, ASTGNN enhances the accuracy of crop yield prediction by incorporating geospatial characteristics. This research has implications for improving agricultural production management, promoting the development of digital agriculture, and addressing climate change in agriculture.

1. Introduction

Agricultural production plays a crucial role in promoting global economic development, reducing poverty, and ensuring food security. The accurate prediction of crop yields not only helps farmers and agricultural managers to optimize production management, improve production efficiency, and reduce production costs but also serves to maintain the market supply–demand balance and stabilize the prices of agricultural products [1,2]. Based on crop yield prediction data, governments can formulate and adjust agricultural policies and take reasonable agricultural security measures to ensure food security [3,4]. Therefore, the timely and accurate prediction of winter wheat yields has great significance.
Crop yields depend on diverse factors, including environmental conditions (e.g., soil and meteorological conditions) [5,6,7,8], crop variety [9,10,11], management practices (e.g., fertilization, irrigation, and crop rotation) [12,13], and plant diseases [14,15,16]. The correlation between these variables and crop yields is usually non-linear, requiring environmental data (e.g., climate and soil data) to reveal the physical conditions associated with crop growth. In recent decades, numerous studies have focused on crop yield prediction. Traditional crop yield estimation methods (e.g., sampling surveys and field observations) are inefficient, unsuitable for large-scale yield prediction, and hindered by insufficient accuracy [17,18,19]. In contrast, the continued development and innovation of intelligent and efficient remote sensing technologies has provided high-resolution, large-scale, and spatially continuous environmental information on crops. Additionally, remote sensing data can present different spectral reflectance information of crops at different key growth stages, providing a useful reference for crop yield estimation. Therefore, remote sensing technologies are among the most effective solutions for crop yield estimation [20,21,22,23,24,25,26]. Previous studies have used remote sensing data and meteorological data alone, or a combination of multi-source data, for crop yield prediction [27,28,29]. However, multi-source data may differ in quality and accuracy because they are typically acquired from different platforms or sensors, which adversely affects the accuracy and stability of the final prediction results [30,31,32,33]. In terms of large-scale crop yield prediction, few studies have determined the relative performance of prediction models based on multi-source data and single-source data.
Statistical models and process-based crop models are two common methods for predicting crop yields [34,35,36,37,38,39,40]. Process-based models simulate the physical processes of crop growth, allowing for the exploration of interactions between yield and environmental factors. However, these models require extensive field data and are computationally intensive [41,42]. In contrast, statistical models establish empirical relationships between yield and predictor variables using existing data, with regression analysis being a widely used method in early applications of remote sensing for yield estimation [43,44,45]. Traditional static regression and mechanistic models are limited in applicability and face significant uncertainties, making it difficult to provide reliable yield predictions [46,47]. While machine learning methods, such as decision trees, random forests, multiple regression, and association rule mining, have been widely applied in crop yield prediction, they typically treat yield as a latent function of input variables such as weather and soil conditions, which may fail to fully capture the complex relationships between these variables [46,48]. Deep learning is increasingly being applied across various fields within agriculture [49,50,51,52]. Deep learning, a hierarchical machine learning method, is particularly well-suited for handling unstructured and unlabeled data. It has demonstrated superior performance in agriculture, especially in feature extraction from large datasets and modeling nonlinear relationships compared to traditional machine learning methods [53]. Given that crop yield prediction depends on a variety of complex factors influencing crop growth, deep learning offers a powerful capability to extract key features from existing data, thereby improving the accuracy of yield predictions [54,55,56,57,58]. For example, Nevavuori et al. [59] used convolutional neural networks (CNNs) based on normalized difference vegetation index (NDVI) and RGB data acquired by unmanned aerial vehicles to predict wheat and barley yields with high accuracy. Moreover, Jiang et al. [60] presented a long short-term memory network (LSTM) method for crop yield prediction, which performed well in learning temporal characteristics compared with machine learning methods such as random forest and least absolute shrinkage and selection operator algorithms. Wang et al. [61] investigated the impact of different combinations of remote sensing, meteorological, and soil data on model performance, finding that incorporating soil data improved the model’s ability to capture yield spatial variability. Tian et al. [62] explored combinations of the Vegetation Temperature Condition Index (VTCI) with remote sensing data and VTCI with both meteorological and remote sensing data to determine the optimal wheat yield estimation model. Zhang et al. [63] demonstrated that combining various input data, including vegetation indices (VIs) and climate indices (CIs), enhanced crop yield prediction, identifying the best combinations of CIs, EIs, and VIs. Sharma et al. [64] applied the CNN-LSTM method for crop yield estimation, comparing models with and without background information (e.g., water bodies, farmland, and urban landscapes), and found that incorporating background information improved yield predictions. Gastli et al. [65] combined Gaussian Processes (GP) with CNN-LSTM to predict crop yields in California, showing better performance with GP-enhanced CNN-LSTM compared to CNN and LSTM alone. Sun et al. [66] used CNN-LSTM for seasonal crop yield predictions and compared it with CNN and LSTM methods, finding that CNN-LSTM performed better over five years, with MODIS surface reflectance significantly affecting model performance.
However, previous studies often treat counties as independent units, neglecting the geographical relationships between neighboring counties, which prevents effective capture of the interactions between regions. In reality, neighboring counties or areas may share similar environmental conditions, and this interconnection significantly impacts crop yield. Traditional deep learning methods, such as CNNs, do not effectively consider and learn the relationships between a node and its neighboring nodes in an image [67]. Graph Neural Networks (GNNs) can capture the interdependencies between a sample and the samples it interacts with, providing a feasible solution to the aforementioned problem [68]. Furthermore, attention mechanism-based methods can better identify the spatio-temporal characteristics of crops [69,70], thus improving the accuracy of crop yield prediction. Therefore, the objectives of this study are as follows. (1) To design a novel Attention-based Spatiotemporal Graph Neural Network (ASTGNN) model that integrates Graph Neural Networks and attention mechanisms, addressing the limitations of traditional methods in neglecting neighboring geospatial features. (2) To incorporate multi-source data, including satellite remote sensing, meteorological, and soil data, into the model to provide more comprehensive environmental information, thereby significantly improving the accuracy of winter wheat yield estimation. (3) To compare the performance of LSTM, CNN-LSTM, and ASTGNN models in capturing spatiotemporal cumulative data features, and to assess the effectiveness of multi-source versus single-source data in crop yield prediction.

2. Materials and Methodology

2.1. Study Area

Anhui is an important agricultural and grain-producing region in eastern China (Figure 1). The province stretches from 114°54′ to 119°37′ (east longitude) and from 29°41′ to 34°38′ (north latitude). Climatically, it spans temperate and subtropical zones, and abounds with mountainous and plain areas, with four distinctive seas and abundant rainfall, making it suitable for winter wheat growth. Winter wheat in the study area is usually harvested from late October to early November and from late May to early June the following year.

2.2. Dataset

For this study, we collected county-level data on winter wheat, including yields, planting area, satellite-based vegetation indices (e.g., solar-induced chlorophyll fluorescence (SIF), NDVI, and enhanced vegetation index (EVI)), climate conditions, and soil properties (Table 1). To integrate the growth stages of winter wheat, we selected data from October of the current year to May of the following year. This period encompasses the entire growth cycle of winter wheat. Additionally, monthly data were used to express all temporal characteristics, and the analysis was based on data from 2005 to 2020.

2.2.1. Remote Sensing Data

NDVI can help assess the health status and growth potential of vegetation [71]. By monitoring changes in NDVI values, we can gain insights into the growth conditions and variations in vegetation cover. EVI is an important remote sensing monitoring metric that primarily reflects seasonal and interannual variations in vegetation growth, and can be utilized to assess the health status, growth condition, and coverage of vegetation [72]. The NDVI and EVI were calculated based on MOD13Q1 images acquired from the Google Earth Engine platform. The temporal and spatial resolution of MOD13Q1 images is 16 days and 250 m, respectively. SIF can directly reflect the dynamic changes in the actual photosynthetic activity of plants. SIF data were acquired through the GOSIF data product [73]. The GOSIF data product is a global SIF data product with a resolution of 0.05°, which is produced through machine learning based on discrete SIF observations from the Orbiting Carbon Observatory-2. Drawing on previous studies [74,75,76], we selected three sources of vegetation index satellite data (NDVI, EVI, and SIF).

2.2.2. Meteorological Data

In this study, the selected meteorological data included potential evapotranspiration, precipitation, runoff, shortwave radiation, maximum temperature, and minimum temperature. All meteorological data were obtained from the Terraclimate dataset, which is a global high-spatial-resolution dataset that integrates the high-spatial-resolution WorldClim dataset with coarse-spatial-resolution data from the Climate Research Unit Ts4.0 and the Japanese 55-year Reanalysis [77].

2.2.3. Soil Data

SoilGrids is a global soil dataset [78], which mainly contains data on soil organic carbon content, sand content, clay content, and soil bulk density, with a resolution of 250 m and seven depths (0, 5, 15, 30, 60, 100, and 200 cm). Soil variables at different depths are highly correlated [38]. In this study, we selected the average values of the specified four soil variables at seven depths from SoilGrids.

2.2.4. Crop Yield Data and Planting Area Data

County-level crop yield data from 2005 to 2020 were obtained from the Anhui Statistical Yearbook (http://www.stats.gov.cn, accessed on 3 September 2024), compiled by the Anhui General Bureau of Investigation of the National Bureau of Statistics on the basis of field survey data. Crop planting area data were acquired from the National Ecosystem Science Data Center (http://www.nesdc.org.cn, accessed on 3 September 2024) [79]. We performed a preliminary quality inspection on the crop yield data to identify and remove the following outliers: (i) counties that lacked yield records in certain years during the study period and (ii) values outside the range of average values from 2005 to 2020 plus or minus the doubled variance.

2.3. Methodology

2.3.1. Technological Roadmap

To address the issue of varying pixel sizes in multi-source data, which can introduce uncertainty in validation results [80], we first standardized the spatial resolution of remote sensing data and environmental variables to 1 km and the temporal resolution to one month [81]. Continuous data, such as remote sensing and meteorological data, were resampled to 1 km using bilinear interpolation, while static data, such as soil data, were resampled using nearest neighbor interpolation. We then converted the resampled winter wheat planting area map, with a 1 km spatial resolution, into a binary image. Satellite vegetation indices and environmental variables were masked on an image-by-image basis according to winter wheat planting locations. These data were aggregated into monthly average variables at the county level before being imported into the deep learning model. Figure 2 shows the technical roadmap of this study.

2.3.2. GraphSAGE

GNNs provide a deep learning method for processing graph-structured data. A graph is a data structure that consists of nodes and edges. A GNN identifies the complex relationships and patterns in a graph structure by iteratively aggregating the characteristics of nodes and their neighbors, thereby implementing tasks such as node classification, link prediction, and graph classification. The characteristics of each node are updated at each layer by aggregating the characteristics of its neighbors. The aggregation is expressed as follows:
h v ( k ) = Aggregate { h u ( k 1 ) , u N ( v ) }
where h v ( k ) denotes the representation of node v at the k -th layer and N ( v ) denotes the neighbor set of node v .
However, most existing GNN methods are based on transductive learning (i.e., learning node embeddings directly on a fixed graph), whereas most graphs continually evolve. When the network structure changes or new nodes are available, transductive learning requires retraining (this is highly complex and may cause embedding to drift), making it difficult to implement transductive learning in machine learning systems that must rapidly generate embeddings for unknown nodes. Hence, transductive learning cannot be directly generalized to unknown nodes.
Graph sample and aggregation (GraphSAGE) is a general framework for learning node embeddings based on node characteristic information and aggregating the local neighbors of nodes [82]. In our study, “vertex” refers to the vertices of the county boundaries, which form the polygonal structure of the county. We use the centroid of the county polygon as the node, representing the central location of the county. The coordinates used are projected geographic coordinates to ensure spatial consistency and accuracy in graph computations. Specifically, GraphSAGE can sample the local neighbors of a vertex and aggregate vertex characteristics, rather than training a separate embedding for each vertex, thereby learning and aggregating neighbor information by sampling a portion of nodes. By learning the information of a node, GraphSAGE can construct an aggregation function based on the characteristics of its neighbor nodes. Moreover, by simulating characteristics and neighbor relationships through the forward propagation process, GraphSAGE can obtain the representation results of new nodes. Given that the adjacency matrices of the target county are sparse and the yield graphs of different counties continually evolve, GraphSAGE is particularly suitable for crop yield prediction. The aggregation function of GraphSAGE can involve mean aggregation, pooling aggregation, or LSTM aggregation, effectively capturing the complex relationships between nodes and enhancing the model’s adaptability to dynamic changes in graph structure. The forward propagation process of the GraphSAGE model is expressed as follows:
h v ( k ) = σ W Aggregate { h u ( k 1 ) , u Sample ( N ( v ) ) } + b
where W and b denote the learnable weight and bias and σ   denotes the activation function.

2.3.3. Constructing the ASTGNN Model

The key to crop yield forecasting lies in extracting complex non-linear relationships from historical data. Identifying the dynamic, local, and global characteristics of winter wheat yield variation is a substantial challenge. In this regard, previous models (e.g., CNNs and recurrent neural networks) have certain limitations. Owing to the limited receptive field, CNNs can only obtain local information, whereas recurrent neural networks face problems such as a vanishing gradient and explosion so cannot learn long-term dependent relationships and are only suitable for learning static time-dependent relationships; that is, they are not adaptable to the dynamic crop growth process. Therefore, we employ an ASTGNN framework, which is based on the attention mechanism, GNNs, and LSTM networks (Figure 3). The ASTGNN framework mainly comprises the following aspects: (1) input of multi-source data; (2) aggregation of node relationships between the target county and neighboring counties through the GraphSAGE module in the GNN; (3) two self-attention modules at the spatio-temporal attention layer for further capturing spatial and temporal characteristics; and (4) conversion and output of prediction results at the fully-connected layer.
Remote sensing and meteorological data obviously change in the short term, which is a dynamic change. In contrast, soil attributes remain nearly unchanged over time [83]. Therefore, in Figure 3, we use CNN(a) for extracting the dynamic change information of remote sensing and meteorological data, CNN(b) for extracting the static information of soil attributes that change insignificantly, and LSTMt for extracting the temporal features in the t-th year. Hc,t denotes the optimized embedding value after the information of adjacent counties of county c for the t-th year is aggregated by the GraphSAGE module; Zc,t denotes the new embedding value optimized by the spatio-temporal attention layer every year; and Yc,t denotes the predicted yield, which is outputted according to annual embedding values. The self-attention mechanism determines weights by calculating the similarity between queries, keys, and values as follows:
Q t = X t W Q t ,   K t = X t W K t ,   V t = X t W K t
where X t denotes the time step t at the temporal attention layer and input characteristics of spatial location t and W Q t , W K t , and W K t denote the weight matrices of query, key, and value, respectively. The self-attention score is expressed as follows:
Attention ( Q t , K , V ) = softmax Q t K T d k V
where K denotes the key vector matrix of all time steps or spatial locations and V denotes the value vector matrix of all time steps or spatial locations.

2.3.4. Model Training and Statistical Assessment

In this study, the Adam optimization algorithm was used for network training, which has the advantage of being able to adaptively adjust the learning rate to speed up the training process and improve stability. The initial learning rate is set to 0.01, and the total number of training rounds (epoch) is 100. In terms of hardware configuration, an NVIDIA 4090 GPU is used for model training, and the training time is about 3 h. To prevent model overfitting, regularization is introduced during training and combined with Early Stopping to automatically terminate training to ensure the model performs optimally on the validation set.
To quantify the effectiveness of the yield prediction model, we assessed the model performance by calculating the determination coefficient (R2), root mean square error (RMSE), and mean absolute error (MAE).
R 2 = 1 i = 1 n   y i y ^ i 2 i = 1 n   y i y ¯ 2
R M S E = i = 1 n   y i y ^ i 2 n
M A E = i = 1 n   | y i y ^ i | n
where y i and y ^ i denote the statistical crop yield and predicted crop yield of the i-th county in a specific year, respectively; y ¯ denotes the annual crop yield in a specific year; and n denotes the total number of counties.
To capture cumulative effects, this study starts annual yield forecasting from 2011 and incrementally incorporates data from 2005 to 2010. Specifically, data from all years prior to the prediction year are randomly divided into training (70%) and validation (30%) sets, while the data from the prediction year itself are used as the test set [58]. For example, when predicting for the year 2011, data from 2005 to 2010 are randomly divided into training and validation sets, while data from 2011 are used as the test set, and this process is repeated for subsequent years (Figure 4). This approach effectively utilizes historical data to develop a model with improved predictive performance while minimizing overfitting. Additionally, it retains the data sequences of each year, facilitating subsequent calibration of spatial changes. To further assess the model’s performance on time series data, we selected the monthly increment data for 2020 for evaluation.

3. Results

3.1. Temporal Distribution of Crop Yield Prediction

Table 2 and Figure 5 present the yield prediction results. As data accumulated over time, model performance continued to improve. The ASTGNN model exhibited the best robustness; annual and average R2 values were higher than those of the LSTM and CNN-LSTM models. Moreover, the annual R2 values of the LSTM and CNN-LSTM models were alternately higher in different years, whereas the average R2 value of the CNN-LSTM model was higher than that of the LSTM model. The RMSE and MAE values showed the same variation trends. According to the yield prediction performance of the three models at different growth stages using the monthly dataset of 2020 (Figure 6), the ASTGNN model again showed higher prediction performance than the other two models. The prediction accuracy of all three models increased steadily at the early growth stage (before February) and remained relatively stable from March to May. This indicates that the yield prediction results were reliable up to three months before the winter wheat reached maturity.

3.2. Spatial Distribution of Crop Yield Prediction

The predicted winter wheat yield of Anhui in 2020 was highly consistent with statistical data (Figure 7). The winter wheat-producing areas of Anhui are mainly distributed in northern Anhui, whereas winter wheat yields are low in southern Anhui owing to complex topography. Figure 8 shows the yield prediction errors of the three models compared with actual yields, which were less than 5% in most winter wheat-producing areas. Among the three models, the ASTGNN model provided the most accurate prediction results, whereas the LSTM and CNN-LSTM models exhibited large errors. The prediction errors for northern Anhui were relatively large for all models, whereas those for southern Anhui were relatively small. Moreover, large differences in yields between certain counties and their neighboring counties contributed to the prediction errors in LSTM and CNN-LSTM models but had a relatively small impact on the ASTGNN model results. This suggests that prediction accuracy can be improved by considering the geospatial information of neighboring counties.

3.3. Effect of Different Data Sources on Yield Prediction

Here, we used the ASTGNN model to assess the accuracy of winter wheat yield prediction based on multi-source data and single-source data (including remote sensing, meteorological, and soil property data). The prediction accuracy was highest when using multi-source data in 2020 (R2 = 0.70, RMSE = 0.21 t/ha, and MAE = 0.17 t/ha) (Table 3 and Figure 9). Among the single-source data types, the prediction accuracy was highest when using remote sensing data (R2 = 0.68, RMSE = 0.28 t/ha, and MAE = 0.23 t/ha). The accuracy of predictions based on soil property and meteorological data was lower than that based on remote sensing data (R2 = 0.64, RMSE = 0.30 t/ha, and MAE = 0.26 t/ha and R2 = 0.61, RMSE = 0.33 t/ha, and MAE = 0.26 t/ha, respectively). This is attributed to two reasons: (1) soil property data are static data, failing to reflect real-time changes in crop growth; and (2) meteorological data involve significant spatio-temporal changes, resulting in an indirect and complex effect on crop growth. Despite differences in the prediction accuracy between different types of single-source data, the data of different sources are complementary. This indicates that each single factor is correlated with, but has no decisive effect on, the crop yield prediction accuracy of the ASTGNN model. The overall accuracy of yield prediction can be improved by using a combination of multi-source data.

4. Discussion

We conducted an ablation study on the designed ASTGNN model. As shown in Table 4, when the spatiotemporal attention layer was not included, the model achieved an R2 of 0.67, an RMSE of 0.23 t/ha, and an MAE of 0.22 t/ha. After incorporating the spatiotemporal attention layer, the R2, RMSE, and MAE improved by approximately 4.48%, 8.70%, and 22.73%, respectively. These results demonstrate that the addition of the spatiotemporal attention layer significantly enhances the model’s accuracy in predicting winter wheat yield.
To date, numerous studies have predicted crop yields using deep-learning models. However, county-scale studies typically only consider the effects of multi-source data (e.g., remote sensing, meteorology, and soil type data) on crop yields, ignoring the effects of geospatial correlation and heterogeneity, which results in large model uncertainty. The accuracy of crop yield prediction can be improved significantly by introducing geospatial information into prediction models based on the graph embedding method, proving the effectiveness of geospatial information in crop yield prediction [67,84]. GNNs also show great potential for addressing spatial correlation [85,86].
A comparison of ASTGNN, LSTM, and CNN-LSTM models showed that model performance continued to improve as data were accumulated over time. That is, the use of more historical data can improve the generalization ability and prediction accuracy of prediction models [87]. In terms of annual crop yield prediction, the ASTGNN model showed higher robustness and accuracy than the LSTM and CNN-LSTM models. This suggests that the ASTGNN model has significant advantages in addressing complex spatio-temporal relationships and multi-source data. In addition, the ASTGNN model achieves higher R-squared and shows better stability than the LSTM and CNN-LSTM models three months before crop maturity. This suggests that the ASTGNN model showed advantages for predicting crop yields at different growth stages, especially before crops reached maturity. This finding is consistent with that of You et al. [88], namely, that prediction models based on spatio-temporal characteristics can better capture dynamic changes in crop growth.
The spatial distribution of the yield prediction results for 2020 showed the overestimation and underestimation of yields in the high-yield areas of northern Anhui and low-yield areas of southern Anhui, respectively. This is because northern Anhui has a large area of cultivated land, making it difficult to collect yield data and leading to data omissions or deviations. Moreover, the sample size of high-yield areas was limited, and the weighted average method was adopted to calculate the relative yield of each county, which ignored the specific trend in certain counties [58]. Additionally, prediction errors produced by the ASTGNN model were small, especially for southern Anhui, which has complex topography. This indicates that the prediction accuracy of the prediction model can be effectively improved by introducing geospatial neighbor information.
Among different sources of data, the prediction accuracy was highest when using multi-source data. Among single-source data, remote sensing data provided higher accuracy than soil property and meteorological data. This is because remote sensing data can capture the crop growth status (e.g., chlorophyll content and vegetation cover) of large-scale crops in real time and with high spatio-temporal resolution through satellite images. Specifically, remote sensing data can capture subtle changes in crop growth and directly indicate both the current health status and future growth trend of crops. Soil property and meteorological data are also major factors influencing crop growth. However, soil property data typically change slowly over time, making it insufficient to adequately reflect short-term variations within the crop growth cycle. Additionally, meteorological data, which cover a wide spatial scope and vary substantially over time, cannot provide crop growth information as timely and accurately as remote sensing data. Furthermore, multi-source data can provide more comprehensive and accurate crop growth information by combining the advantages of different sources of data (including remote sensing, soil property, and meteorological data) [83,89]. Therefore, our research findings suggest adopting a hybrid modeling strategy to achieve better prediction accuracy.
Although the yield prediction model developed in this study demonstrates strong performance, it has several limitations that need to be addressed. The study’s focus on winter wheat in Anhui Province, China, chosen for its complex geography—encompassing both hilly and plain areas—enhances the representativeness of the research. However, this focus may limit the applicability of the findings to other crops and regions. Future research should expand to diverse geographic areas and crop types to assess the generalizability of the proposed methods. Applying similar models in regions with varying climate and soil conditions could provide valuable insights into their effectiveness and adaptability. Additionally, issues related to data accuracy and security must be addressed [90]. The impact of certain features may be overlooked during the processing of multi-source data, potentially compromising the model’s performance in specific contexts. The model’s intricate structure and numerous parameters result in high computational costs, which can limit its applicability to large-scale datasets. Additionally, the complexity of spatial correlations might be underestimated, leading to prediction errors in both high-yield and low-yield areas, highlighting the need for further optimization to enhance predictive accuracy. Given the complexity of the dataset, future research should further explore the impact of different models and combinations of data sources on yield prediction. We acknowledge that the model’s predictive capability still has room for improvement. Future research should focus on refining multi-source data fusion techniques and enhancing data quality and coverage to minimize biases. Streamlining and simplifying model structures could significantly reduce computational costs and improve feasibility for large-scale applications. Furthermore, incorporating additional geographic and neighborhood features could enhance the model’s adaptability to complex terrains and spatial heterogeneity. Addressing concerns related to data accuracy and security, including the protection of privacy and information security during data collection, storage, and processing, is also crucial. Lastly, improving long-term prediction capabilities by integrating dynamic factors such as climate change and land use changes will bolster the model’s ability to forecast future variations, leading to more accurate crop yield predictions.

5. Conclusions

This study presents an innovative ASTGNN model that effectively integrates Graph Neural Networks and attention mechanisms, addressing the limitations of traditional methods in considering neighboring geospatial features. We combined multi-source data—including remote sensing, meteorological, and soil property data—with the ASTGNN model to predict winter wheat yields in Anhui, China. The ASTGNN model demonstrated superior performance in yield predictions from 2005 to 2020 (R2 = 0.70, RMSE = 0.21 t/ha, MAE = 0.17 t/ha), accurately forecasting yields three months prior to harvest. The incorporation of multi-source data and geospatial characteristics significantly enhanced the accuracy and reliability of the predictions. Our research not only enriches the existing literature but also provides valuable insights for accurate yield predictions of other crops. This study has significant implications for improving agricultural production management, advancing digital agriculture, and addressing the impacts of climate change on agriculture.

Author Contributions

Conceptualization, methodology, software, validation, formal analysis, investigation, writing—original draft, Z.Y.; conceptualization, methodology, software, validation, formal analysis, X.Z.; writing—review and editing, visualization, funding acquisition, Q.W., T.S., X.L., Y.H., L.W. and L.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key R&D Program of China (2023YFD1702105), the key project of the Joint Fund of the National Natural Science Foundation of China (U22A20567), and the Scientific Research Key Project of Anhui Province University (grant number: 2022AH053109).

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be addressed to the corresponding author.

Acknowledgments

We would like to thank the editors and reviewers for their professional advice on this manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Meghraoui, K.; Sebari, I.; Pilz, J.; Ait El Kadi, K.; Bensiali, S. Applied Deep Learning-Based Crop Yield Prediction: A Systematic Analysis of Current Developments and Potential Challenges. Technologies 2024, 12, 43. [Google Scholar] [CrossRef]
  2. Duarte Recalde, L.R.; Feierherd, G.; Mangonnet, J.; Murillo, M.V. Peasant Resistance in Times of Economic Affluence: Lessons From Paraguay. Comp. Political Stud. 2024, 00104140241237478. [Google Scholar] [CrossRef]
  3. Zhou, W.; Liu, Y.; Ata-Ul-Karim, S.T.; Ge, Q.; Li, X.; Xiao, J. Integrating Climate and Satellite Remote Sensing Data for Predicting County-Level Wheat Yield in China Using Machine Learning Methods. Int. J. Appl. Earth Obs. Geoinf. 2022, 111, 102861. [Google Scholar] [CrossRef]
  4. Ansarifar, J.; Wang, L.; Archontoulis, S.V. An Interaction Regression Model for Crop Yield Prediction. Sci. Rep. 2021, 11, 17754. [Google Scholar] [CrossRef] [PubMed]
  5. Warrick, A.W.; Gardner, W.R. Crop Yield as Affected by Spatial Variations of Soil and Irrigation. Water Resour. Res. 1983, 19, 181–186. [Google Scholar] [CrossRef]
  6. Karimzadeh Soureshjani, H.; Ghorbani Dehkordi, A.; Bahador, M. Temperature Effect on Yield of Winter and Spring Irrigated Crops. Agric. For. Meteorol. 2019, 279, 107664. [Google Scholar] [CrossRef]
  7. Maitah, M.; Malec, K.; Maitah, K. Influence of Precipitation and Temperature on Maize Production in the Czech Republic from 2002 to 2019. Sci. Rep. 2021, 11, 10467. [Google Scholar] [CrossRef] [PubMed]
  8. Madhukar, A.; Kumar, V.; Dashora, K. Temperature and Precipitation Are Adversely Affecting Wheat Yield in India. J. Water Clim. Chang. 2022, 13, 1631–1656. [Google Scholar] [CrossRef]
  9. Huang, T.; Döring, T.F.; Zhao, X.; Weiner, J.; Dang, P.; Zhang, M.; Zhang, M.; Siddique, K.H.M.; Schmid, B.; Qin, X. Cultivar Mixtures Increase Crop Yields and Temporal Yield Stability Globally. A Meta-Analysis. Agron. Sustain. Dev. 2024, 44, 28. [Google Scholar] [CrossRef]
  10. Talbot, M.; Milner, A.D.; Nutkins, M.A.E.; Law, J.R. Effect of Interference between Plots on Yield Performance in Crop Variety Trials. J. Agric. Sci. 1995, 124, 335–342. [Google Scholar] [CrossRef]
  11. Laidig, F.; Feike, T.; Klocke, B.; Macholdt, J.; Miedaner, T.; Rentel, D.; Piepho, H.P. Yield Reduction Due to Diseases and Lodging and Impact of Input Intensity on Yield in Variety Trials in Five Cereal Crops. Euphytica 2022, 218, 150. [Google Scholar] [CrossRef]
  12. Jehangir, I.A.; Hussain, A.; Sofi, N.R.; Wani, S.H.; Ali, O.M.; Abdel Latef, A.A.H.; Raja, W.; Bhat, M.A. Crop Establishment Methods and Weed Management Practices Affect Grain Yield and Weed Dynamics in Temperate Rice. Agronomy 2021, 11, 2137. [Google Scholar] [CrossRef]
  13. Araya, A.; Prasad, P.V.V.; Ciampitti, I.A.; Jha, P.K. Using Crop Simulation Model to Evaluate Influence of Water Management Practices and Multiple Cropping Systems on Crop Yields: A Case Study for Ethiopian Highlands. Field Crops Res. 2021, 260, 108004. [Google Scholar] [CrossRef]
  14. Bregaglio, S.; Willocquet, L.; Kersebaum, K.C.; Ferrise, R.; Stella, T.; Ferreira, T.B.; Pavan, W.; Asseng, S.; Savary, S. Comparing Process-Based Wheat Growth Models in Their Simulation of Yield Losses Caused by Plant Diseases. Field Crops Res. 2021, 265, 108108. [Google Scholar] [CrossRef]
  15. Loit, K.; Soonvald, L.; Adamson, K.; Runno-Paurson, E.; Tedersoo, L.; Astover, A. Assessing the Effect of Commercial Arbuscular Mycorrhizal Fungal Inoculum on Potato Plant Disease Incidence, Yield and the Indigenous Root Fungal Community Composition. BioControl 2023, 68, 537–547. [Google Scholar] [CrossRef]
  16. Mehmood, N.; Saeed, M.; Zafarullah, S.; Hyder, S.; Rizvi, Z.F.; Gondal, A.S.; Jamil, N.; Iqbal, R.; Ali, B.; Ercisli, S.; et al. Multifaceted Impacts of Plant-Beneficial Pseudomonas Spp. in Managing Various Plant Diseases and Crop Yield Improvement. ACS Omega 2023, 8, 22296–22315. [Google Scholar] [CrossRef] [PubMed]
  17. Geipel, J.; Link, J.; Claupein, W. Combined Spectral and Spatial Modeling of Corn Yield Based on Aerial Images and Crop Surface Models Acquired with an Unmanned Aircraft System. Remote Sens. 2014, 6, 10335–10355. [Google Scholar] [CrossRef]
  18. Alesso, C.A.; Cipriotti, P.A.; Bollero, G.A.; Martin, N.F. Experimental Designs and Estimation Methods for On-Farm Research: A Simulation Study of Corn Yields at Field Scale. Agron. J. 2019, 111, 2724–2735. [Google Scholar] [CrossRef]
  19. Willocquet, L.; Meza, W.R.; Dumont, B.; Klocke, B.; Feike, T.; Kersebaum, K.C.; Meriggi, P.; Rossi, V.; Ficke, A.; Djurle, A.; et al. An Outlook on Wheat Health in Europe from a Network of Field Experiments. Crop Prot. 2021, 139, 105335. [Google Scholar] [CrossRef]
  20. Zhou, K.; Cao, L.; Shen, X.; Wang, G. Novel Spectral Indices for Enhanced Estimations of 3-Dimentional Flavonoid Contents for Ginkgo Plantations Using UAV-Borne LiDAR and Hyperspectral Data. Remote Sens. Environ. 2023, 299, 113882. [Google Scholar] [CrossRef]
  21. Vintrou, E.; Desbrosse, A.; Bégué, A.; Traoré, S.; Baron, C.; Lo Seen, D. Crop Area Mapping in West Africa Using Landscape Stratification of MODIS Time Series and Comparison with Existing Global Land Products. Int. J. Appl. Earth Obs. Geoinf. 2012, 14, 83–93. [Google Scholar] [CrossRef]
  22. Sapkota, S.; Paudyal, D.R. Growth Monitoring and Yield Estimation of Maize Plant Using Unmanned Aerial Vehicle (UAV) in a Hilly Region. Sensors 2023, 23, 5432. [Google Scholar] [CrossRef]
  23. Gumma, M.K.; Kadiyala, M.D.M.; Panjala, P.; Ray, S.S.; Akuraju, V.R.; Dubey, S.; Smith, A.P.; Das, R.; Whitbread, A.M. Assimilation of Remote Sensing Data into Crop Growth Model for Yield Estimation: A Case Study from India. J. Indian. Soc. Remote Sens. 2022, 50, 257–270. [Google Scholar] [CrossRef]
  24. Hassan, S.S.; Goheer, M.A. Modeling and Monitoring Wheat Crop Yield Using Geospatial Techniques: A Case Study of Potohar Region, Pakistan. J. Indian Soc. Remote Sens. 2021, 49, 1331–1342. [Google Scholar] [CrossRef]
  25. Yu, B.; Shang, S. Multi-Year Mapping of Major Crop Yields in an Irrigation District from High Spatial and Temporal Resolution Vegetation Index. Sensors 2018, 18, 3787. [Google Scholar] [CrossRef]
  26. Ilyas, Q.M.; Ahmad, M.; Mehmood, A. Automated Estimation of Crop Yield Using Artificial Intelligence and Remote Sensing Technologies. Bioengineering 2023, 10, 125. [Google Scholar] [CrossRef]
  27. Dubey, S.K.; Sharma, D. Assessment of Climate Change Impact on Yield of Major Crops in the Banas River Basin, India. Sci. Total Environ. 2018, 635, 10–19. [Google Scholar] [CrossRef] [PubMed]
  28. Bolton, D.K.; Friedl, M.A. Forecasting Crop Yield Using Remotely Sensed Vegetation Indices and Crop Phenology Metrics. Agric. For. Meteorol. 2013, 173, 74–84. [Google Scholar] [CrossRef]
  29. Stepanov, A.; Dubrovin, K.; Sorokin, A.; Aseeva, T. Predicting Soybean Yield at the Regional Scale Using Remote Sensing and Climatic Data. Remote Sens. 2020, 12, 1936. [Google Scholar] [CrossRef]
  30. Kolovos, A.; Smith, L.M.; Schwab-McCoy, A.; Gengler, S.; Yu, H.-L. Emerging Patterns in Multi-Sourced Data Modeling Uncertainty. Spat. Stat. 2016, 18, 300–317. [Google Scholar] [CrossRef]
  31. Chen, C.; Liang, J.; Ma, F.; Glass, L.M.; Sun, J.; Xiao, C. UNITE: Uncertainty-Based Health Risk Prediction Leveraging Multi-Sourced Data. Proc. Web Conf. 2021, 2021, 217–226. [Google Scholar]
  32. Ermolieva, T.; Havlik, P.; Lessa-Derci-Augustynczik, A.; Frank, S.; Balkovic, J.; Skalsky, R.; Deppermann, A.; Nakhavali, M.; Komendantova, N.; Kahil, T.; et al. Tracking the Dynamics and Uncertainties of Soil Organic Carbon in Agricultural Soils Based on a Novel Robust Meta-Model Framework Using Multisource Data. Sustainability 2024, 16, 6849. [Google Scholar] [CrossRef]
  33. Liu, Y.; Guo, Z.; Zhang, X.; Zhang, R.; Zhou, Z. (ChinaVis 2019) Uncertainty Visualization in Stratigraphic Correlation Based on Multi-Source Data Fusion. J. Vis. 2019, 22, 1021–1038. [Google Scholar] [CrossRef]
  34. Debaeke, P.; Attia, F.; Champolivier, L.; Dejoux, J.-F.; Micheneau, A.; Bitar, A.A.; Trépos, R. Forecasting Sunflower Grain Yield Using Remote Sensing Data and Statistical Models. Eur. J. Agron. 2023, 142, 126677. [Google Scholar] [CrossRef]
  35. Romanovska, P.; Schauberger, B.; Gornott, C. Wheat Yields in Kazakhstan Can Successfully Be Forecasted Using a Statistical Crop Model. Eur. J. Agron. 2023, 147, 126843. [Google Scholar] [CrossRef]
  36. Lecerf, R.; Ceglar, A.; López-Lozano, R.; Van Der Velde, M.; Baruth, B. Assessing the Information in Crop Model and Meteorological Indicators to Forecast Crop Yield over Europe. Agric. Syst. 2019, 168, 191–202. [Google Scholar] [CrossRef]
  37. Shuai, G.; Basso, B. Subfield Maize Yield Prediction Improves When In-Season Crop Water Deficit Is Included in Remote Sensing Imagery-Based Models. Remote Sens. Environ. 2022, 272, 112938. [Google Scholar] [CrossRef]
  38. Zhuang, H.; Zhang, Z.; Cheng, F.; Han, J.; Luo, Y.; Zhang, L.; Cao, J.; Zhang, J.; He, B.; Xu, J.; et al. Integrating Data Assimilation, Crop Model, and Machine Learning for Winter Wheat Yield Forecasting in the North China Plain. Agric. For. Meteorol. 2024, 347, 109909. [Google Scholar] [CrossRef]
  39. Cai, Y.; Guan, K.; Lobell, D.; Potgieter, A.B.; Wang, S.; Peng, J.; Xu, T.; Asseng, S.; Zhang, Y.; You, L.; et al. Integrating Satellite and Climate Data to Predict Wheat Yield in Australia Using Machine Learning Approaches. Agric. For. Meteorol. 2019, 274, 144–159. [Google Scholar] [CrossRef]
  40. Yu, W.; Yang, G.; Li, D.; Zheng, H.; Yao, X.; Zhu, Y.; Cao, W.; Qiu, L.; Cheng, T. Improved Prediction of Rice Yield at Field and County Levels by Synergistic Use of SAR, Optical and Meteorological Data. Agric. For. Meteorol. 2023, 342, 109729. [Google Scholar] [CrossRef]
  41. Rafique, R.; Ahmad, T.; Ahmed, M.; Khan, M.A. Adapting the Process Based STICS Model to Simulate Phenology and Yield of Table Grapes- a High Value Fruit Crop in a New Emerging Viticulture Zone of South Asia. Sci. Hortic. 2024, 336, 113419. [Google Scholar] [CrossRef]
  42. Fattori Junior, I.M.; dos Santos Vianna, M.; Marin, F.R. Assimilating Leaf Area Index Data into a Sugarcane Process-Based Crop Model for Improving Yield Estimation. Eur. J. Agron. 2022, 136, 126501. [Google Scholar] [CrossRef]
  43. Li, Y.; Guan, K.; Yu, A.; Peng, B.; Zhao, L.; Li, B.; Peng, J. Toward Building a Transparent Statistical Model for Improving Crop Yield Prediction: Modeling Rainfed Corn in the U.S. Field Crops Res. 2019, 234, 55–65. [Google Scholar] [CrossRef]
  44. Blanc, É. Statistical Emulators of Maize, Rice, Soybean and Wheat Yields from Global Gridded Crop Models. Agric. For. Meteorol. 2017, 236, 145–161. [Google Scholar] [CrossRef]
  45. Gavahi, K.; Abbaszadeh, P.; Moradkhani, H. DeepYield: A Combined Convolutional Neural Network with Long Short-Term Memory for Crop Yield Forecasting. Expert. Syst. Appl. 2021, 184, 115511. [Google Scholar] [CrossRef]
  46. Basso, B.; Cammarano, D.; Carfagna, E. Review of Crop Yield Forecasting Methods and Early Warning Systems. In Proceedings of the First Meeting of the Scientific Advisory Committee of the Global Strategy to Improve Agricultural and Rural Statistics; FAO Headquarters: Rome, Italy, 2013. [Google Scholar]
  47. Horie, T.; Yajima, M.; Nakagawa, H. Yield Forecasting. Agric. Syst. 1992, 40, 211–236. [Google Scholar] [CrossRef]
  48. Jeong, J.H.; Resop, J.P.; Mueller, N.D.; Fleisher, D.H.; Yun, K.; Butler, E.E.; Timlin, D.J.; Shim, K.-M.; Gerber, J.S.; Reddy, V.R.; et al. Random Forests for Global and Regional Crop Yield Predictions. PLoS ONE 2016, 11, e0156571. [Google Scholar] [CrossRef] [PubMed]
  49. Dhaka, V.S.; Kundu, N.; Rani, G.; Zumpano, E.; Vocaturo, E. Role of Internet of Things and Deep Learning Techniques in Plant Disease Detection and Classification: A Focused Review. Sensors 2023, 23, 7877. [Google Scholar] [CrossRef] [PubMed]
  50. Disease detection, severity prediction, and crop loss estimation in MaizeCrop using deep learning. Artif. Intell. Agric. 2022, 6, 276–291. [CrossRef]
  51. Gangwar, A.; Dhaka, V.S.; Rani, G.; Khandelwal, S.; Zumpano, E.; Vocaturo, E. Time and Space Efficient Multi-Model Convolution Vision Transformer for Tomato Disease Detection from Leaf Images with Varied Backgrounds. Comput. Mater. Contin. 2024, 79, 117–142. [Google Scholar] [CrossRef]
  52. Gill, H.S.; Bath, B.S.; Singh, R.; Riar, A.S. Wheat Crop Classification Using Deep Learning. Multimed. Tools Appl. 2024. [Google Scholar] [CrossRef]
  53. Kuwata, K.; Shibasaki, R. Estimating Crop Yields with Deep Learning and Remotely Sensed Data. In Proceedings of the 2015 IEEE International Geoscience and Remote Sensing Symposium (IGARSS), Milan, Italy, 26–31 July 2015; pp. 858–861. [Google Scholar]
  54. Li, X.; Geng, H.; Zhang, L.; Peng, S.; Xin, Q.; Huang, J.; Li, X.; Liu, S.; Wang, Y. Improving Maize Yield Prediction at the County Level from 2002 to 2015 in China Using a Novel Deep Learning Approach. Comput. Electron. Agric. 2022, 202, 107356. [Google Scholar] [CrossRef]
  55. Wang, J.; Wang, P.; Tian, H.; Tansey, K.; Liu, J.; Quan, W. A Deep Learning Framework Combining CNN and GRU for Improving Wheat Yield Estimates Using Time Series Remotely Sensed Multi-Variables. Comput. Electron. Agric. 2023, 206, 107705. [Google Scholar] [CrossRef]
  56. Sun, J.; Di, L.; Sun, Z.; Shen, Y.; Lai, Z. County-Level Soybean Yield Prediction Using Deep CNN-LSTM Model. Sensors 2019, 19, 4363. [Google Scholar] [CrossRef] [PubMed]
  57. Zhou, S.; Xu, L.; Chen, N. Rice Yield Prediction in Hubei Province Based on Deep Learning and the Effect of Spatial Heterogeneity. Remote Sens. 2023, 15, 1361. [Google Scholar] [CrossRef]
  58. Chen, P.; Li, Y.; Liu, X.; Tian, Y.; Zhu, Y.; Cao, W.; Cao, Q. Improving Yield Prediction Based on Spatio-Temporal Deep Learning Approaches for Winter Wheat: A Case Study in Jiangsu Province, China. Comput. Electron. Agric. 2023, 213, 108201. [Google Scholar] [CrossRef]
  59. Nevavuori, P.; Narra, N.; Lipping, T. Crop Yield Prediction with Deep Convolutional Neural Networks. Comput. Electron. Agric. 2019, 163, 104859. [Google Scholar] [CrossRef]
  60. Jiang, H.; Hu, H.; Zhong, R.; Xu, J.; Xu, J.; Huang, J.; Wang, S.; Ying, Y.; Lin, T. A Deep Learning Approach to Conflating Heterogeneous Geospatial Data for Corn Yield Estimation: A Case Study of the US Corn Belt at the County Level. Glob. Chang. Biol. 2020, 26, 1754–1766. [Google Scholar] [CrossRef]
  61. Wang, X.; Huang, J.; Feng, Q.; Yin, D. Winter Wheat Yield Prediction at County Level and Uncertainty Analysis in Main Wheat-Producing Regions of China with Deep Learning Approaches. Remote Sens. 2020, 12, 1744. [Google Scholar] [CrossRef]
  62. Tian, H.; Wang, P.; Tansey, K.; Zhang, J.; Zhang, S.; Li, H. An LSTM Neural Network for Improving Wheat Yield Estimates by Integrating Remote Sensing Data and Meteorological Data in the Guanzhong Plain, PR China. Agric. For. Meteorol. 2021, 310, 108629. [Google Scholar] [CrossRef]
  63. Zhang, L.; Zhang, Z.; Luo, Y.; Cao, J.; Xie, R.; Li, S. Integrating Satellite-Derived Climatic and Vegetation Indices to Predict Smallholder Maize Yield Using Deep Learning. Agric. For. Meteorol. 2021, 311, 108666. [Google Scholar] [CrossRef]
  64. Sharma, S.; Rai, S.; Krishnan, N.C. Wheat Crop Yield Prediction Using Deep LSTM Model. arXiv 2020, arXiv:2011.01498. [Google Scholar]
  65. Gastli, M.S.; Nassar, L.; Karray, F. Satellite Images and Deep Learning Tools for Crop Yield Prediction and Price Forecasting. In Proceedings of the 2021 International Joint Conference on Neural Networks (IJCNN), Shenzhen, China, 18–22 July 2021; Available online: https://ieeexplore.ieee.org/abstract/document/9534388 (accessed on 19 September 2024).
  66. Sun, J.; Lai, Z.; Di, L.; Sun, Z.; Tao, J.; Shen, Y. Multilevel Deep Learning Network for County-Level Corn Yield Estimation in the U.S. Corn Belt. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2020, 13, 5048–5060. [Google Scholar] [CrossRef]
  67. Han, D.; Wang, P.; Tansey, K.; Zhang, Y.; Li, H. A Graph-Based Deep Learning Framework for Field Scale Wheat Yield Estimation. Int. J. Appl. Earth Obs. Geoinf. 2024, 129, 103834. [Google Scholar] [CrossRef]
  68. Scarselli, F.; Gori, M.; Tsoi, A.C.; Hagenbuchner, M.; Monfardini, G. The Graph Neural Network Model. IEEE Trans. Neural Netw. 2009, 20, 61–80. [Google Scholar] [CrossRef]
  69. Tian, H.; Wang, P.; Tansey, K.; Han, D.; Zhang, J.; Zhang, S.; Li, H. A Deep Learning Framework under Attention Mechanism for Wheat Yield Estimation Using Remotely Sensed Indices in the Guanzhong Plain, PR China. Int. J. Appl. Earth Obs. Geoinf. 2021, 102, 102375. [Google Scholar] [CrossRef]
  70. Priya, S.S.; Dathan, R.A. Attention Based Peephole LSTM Model for Soybean Crop Yield Prediction. J. Phys. Conf. Ser. 2023, 2571, 012013. [Google Scholar] [CrossRef]
  71. Jordan, C.F. Derivation of Leaf-Area Index from Quality of Light on the Forest Floor. Ecology 1969, 50, 663–666. [Google Scholar] [CrossRef]
  72. Roy, B. Optimum Machine Learning Algorithm Selection for Forecasting Vegetation Indices: MODIS NDVI & EVI. Remote Sens. Appl. : Soc. Environ. 2021, 23, 100582. [Google Scholar] [CrossRef]
  73. Li, X.; Xiao, J. A Global, 0.05-Degree Product of Solar-Induced Chlorophyll Fluorescence Derived from OCO-2, MODIS, and Reanalysis Data. Remote Sens. 2019, 11, 517. [Google Scholar] [CrossRef]
  74. Joshi, A.; Pradhan, B.; Chakraborty, S.; Behera, M.D. Winter Wheat Yield Prediction in the Conterminous United States Using Solar-Induced Chlorophyll Fluorescence Data and XGBoost and Random Forest Algorithm. Ecol. Inform. 2023, 77, 102194. [Google Scholar] [CrossRef]
  75. Li, Z.; Ding, L.; Xu, D. Exploring the Potential Role of Environmental and Multi-Source Satellite Data in Crop Yield Prediction across Northeast China. Sci. Total Environ. 2022, 815, 152880. [Google Scholar] [CrossRef] [PubMed]
  76. Zhang, L.; Zhang, Z.; Luo, Y.; Cao, J.; Tao, F. Combining Optical, Fluorescence, Thermal Satellite, and Environmental Data to Predict County-Level Maize Yield in China Using Machine Learning Approaches. Remote Sens. 2019, 12, 21. [Google Scholar] [CrossRef]
  77. Abatzoglou, J.T.; Dobrowski, S.Z.; Parks, S.A.; Hegewisch, K.C. TerraClimate, a High-Resolution Global Dataset of Monthly Climate and Climatic Water Balance from 1958–2015. Sci. Data 2018, 5, 170191. [Google Scholar] [CrossRef] [PubMed]
  78. Hengl, T.; Jesus, J.M.; Heuvelink, G.B.M.; Gonzalez, M.R.; Kilibarda, M.; Blagotić, A.; Shangguan, W.; Wright, M.N.; Geng, X.; Bauer-Marschallinger, B.; et al. SoilGrids250m: Global gridded soil information based on machine learning. PLoS ONE 2017, 12, e0169748. [Google Scholar] [CrossRef] [PubMed]
  79. Cao, J.; Zhang, Z.; Tao, F.; Zhang, L.; Luo, Y.; Han, J.; Li, Z. Identifying the Contributions of Multi-Source Data for Winter Wheat Yield Prediction in China. Remote Sens. 2020, 12, 750. [Google Scholar] [CrossRef]
  80. Bayat, B.; Camacho, F.; Nickeson, J.; Cosh, M.; Bolten, J.; Vereecken, H.; Montzka, C. Toward Operational Validation Systems for Global Satellite-Based Terrestrial Essential Climate Variables. Int. J. Appl. Earth Obs. Geoinf. 2021, 95, 102240. [Google Scholar] [CrossRef]
  81. Wang, Y.Q.; Leng, P.; Shang, G.F.; Zhang, X.; Li, Z.L. Sun-Induced Chlorophyll Fluorescence Is Superior to Satellite Vegetation Indices for Predicting Summer Maize Yield under Drought Conditions. Comput. Electron. Agric. 2023, 205, 107615. [Google Scholar] [CrossRef]
  82. Hamilton, W.L.; Ying, R.; Leskovec, J. Inductive Representation Learning on Large Graphs. In Proceedings of the 31st Annual Conference on Neural Information Processing Systems (NIPS 2017), Long Beach, CA, USA, 4–9 December 2017. [Google Scholar]
  83. Huang, H.; Huang, J.; Feng, Q.; Liu, J.; Li, X.; Wang, X.; Niu, Q. Developing a Dual-Stream Deep-Learning Neural Network Model for Improving County-Level Winter Wheat Yield Estimates in China. Remote Sens. 2022, 14, 5280. [Google Scholar] [CrossRef]
  84. Fan, J.; Bai, J.; Li, Z.; Ortiz-Bobea, A.; Gomes, C.P. A GNN-RNN Approach for Harnessing Geospatial and Temporal Information: Application to Crop Yield Prediction. Proc. AAAI Conf. Artif. Intell. 2022, 36, 11873–11881. [Google Scholar] [CrossRef]
  85. Zhao, B.; He, X.; Ran, S.; Zhang, Y.; Cheng, C. Spatial Correlation Learning Based on Graph Neural Network for Medium-Term Wind Power Forecasting. Energy 2024, 296, 131164. [Google Scholar] [CrossRef]
  86. Guo, Y.; Peng, Y.; Hao, R.; Tang, X. Capturing Spatial–Temporal Correlations with Attention Based Graph Convolutional Network for Network Traffic Prediction. J. Netw. Comput. Appl. 2023, 220, 103746. [Google Scholar] [CrossRef]
  87. Chen, L.; Min, Y.; Zhang, M.; Karbasi, A. More Data Can Expand the Generalization Gap between Adversarially Robust and Standard Models. In Proceedings of the International Conference on Machine Learning, Virtual, 13–18 July 2020. [Google Scholar]
  88. You, Y.; Tian, H.; Pan, S.; Shi, H.; Bian, Z.; Gurgel, A.; Huang, Y.; Kicklighter, D.; Liang, X.-Z.; Lu, C.; et al. Incorporating Dynamic Crop Growth Processes and Management Practices into a Terrestrial Biosphere Model for Simulating Crop Production in the United States: Toward a Unified Modeling Framework. Agric. For. Meteorol. 2022, 325, 109144. [Google Scholar] [CrossRef]
  89. Wang, Y.; Zhang, Z.; Feng, L.; Du, Q.; Runge, T. Combining Multi-Source Data and Machine Learning Approaches to Predict Winter Wheat Yield in the Conterminous United States. Remote Sens. 2020, 12, 1232. [Google Scholar] [CrossRef]
  90. Vocaturo, E.; Rani, G.; Dhaka, V.S. AI-Driven Agriculture: Opportunities and Challenges. In Proceedings of the 2023 IEEE International Conference on Big Data (BigData), Sorrento, Italy, 15–18 December 2023. [Google Scholar]
Figure 1. Location of the study area.
Figure 1. Location of the study area.
Agronomy 14 02262 g001
Figure 2. Technological roadmap of this study. EVI: enhanced vegetation index; NDVI: normalized difference vegetation index; SIF: solar-induced chlorophyll fluorescence; LSTM: long short-term memory network; ASTGNN: attention-based spatio-temporal Graph Neural Network; CNN-LSTM: convolutional neural network-LSTM; R2: coefficient of determination; RMSE: root mean squared error; MAE: mean absolute error.
Figure 2. Technological roadmap of this study. EVI: enhanced vegetation index; NDVI: normalized difference vegetation index; SIF: solar-induced chlorophyll fluorescence; LSTM: long short-term memory network; ASTGNN: attention-based spatio-temporal Graph Neural Network; CNN-LSTM: convolutional neural network-LSTM; R2: coefficient of determination; RMSE: root mean squared error; MAE: mean absolute error.
Agronomy 14 02262 g002
Figure 3. Schematic of the ASTGNN model employed in this study. (a) Spatial Attention, (b) Temporal Attention.
Figure 3. Schematic of the ASTGNN model employed in this study. (a) Spatial Attention, (b) Temporal Attention.
Agronomy 14 02262 g003
Figure 4. Data partitioning diagram (the dashed lines in the figure are not drawn to scale).
Figure 4. Data partitioning diagram (the dashed lines in the figure are not drawn to scale).
Agronomy 14 02262 g004
Figure 5. Comparison of true and predicted values of different models for different years in different counties and districts. (a) LSTM. (b) CNN-LSTM. (c) ASTGNN.
Figure 5. Comparison of true and predicted values of different models for different years in different counties and districts. (a) LSTM. (b) CNN-LSTM. (c) ASTGNN.
Agronomy 14 02262 g005
Figure 6. Monthly prediction performance of LSTM, CNN-LSTM, and ASTGNN models in 2020 (M1 to M8: October 2019 to May 2020).
Figure 6. Monthly prediction performance of LSTM, CNN-LSTM, and ASTGNN models in 2020 (M1 to M8: October 2019 to May 2020).
Agronomy 14 02262 g006
Figure 7. Spatial distribution of winter wheat yields in 2020: (a) actual yield; (b) LSTM model; (c) CNN-LSTM model; (d) ASTGNN model.
Figure 7. Spatial distribution of winter wheat yields in 2020: (a) actual yield; (b) LSTM model; (c) CNN-LSTM model; (d) ASTGNN model.
Agronomy 14 02262 g007
Figure 8. Relative errors of yield prediction of different models in 2020: (a) LSTM model; (b) CNN-LSTM model; (c) ASTGNN model.
Figure 8. Relative errors of yield prediction of different models in 2020: (a) LSTM model; (b) CNN-LSTM model; (c) ASTGNN model.
Agronomy 14 02262 g008
Figure 9. Scatterplot of four data sources: (a) multi-source data; (b) remote sensing data; (c) soil data; (d) meteorological data.
Figure 9. Scatterplot of four data sources: (a) multi-source data; (b) remote sensing data; (c) soil data; (d) meteorological data.
Agronomy 14 02262 g009
Table 1. Data used in this study.
Table 1. Data used in this study.
DataAbbreviationSourceSpatial ResolutionTemporal Resolution
Remote sensing data
NDVI MODIS (MOD13Q1)250 m16 days
EVI MODIS (MOD13Q1)250 m1 days
SIF GOSIF0.05°monthly
Meteorological data
Potential evapotranspirationpetTerraClimate4 kmmonthly
PrecipitationpptTerraClimate4 kmmonthly
RunoffqTerraClimate4 kmmonthly
Shortwave radiationsradTerraClimate4 kmmonthly
Maximum temperaturetmaxTerraClimate4 kmmonthly
Minimum temperaturetminTerraClimate4 kmmonthly
Soil data
Organic carbon contentsocSoilGrids 100 m
Sand contentsandSoilGrids100 m
Clay contentclaySoilGrids100 m
Soil bulk densitybulkSoilGrids100 m
Yield data Anhui Statistical Yearbookcounty-scaleannual
Planting area data National Ecosystem Science Data Center30 m
Table 2. Annual prediction performance of LSTM, CNN-LSTM, and ASTGNN models from 2011 to 2020.
Table 2. Annual prediction performance of LSTM, CNN-LSTM, and ASTGNN models from 2011 to 2020.
YearLSTMCNN-LSTMASTGNN
R2RMSE
(t/ha)
MAE (t/ha)R2RMSE (t/ha)MAE (t/ha)R2RMSE
(t/ha)
MAE (t/ha)
20110.49 0.44 0.31 0.55 0.34 0.26 0.56 0.35 0.27
20120.57 0.35 0.25 0.60 0.28 0.25 0.68 0.26 0.22
20130.56 0.35 0.25 0.57 0.31 0.24 0.60 0.28 0.24
20140.62 0.28 0.24 0.65 0.27 0.23 0.67 0.27 0.22
20150.67 0.27 0.23 0.66 0.27 0.23 0.68 0.26 0.22
20160.68 0.26 0.22 0.65 0.27 0.22 0.70 0.23 0.21
20170.59 0.31 0.26 0.70 0.24 0.21 0.72 0.21 0.20
20180.66 0.27 0.23 0.70 0.24 0.17 0.73 0.21 0.18
20190.68 0.27 0.23 0.68 0.27 0.23 0.71 0.22 0.19
20200.66 0.27 0.23 0.69 0.26 0.22 0.70 0.21 0.17
Average0.62 0.31 0.25 0.65 0.28 0.23 0.68 0.25 0.21
Table 3. Accuracy of winter wheat yield prediction based on four types of data sources.
Table 3. Accuracy of winter wheat yield prediction based on four types of data sources.
Data TypeR2RMSE (t/ha)MAE (t/ha)
Multi-source data0.700.210.17
Remote sensing data0.680.280.23
Soil data0.640.300.26
Meteorological data0.610.330.26
Table 4. Ablation analysis of the spatiotemporal attention layer.
Table 4. Ablation analysis of the spatiotemporal attention layer.
Ablation AnalysesR2RMSE (t/ha)MAE (t/ha)
Attention Deficit Layer (ADL)0.670.230.22
Attention Span Layer (MSL)0.700.210.17
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ye, Z.; Zhai, X.; She, T.; Liu, X.; Hong, Y.; Wang, L.; Zhang, L.; Wang, Q. Winter Wheat Yield Prediction Based on the ASTGNN Model Coupled with Multi-Source Data. Agronomy 2024, 14, 2262. https://doi.org/10.3390/agronomy14102262

AMA Style

Ye Z, Zhai X, She T, Liu X, Hong Y, Wang L, Zhang L, Wang Q. Winter Wheat Yield Prediction Based on the ASTGNN Model Coupled with Multi-Source Data. Agronomy. 2024; 14(10):2262. https://doi.org/10.3390/agronomy14102262

Chicago/Turabian Style

Ye, Zhicheng, Xu Zhai, Tianlong She, Xiaoyan Liu, Yuanyuan Hong, Lihui Wang, Lili Zhang, and Qiang Wang. 2024. "Winter Wheat Yield Prediction Based on the ASTGNN Model Coupled with Multi-Source Data" Agronomy 14, no. 10: 2262. https://doi.org/10.3390/agronomy14102262

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop