Next Article in Journal
Assessing Yield, Biomass Production, and Forage Quality of Red Clover (Trifolium pratense L.) in Agroforestry System: One-Year Study in Szarvas, Hungary
Previous Article in Journal
Preparation, Characterization, and Testing of Compost Tea Derived from Seaweed and Fish Residues
Previous Article in Special Issue
Research on Rapeseed Above-Ground Biomass Estimation Based on Spectral and LiDAR Data
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Remote Sensing Data Assimilation in Crop Growth Modeling from an Agricultural Perspective: New Insights on Challenges and Prospects

College of Information Science and Technology, Gansu Agricultural University, Lanzhou 730070, China
*
Author to whom correspondence should be addressed.
Agronomy 2024, 14(9), 1920; https://doi.org/10.3390/agronomy14091920
Submission received: 18 July 2024 / Revised: 21 August 2024 / Accepted: 23 August 2024 / Published: 27 August 2024
(This article belongs to the Special Issue Remote Sensing Applications in Crop Monitoring and Modelling)

Abstract

:
The frequent occurrence of global climate change and natural disasters highlights the importance of precision agricultural monitoring, yield forecasting, and early warning systems. The data assimilation method provides a new possibility to solve the problems of low accuracy of yield prediction, strong dependence on the field, and poor adaptability of the model in traditional agricultural applications. Therefore, this study makes a systematic literature retrieval based on Web of Science, Scopus, Google Scholar, and PubMed databases, introduces in detail the assimilation strategies based on many new remote sensing data sources, such as satellite constellation, UAV, ground observation stations, and mobile platforms, and compares and analyzes the progress of assimilation models such as compulsion method, model parameter method, state update method, and Bayesian paradigm method. The results show that: (1) the new remote sensing platform data assimilation shows significant advantages in precision agriculture, especially in emerging satellite constellation remote sensing and UAV data assimilation. (2) SWAP model is the most widely used in simulating crop growth, while Aquacrop, WOFOST, and APSIM models have great potential for application. (3) Sequential assimilation strategy is the most widely used algorithm in the field of agricultural data assimilation, especially the ensemble Kalman filter algorithm, and hierarchical Bayesian assimilation strategy is considered to be a promising method. (4) Leaf area index (LAI) is considered to be the most preferred assimilation variable, and the study of soil moisture (SM) and vegetation index (VIs) has also been strengthened. In addition, the quality, resolution, and applicability of assimilation data sources are the key bottlenecks that affect the application of data assimilation in the development of precision agriculture. In the future, the development of data assimilation models tends to be more refined, diversified, and integrated. To sum up, this study can provide a comprehensive reference for agricultural monitoring, yield prediction, and crop early warning by using the data assimilation model.

1. Introduction

Real-time monitoring and precise prediction of large-scale, long-term crop growth processes constitute a pivotal element in ensuring food security and optimizing agricultural production strategies [1]. While traditional empirical statistical models, regression models, and Markov models have provided some degree of support for agricultural production, their limitations in addressing extreme environmental stressors, evaluating crop variety adaptability, and predicting growth cycles have increasingly come to light [2,3]. In response, researchers are actively exploring the integration of Crop Growth Models (CGMs) with Remote Sensing (RS) technology to develop a monitoring and forecasting platform that closely approximates real-world conditions while offering fine-grained spatial information on crop growth [4].
This fusion strategy, integrating RS data with ground-based biomass predictions, enables cross-scale accurate estimation of critical parameters such as crop yield, pest, and disease thresholds [5,6]. This empowers agricultural decision-makers to swiftly respond to various risks and implement precision management measures, thereby enhancing both economic and ecological benefits of agricultural production [7]. Furthermore, delving deeper into the practical applications of such assimilation strategies in precision agriculture not only contributes to the transformation and upgrading of agricultural paradigms but also holds profound practical significance for refining agricultural production systems and fortifying the defense lines of food security [8].
Since the 1970s, RS technology has achieved remarkable advancements across numerous fields, particularly demonstrating immense potential in agricultural monitoring, crop yield estimation, and precision agriculture applications. This is largely due to its unique advantages such as synchronicity, timeliness, continuous coverage, and extensive observation ranges [9,10]. With further technological development, a series of quantitative inversion products closely related to crop physiological and biochemical parameters have emerged, providing rich and valuable data resources for agricultural early warning, monitoring, and yield assessment [11,12,13,14,15]. For example, in agriculture, horticulture, plant science, and other fields, the use of remote sensing data for plant growth modeling can significantly improve the accuracy and efficiency of the model [16,17,18,19].
Moreover, studies have shown that RS technology also exhibits broad application value in revealing environmental change dynamics, agricultural forecasting, and monitoring [20,21,22,23,24]. However, it still faces significant limitations in deeply uncovering crop growth mechanisms and intricately characterizing the complex interactions between crops and their environment [14]. Externally, this manifests as issues with poor temporal and spatial continuity, limited model fitting accuracy, and insufficient generality [4,25]. Additionally, the substantial costs and time required to process vast amounts of RS data represent a major challenge hindering its widespread application in the agricultural sector.
Compared to RS technology, CGMs operate within a rigorous mathematical framework, enabling them to meticulously track the dynamic physiological and ecological processes of crops [26]. Studies demonstrate that CGMs, through their deep understanding of the intertwined effects of intrinsic crop growth patterns and environmental factors, have become effective tools for simulating crop growth dynamics and multi-dimensional yield prediction [27]. Different from models that simply rely on meteorological statistics, CGMs comprehensively consider atmospheric conditions, soil characteristics, crop genetic traits, and field management practices, which can more accurately simulate the actual growth trajectory of crops and effectively solve the shortcomings of complex system modeling [28]. The process-oriented nature, mechanism analysis, and temporal dynamics of CGMs are crucial for understanding the intricate behaviors of agricultural production systems [25]. To delve further into the latest developments concerning CGMs’ application in data assimilation, we conducted a literature review focusing on key terms like “data assimilation”, “Crop Growth Models”, and “precision agriculture”. The results revealed the frequency distribution of various models in agricultural applications, as depicted in Figure 1. Evident from the figure, the SWAP model occupies a leading position in assimilation applications with an adoption rate exceeding 30%, while EPIC, WOFOST, and APSIM models are also widely used.
On another front, the most recent CGMs research in agricultural Data Assimilation (DA) emphasizes the focus on precise, detailed, and integrated simulation of crop growth processes, aiming to achieve meticulous operations for crop growth status, yield, and even the entire agricultural production system [29]. Table 1 summarizes several common CGMs, showcasing their exceptional applicability in the domain of precision agriculture. To a certain extent, CGMs successfully complement the limitations of RS technology in acquiring and processing crop parameter information, especially in elucidating crop growth mechanisms, simulating crop responses under complex environmental conditions, and improving the accuracy and consistency of data interpretation [12,30]. For instance, multiple studies explicitly highlight that in critical areas such as grain yield forecasting, crop fine-scale management, and regional optimization of agricultural resource allocation, CGMs are gradually emerging as vital instruments for assessing and investigating the performance of agricultural production systems [25,26,31,32].
However, despite their excellent performance in theoretical explorations and small-scale applications, CGMs encounter considerable challenges when scaling up to broader geographical or regional levels. Many CGMs are initially designed with a focus on crop growth at the level of individual plots or fields, and their parameter settings and structures tend to be relatively simplified, which makes it difficult to fully capture the complex uncertainties and interactions of meteorological conditions, soil characteristics, and agricultural management practices at the regional scale [46,47]. Therefore, how to overcome these limitations while maintaining the model prediction accuracy, so that CGMs can be effectively applied in precision agriculture scale, is a key research topic in this field.
Fortunately, DA strategies offer a promising solution to effectively integrate the strengths of RS and CGMs and address the aforementioned challenges. DA methods ingeniously combine multisource RS observational data with CGMs simulation outputs, aiming to harness the full potential of both to achieve a more precise and comprehensive understanding of crop growth processes [30,48]. Compared with the methods using RS technology or crop model alone, the data assimilation method has greater flexibility, computational efficiency, and versatility. RS data have injected rich spatial information into CGMs, filling the spatial dimension gap that models usually lack. On the other hand, CGM can take advantage of the rich details of RS to perform more accurate simulations and predictions [49]. This complementary relationship significantly improves the accuracy of crop yield prediction and agricultural situation monitoring and provides a strong technical basis for the transformation of agricultural production to precision and high efficiency [4,30,50]. Therefore, in-depth study of the application of data assimilation in precision agriculture not only helps to overcome the limitations of existing technologies in agricultural monitoring and yield prediction but is also an important driving force to promote the modernization of agricultural production system.
In view of this, we rely on Web of Science, Scopus, and Google Scholar tools to carry out in-depth retrieval around keywords such as “Data assimilation”, “Crop simulation”, and “assimilation algorithm”. The results are shown in Figure 2. We find that the research situation of DA in the field of agriculture presents the following remarkable characteristics. From the time dimension, the research heat of DA in the field of agriculture shows a continuous upward trend, reaching an historical peak in 2022 (see Figure 2a). This shows that with the progress of science and technology and the growth of the demand for agricultural modernization, the application research of DA technology in agriculture is receiving more and more attention and investment. From the spatial dimension, China, the United States, Canada, and other countries are particularly active in the field of agricultural DA research, with fruitful results (see Figure 2b) that may be related to the advantages of agricultural scientific research strength, technological innovation, and policy support in these countries. However, the current research on the DA method in the field of agriculture is still relatively limited, which may be closely related to the consideration of assimilation cost, the difficulty of technology implementation, the complexity and variability of agricultural production itself, and so on. Although some scholars have combed and summarized the progress of the application of DA in agriculture in detail, and achieved a series of important research results [30,51,52], it is undeniable that there are still significant gaps in the following aspects:
  • The choice of emerging RS platform DA: in the face of an endless stream of new generation RS data (such as satellite constellation, UAV, new RS platform, high space-time resolution, etc.), how to make use of the characteristics of multi-source data, consider application scenarios and cost-effectiveness and other factors, select and effectively integrate these data, and further solve the data source problem of DA is an urgent need for the current research.
  • In-depth understanding of the internal mechanism of DA model: there is a lack of discussion on parameter selection and optimization, uncertainty quantification, multi-model fusion, and fine management of the DA model in existing precision agriculture research, which restricts the further improvement of DA model performance and the expansion of agricultural application.
  • The applicability of new challenges to DA: under the background of extreme climate change, precision agriculture, and the rise of the agriculture 4.0 digital era, challenges such as the processing of new RS sources of DA, multi-dimensional application scenarios, robust models, and new requirements for high-precision DA have not been fully studied and addressed.
Based on the above analysis, this study will focus on the above issues and provide new ideas for the future development of DA in the field of precision agriculture. Through multi-dimensional literature retrieval and strengthening case analysis, this paper summarizes and analyzes the possible new applications of DA in order to promote the wide application of DA in precision agriculture and help to realize the precision, intelligence, and sustainability of agricultural production.
The structure of the rest of this paper is as follows: the Section 2 describes the remote sensing data source, platform, and framework of data assimilation in detail. The Section 3 systematically analyzes the latest development of precision agriculture data assimilation model strategy. The Section 4 discusses the main challenges and future prospects of data assimilation. The Section 5 part is the conclusion. Finally, there is a list of references.

2. Materials and Methods

2.1. Assimilation of Agricultural Data Sources by Remote Sensing

In recent years, DA has made remarkable progress in agricultural practice fields such as crop monitoring, yield prediction, and evaluation [53,54,55,56]. On the one hand, RS data, as an important data source of DA, play an important role in promoting scientific agricultural decision-making; on the other hand, RS data quality, acquisition, management, full chain error, and scale effect all constitute a key bottleneck restricting the effectiveness of DA [48]. This means that selecting and evaluating the universality of RS data is very important to maximize the accuracy of crop yield prediction and agricultural management decision-making. Fortunately, with the rapid development of RS technology, the spatial resolution has leapt from several hundred meters in the past to several meters or even better today, which not only brings unprecedented spatial detail richness and feature recognition refinement to DA but also greatly broadens its depth and breadth in agricultural applications (see Figure 3). In this context, Multispectral data [57,58,59], Hyperspectral data [60,61], Visible light and near-infrared spectra [62,63,64,65,66], as well as various types of RS data such as LiDAR [67,68], are being widely and creatively integrated into the DA process to more comprehensively and accurately characterize the interaction between crop growth status and the environment [69].

2.2. A New Multi-Source Remote Sensing Platform DA

As a new type of RS with the most potential, satellite constellation plays an indispensable role in key fields such as fine water resources management, disaster early warning, and crop yield prediction because of its excellent environmental adaptability, strong comprehensive resolution of multi-parameters, and dual characteristics of wide area coverage and fine monitoring, and provides unprecedented spatial information support for DA work [71]. The satellite constellation consists of a set of carefully planned satellite units that orbit the Earth in a specific orbital layout and configuration, ensuring seamless global coverage and efficient revisit capabilities [72]. It is particularly worth mentioning that the constellation integrates a wide range of sensor devices, including optical cameras (such as panchromatic, multispectral, and hyperspectral cameras), thermal infrared sensors, and microwave radar (such as synthetic Aperture Radar, SAR), etc., which can capture land surface information in an omni-directional and multi-band manner, laying a rich and accurate data foundation for various DA applications [71,73].
Assimilation research in recent years has shown great potential in improving the accuracy and refinement of data analysis. For example, Bouras et al. effectively assimilated Sentinel-2 satellite LAI data into the SAFY model by using ensemble Kalman filtering method, which significantly improved the estimation accuracy of spatial variability of winter wheat yield by more than 70% [74]. Meanwhile, Manivasagam et al. innovatively combined PlanetScope and Sentinel-2 imagery to create a high-density time-series LAI dataset, accurately simulating the dynamic changes in crop yields within fields with notable research outcomes. Additionally, the sub-meter to tens-of-meters high spatial resolution provided by satellite constellations meets the demanding requirements for precise identification of subtle agricultural features [75]. Gilardelli et al. adeptly used remote sensing LAI data to realize fine-grained simulation of rice yields at the sub-field scale [76]. Moreover, Ogweno et al. leveraged the high-resolution datasets from PlanetScope and Sentinel-2, along with advanced machine learning (ML) algorithms such as Random Forest, Support Vector Machines, and Regression Trees, to conduct large-scale Land Use and Land Cover (LULC) mapping with striking classification accuracies, reaching up to 93% and 91%, respectively [77].
Looking ahead, with the successful deployment of Landsat-9 and the anticipated efficient operation of its successor Landsat-Next, coupled with the extensive integration of Dove small satellites from Planet Labs into the DA field, we can reasonably anticipate near-real-time, daily global coverage of RS monitoring, offering timely insights into crop growth conditions [78]. Such advancements will not only elevate the timeliness and accuracy of DA but also substantially bolster the capacity of agriculture and environmental sectors to tackle global challenges, providing robust data support for informed decisions toward sustainable development.
As an important part of modern remote sensing technology, unmanned aerial vehicles (UAVs) have significant advantages compared with remote sensing satellite sensors in terms of mobility, flexibility, safety, cost, ease of operation, response speed, and high-resolution image acquisition ability, and can efficiently capture high-precision details of farmland microtopography, vegetation structure, and water status. UAVs have brought significant innovation and development to the field of data analysis [79,80]. For instance, Yamagishi et al. demonstrated the enormous potential of UAV imaging technology in rapid and non-destructive field phenotyping [81]. Gu et al. utilized a high spatio-temporal resolution UAV-based point spectrometer to significantly improve the accuracy of Sun-Induced Chlorophyll Fluorescence (SIF) measurements at the field scale, opening up a new technological pathway for refined monitoring in agriculture [82]. Concurrently, Xie et al. leveraged UAV multispectral technology to accurately estimate SPAD values across multiple growth stages of lychees, providing a powerful tool for rapid assessment of crop nutritional status [83].
New DA research reveals that the assimilation application value of UAVs is especially pronounced in small-scale, refined agricultural management scenarios. Philippe et al. integrated DA with UAV monitoring and plant parameter estimation techniques to enable real-time monitoring of vegetation health status, providing a scientific basis for precision agricultural management [84]. Additionally, Zhang et al. proposed partial least square method and random forest method to estimate winter wheat grain filling rate (GFR) by using leaf chlorophyll content (LCC) and LAI extracted from UAV images, providing a new perspective for dynamic monitoring of crop growth [85]. Numerous studies have also relied on key parameters related to crop growth status and health obtained via UAVs to provide timely and effective guidance for crop management, propelling DA towards increased intelligence and refinement [86,87,88]. It is worth noting that UAVs possess the capability to rapidly cover vast areas of surface information and can operate in complex terrains and hazardous areas, greatly enhancing the efficiency and accuracy of RS data acquisition [34,89,90]. In summary, the application prospects of UAVs in the precise agriculture DA domain are set to become increasingly broad, and their role in promoting agricultural modernization and boosting agricultural productivity will be even more pronounced.
It is worth noting that the latest technological innovation of the RS platform has injected a strong impetus into the research of DA. The emergence of diversified devices such as ground observation stations, vehicle mobile platforms, and handheld spectrometers greatly expands the breadth and depth of RS data sources, which can effectively make up for the lack of satellite and UAV data and provide near-surface information closer to the actual state of crops, thus significantly improving the accuracy of model assimilation. For example, some researchers have creatively proposed a method of combining multi-sensor ground mapping and representation, which skillfully integrates a variety of heterogeneous but complementary sensor data carried on unmanned rovers to generate a multi-level three-dimensional map of the environment [91]. This technique not only significantly enhances the resolution of mapping texture details, but also opens up a new research approach for DA research. At the same time, Kursch et al. successfully achieved accurate estimation of outdoor vineyard yield by using a consumer-grade RGB-D camera mounted on a mobile robot platform. In addition, they discussed the potential of handheld mid-infrared spectrometers in field fresh soil sampling and real-time analysis, providing strong technical support for soil science research and agricultural precision management [92].
Looking ahead, we can expect greater results from these methods, with the emergence of more high-quality ground observation station data, such as the high-resolution satellite-driven automatic soil observation network SONTE-China, and the multi-spectral sensor system DIEGO deployed on the International Space Station [93]. Zhang et al.’s ground-based RS platform will play an increasingly important role in the field of DA research, starting a new upsurge of scientific research and helping to achieve comprehensive and accurate monitoring and management of agro-ecosystems [94].

2.3. High Spatio-Temporal Resolution Remote Sensing DA

The rise of high temporal resolution satellite constellations has ushered in a stream of high-frequency revisit data, enabling the continuous generation of time series for critical biophysical parameters such as VIs, LAI, and phenology, thereby facilitating seamless and dynamic tracking of crop growth processes [95,96]. This groundbreaking advancement provides an unprecedented level of detailed time series inputs for model assimilation, significantly enhancing the precision modeling capabilities of agricultural ecosystems. In cutting-edge high spatio-temporal resolution DA research, scientists employ Enhanced Spatio-Temporal Adaptive Reflectance Fusion Model to skillfully merge Landsat and MODIS data, yielding composite products with both high spatial and temporal resolutions. These are further combined with models like SAFY to accurately quantify crop yields [97]. Additionally, by leveraging RS, reanalysis, and in-situ datasets, advanced point-to-surface fusion techniques like GRASPS create high-resolution, accurate, and seamless maps of surface SM distribution, significantly improving monitoring accuracy of agricultural water status [98].
Of particular note, high spatial resolution satellites demonstrate unparalleled advantages in revealing within-field heterogeneity, rapidly identifying crop disease patches, and precisely assessing the effects of management practices [99,100,101]. This suite of achievements suggests that continuously exploring and innovating in multi-source RS data fusion methods holds the promise of continually enhancing spatial and temporal resolution, thus pushing the application of DA in the agricultural sector into deeper realms [102,103,104]. Specifically, recent studies have proposed a series of innovative models and frameworks, such as Dual-Branch Convolutional Neural Networks (DB-CNN), the Cascade-PSPNET model integrating Bayesian theory with deep learning (DL) algorithms, and the Spatio-Temporal Fusion (STF) framework, all of which have achieved remarkable results in practical applications [105,106,107]. Fundamentally, it is the continuous innovation and progress of RS data sources that provide unprecedented high spatiotemporal resolution data support for DA research, driving the innovation and application leap of agricultural RS technology in the field of DA.

2.4. The Mechanism and Framework of DA

In recent years, the deep integration of RS technology and CGMs has brought significant innovation power to the DA method in the agricultural field and has shown great potential for application is undoubtedly worthy of recognition [44,108,109,110]. The core mechanism of DA is to use mathematical theory and advanced model algorithms to combine the multi-dimensional and high spatio-temporal resolution observation data from RS platform with the internal biophysical process simulation of CGMs so as to bridge the gap between model simulation and actual observation and enhance the simulation accuracy and predictive behavior of the model to complex agricultural systems [52,111]. To some extent, based on proficiency in RS data analysis, CGMs, and other professional skills, and a deep understanding of the internal mechanism of DA, building an efficient and adaptable assimilation framework is the top priority of DA. A simple assimilation framework is shown in Figure 4.
The DA framework encompasses four core steps: RS observation acquisition and processing, model simulation and parameter setting, assimilation algorithm application, and cyclic iterative updating [112]. RS observation focuses on acquiring high-resolution RS images, involving the integration of multi-source RS data, preprocessing (including radiometric calibration, atmospheric correction, geometric correction, and error elimination), and data fusion techniques (such as super-resolution, multi-scale fusion, and spatio-temporal filtering) to inversely derive vegetation bio-chemical parameters for assimilation input. In the model simulation stage, suitable numerical models and observation operators are selected based on input parameters, generating an integration of bio-chemical parameters, establishing model and parameter errors, and employing diverse assimilation strategies to assess the feasibility of observational data, which, when available, are incorporated into the CGMs. The assimilation algorithm stage employs statistical methodologies (e.g., Kalman filtering, variational methods) or physically-based approaches (like 4D-Var, Ensemble Kalman Filtering) to identify the optimal state estimate via the model’s internal residuals, constructing the framework of the assimilation process system. Lastly, during the iterative updating cycle of assimilation, key state variables are selected and updated by the assimilation algorithm, driving the CGMs to update according to the new state until all updates conclude, yielding the final predictive outcome [111,113].

3. Results

Exploring the DA model with low cost, strong versatility, and strict mechanisms has become one of the research hotspots in the field of precision agriculture assimilation. Therefore, by combing the literature, we summarize the mechanism process and evolution trend of the DA model, as shown in Figure 5. The development track of the DA model gradually develops from simplification to refinement, diversification, and integration. From the initial simple forcing method, which is slightly rough due to the accumulation of model errors due to over-reliance on observations, to the innovation of DA model under Bayesian paradigm, such as Markov chain Monte Carlo (MCMC) algorithm, Variational Bayes, and Bayesian coupling, it undoubtedly marks a great improvement in the theoretical framework and a great leap in domain performance. In addition, the development of the DA model continues to expand in the direction of computational efficiency and parallelization, multi-source data fusion and multi-assimilation strategy coupling, model universality enhancement, and so on.

3.1. Forcing Method

Forced mode assimilation, as a straightforward and simple DA approach, has been extensively adopted in crop growth monitoring [114]. Its fundamental principle involves directly utilizing RS observational data as external driving forces for CGMs, bypassing complex mathematical optimizations or probabilistic statistical treatments. Instead, it relies on the model’s inherent dynamical mechanisms to respond to these external signals, aiming to align model outputs with actual observations [30], as shown in Figure 6. The assimilation model has the advantages of intuition, high efficiency, and no iteration, and plays an important role in the study of early crop DA. Its achievements are mainly focused on the accurate simulation of aboveground biomass and yield [115,116,117,118].
Simultaneously, these studies have vividly exposed the limitations of the Forced mode assimilation; for instance, it is highly sensitive to external driving factors but tends to overlook fluctuations in the model’s internal states. Furthermore, it heavily depends on RS observational data and has limited capacity to deal with uncertainties in the model’s internal states. These constraints are evident in the context of precision agriculture applications [119]. Thus, facing potential nonlinear and non-Gaussian processes in future large-scale and complex application scenarios, it is imperative to explore the integration and innovation of the forced mode assimilation with more advanced DA methods, such as variational assimilation and filtering assimilation. Through such integration, we can effectively enhance the accuracy and robustness of DA, thereby addressing the challenges in the agricultural domain in the future [120].
Figure 6. The assimilation strategy framework of yield prediction is driven directly by assimilation model based on crop canopy temperature as input. Notes: DBA is dry biomass accumulation, kg ha−1; FBA is fresh biomass accumulation, kg ha−1; RDBA is relative DBA; RFBA is relative FBA [121].
Figure 6. The assimilation strategy framework of yield prediction is driven directly by assimilation model based on crop canopy temperature as input. Notes: DBA is dry biomass accumulation, kg ha−1; FBA is fresh biomass accumulation, kg ha−1; RDBA is relative DBA; RFBA is relative FBA [121].
Agronomy 14 01920 g006

3.2. Model Parameter Optimization DA

In contrast, model parameter optimization methods (also known as continuous assimilation methods) are more suitable for complex and refined agricultural scenarios. At the heart of this approach lies the iterative adjustment of model initial parameters within a given reanalysis time window T, where all available observation data and model state values are used as a foundation to strive for the best possible fit between the model trajectory and observed data [109,122,123,124]. Particularly noteworthy among these methods is the Four-Dimensional Variational (4D-Var) algorithm, which builds upon the three-dimensional version by incorporating temporally evolving background information. It transforms the assimilation problem into a search for the optimal solution of the dynamic model, effectively integrating all observational results over a specified time period [125,126,127]. The core mechanism of 4D-Var is shown in Formulas (1)–(3) [128,129]:
J X 0 = 1 2 X 0 X 0 b T B 1 X 0 X 0 b + 1 2 i = 1 n H X i Y i T R i 1 H X i Y i
where J X is the cost function of 4D-Var, X and Y represent the model state variable and observation, respectively, X 0 represents the initial value of the state variable, X 0 b is the background field of the state variable, X 0 b is the prior estimate of X 0 , B and R represent the covariance matrix of the background field and the observation field, respectively, and H is the observation operator. X i is the model prediction value at time i and Y i is the observation vector at time i. X i is obtained from the integral prediction equation M:
X i = M i 1 X i 1
In 4D-Var, the minimum objective function J is solved to obtain the optimal X 0 . First, the gradient information of the objective function is required:
J X 0 = B 1 X 0 X 0 b + i = 0 n M 0 i T H T O 1 [ Y i H i ]
where the adjoint of the second term on the right solves the gradient through the inverse integral of the adjoint equation, and the following assumptions are made in the calculation process:
  • M 0 i T represents the mode prediction operator from 0 to i, and has X i = M 0 i T X 0 , and the forward prediction mode can be expressed as the product of the intermediate prediction mode, that is, X i = M i 1 i M 0 1 X 0 .
  • Under the tangent linearity hypothesis, the pattern operator and the observation operator are linearized, that is, X i X 0 M 0 i ( X i b X 0 b ) H ( X i b X 0 b ) , where M 0 i and H are the tangent operators of the pattern operator and the observation operator, respectively, and the minimum value of J is solved by gradient descent or conjugate gradient method. The obtained X 0 is the optimal solution of the initial state.
The current 4D-Var research either focuses on the selection of assimilation parameter variables [44,62,130] or on the optimization of cost function [108,131]. As shown in Figure 7, the assimilation framework based on the 4D-Var model is shown in detail. Based on this model, many scholars have achieved satisfactory results. However, the latest research shows that there are challenges in the application of precision agriculture [132,133,134,135].
First, based on the statistical normal distribution hypothesis and tangent linear hypothesis, 4D-Var simplifies the complexity of the model, but when dealing with strong nonlinear or significant model errors, linearization approximation may lead to gradient information distortion and increase the uncertainty in the process of assimilation [135]. Second, 4D-Var is extremely sensitive to the selection of initial conditions, and improper initial values may cause the algorithm to fall into local optimization or face difficulties when trying to converge, so it is necessary to carefully evaluate the effectiveness of initial values and seek more robust optimization algorithms. Although some scholars consider using Jacobian matrix to optimize the initial conditions and parameters of the model, it controls the gradient information, determines the search direction and step size, and then considers the model continuity [136]. However, for large-scale, high-resolution models, the calculation of the Jacobian matrix often requires significant computing resources and memory, and the effect is not always as expected. Sparsity and parallel computing technology can alleviate this problem to some extent, but for extremely large-scale problems, the computing cost is still high. At the same time, although the finite difference method is often used in gradient approximation, it may have adaptability problems in high-dimensional cases [137,138]. Therefore, the dimensions and characteristics of the data should be fully considered when selecting and applying the gradient approximation method [139]. As the selection of assimilation parameters also has an impact on the results, the selection of parameterization schemes and physical processes will also bring uncertainty to the data analysis results [140]. Based on this, we extract the following possible solutions:
  • Explore the advantages of optimization algorithms such as conjugate gradient method and quasi-Newton method in solving gradient problems. Global optimization techniques such as multiple startup, simulated annealing, and genetic algorithm were used to improve the initialization and optimization strategy of the algorithm and improve the convergence and stability of the algorithm [113]. Try to calibrate the sensitivity error of the initial conditions based on the model.
  • Search for nonlinear framework: To solve nonlinear uncertainty, explore and develop nonlinear variational assimilation frameworks, such as manifold embedding variational method (ME-Var), four-dimensional variational filtering (4D-VARF), etc. For example, the anisotropic background error and time-window four-dimensional variation algorithm constructed by Wu et al. provides us with valuable reference experience [141].
  • Focusing on uncertainty characterization and quantification. Introducing set members to model uncertainty: set Kalman filtering idea, and developing assimilation methods such as set four-dimensional variational (En-4D-Var) [142]. In terms of physical and chemical parameters, we need to pay attention to the quantification of the uncertainty, reduce the uncertainty through the integration of assimilation methods, and improve the adaptability and scalability of assimilation algorithms.
  • Using ML and AI to improve DA efficiency: Using advanced technologies such as GPU acceleration and distributed parallel computing to develop low-rank or reduced-order models to improve computing efficiency and reduce computing costs [143,144,145,146].

3.3. Model Status Update DA

Ensemble Kalman filter (EnKF) is an assimilation algorithm that combines the advantages of ensemble prediction and Kalman filter. It was first proposed by Evensen and later described in detail by Burgers et al. [147,148]. The model state update assimilation process is shown in Figure 8. Its core operational mechanism involves two main stages: prediction and update [149,150]. Its core mechanism is shown in Formulas (4)–(9) [74,151,152]:
  • Forecast
X i , k + 1 f = M k , k + 1 X i , k μ + ω i , k ,   ω i , k ~ N 0 , Q k
where, X i , k + 1 f represents the predicted value of the i th state variable at time k + 1, X i , k μ represents the estimated value of the ith state variable at time k, X i , k μ represents the state transition matrix from time k to time k + 1, ω i , k represents the model error, which obeys the Gaussian distribution with mean 0 and covariance matrix Q k .
  • Update
X i , k + 1 μ = X i , k + 1 f + K k + 1 Y k + 1 o H k + 1 X i , k + 1 f + v i , k ,   v i , k ~   N 0 , Q k
X ¯ k + 1 μ = 1 N i = 1 N X i , k + 1 μ
K k + 1 = P k + 1 f H T H P k + 1 f H T + R k 1
P k + 1 f = 1 N 1 i = 1 N X i , k + 1 f H X ¯ k + 1 f X i , k + 1 f X ¯ k + 1 f T
H P k + 1 f H T = 1 N 1 i = 1 N H X i , k + 1 f H X ¯ k + 1 f H X i , k + 1 f H X ¯ k + 1 f T
where X i , k + 1 μ is the updated estimate of the ith state variable, K k + 1 represents the Kalman gain, Y k + 1 o is the new observation value, H k + 1 is the observation matrix, and v i , k is the observation error, which follows the Gaussian distribution with the mean of 0 and the covariance matrix of Q k . X ¯ k + 1 μ represents the average estimate of all state variables, P k + 1 f represents the prediction error variance matrix, and R k is the observed noise covariance.
It is found that the EnKF model is most widely used in precision agriculture assimilation, especially in combination with WOFOST, CERES, and SAFY swb models [153]. In addition, the model is often assimilated as physicochemical parameters (such as LAI, SM, NDVI) for yield prediction, and more results have been achieved [154]. Recent DA studies show that EnKF focuses on in-depth exploration of uncertainties associated with multi-source RS assimilation and model blending errors. Some researchers have integrated diverse RS data sources, such as MODIS, Landsat-8, HJ-1 A/B, and Sentinel-2, into the EnKF framework to optimize CGMs simulation performance [155,156]. Other studies have combined MCMC algorithms [157] or hybrid complex evolutionary algorithms with EnKF to recalibrate CGMs parameters and assimilate time-series data like LAI into the models, significantly enhancing simulation accuracy [135]. Li et al. effectively improved the estimation method for winter wheat yields by employing an EnKF strategy to assimilate spatiotemporal LAI into the CERES wheat model, considering factors such as observation error, assimilation stage uncertainty, and temporal-spatial scale variability [5]. Despite these advancements, EnKF still faces several challenges in the field of precision agriculture that require resolution.
In general, EnKF is built upon the assumption of Gaussian distributions for system state and observation errors. However, in practical applications, nonlinear factors such as extreme weather, pests and diseases, light, temperature, moisture, and fertilizer may cause state variables to deviate from Gaussian distributions, exhibiting non-Gaussian characteristics like heavy tails or sharp peaks and leading to inaccurate covariance calculations. Although EnKF can calculate the covariance of prior state vectors using samples from the prior distribution, its effectiveness can be limited when dealing with non-Gaussian properties. Consequently, non-parametric methods like Particle Filtering and Iterative Expectation-Maximization (EM) algorithms, all Gaussian assumptions, emerge as strong alternatives to EnKF [158,159].
Furthermore, EnKF usually estimates the covariance matrix by the differences between the set members. However, when dealing with high-dimensional systems, smaller integration sizes can lead to underestimation or overestimation of covariance, manifested by filter divergence or over-reliance on observed data [160,161]. To address this problem, common strategies include introducing an inflation factor to increase Kalman gain and adopting an adaptive mechanism to dynamically adjust covariance matrix parameters based on historical filter performance and observed data characteristics [162]. In the future, the Schur product (also known as the Hadamard product), which refers to a new matrix formed by the product of the corresponding elements of two matrices in the context of data assimilation, can be used to localize the covariance matrix, that is, to preserve only the covariance between the observation points and the model states that are spatially close. The covariance between distant points is set to zero (or smaller) or the Gaspari-Cohn function (Gaspari and Cohn proposed a smooth truncation function in 1999), which is often used as a localized kernel, that is, the smooth decay from the center (i.e., the observed position) outward to zero, ensuring that the covariance of the neighboring region is fully utilized, and the covariance of the distance between the two points is set to zero. Distant covariances are gradually ignored, and other methods are expected to reduce the number of conditions in the covariance matrix, thereby improving the estimation accuracy [163].
In addition, EnKF faces challenges in dealing with discontinuous processes such as threshold effects, saturation phenomena, or discrete events [164], as the assimilation results are no longer significantly updated or improved despite the input of new observational data after the data reach a certain threshold. There are two main coping strategies: the first is to convert non-Gaussian observations to approximate Gaussian distributions through techniques such as piecewise linearization, logarithmic transformations, Box-Cox transformations, and TPCM pairing [165]. The second option is to take the non-Gaussian process model directly and solve for the Kalman gain using the Jacobian matrix or Newton iterative methods, although this may increase the computational cost [136]
It is noteworthy that recent EnKF research highlights that enhancing its scalability and adaptability is crucial when confronted with complex nonlinearity, non-stationarity, uncertainties, and model errors arising from high-resolution RS, unmanned aerial vehicle hyperspectral data, and extreme climates and changing farming systems [166]. Specific strategies include explicitly incorporating model error terms, implementing weighted fusion of multiple models, adopting advanced smoothing-filtering algorithms like Ensemble Square Root Filter (EnSRF) [167], and exploring ML-based automatic parameter adjustment algorithms and feedback mechanisms [168,169]. These measures hold promise to bolster EnKF’s performance in complex agricultural RS applications, preparing it to tackle future challenges.

3.4. Bayesian Normal Form DA

Particle filtering (PF), as a recursive Bayesian filtering algorithm, has gained significant attention due to its superior capabilities in handling non-Gaussian and nonlinear systems [30,60,170]. PF generates a set of particles through random sampling, representing the probability distribution of the system state, and approximates the posterior probability of state variables by weighting and summing these particles [171]. Its core mechanism is shown in Formulas (10)–(15):
Hypothesis from the posterior probability distribution of the state variables P ( x k | z : k ) to extract the N particles x K i , ω k i i = 1 N , the state of the posterior probability density distribution by the following formula approximation obtained [5]:
P ( x k | z 1 : k ) ω k i i = 1 N δ x k x k i
where, δ represents the Dirac function, k represents the time, x k i represents the particle’s state value, x k i represents the particle’s weight, z k represents the observed value, and P ( x k | z 1 : k ) represents the posterior probability distribution. Generally speaking, the posterior probability density of the state variable is unknown, so it is difficult to obtain the posterior distribution density function by direct sampling. To solve this problem, this study adopts Sequential Importance Sampling (SIS), that is, sampling from an importance distribution function q ( x k | z 1 : k ) whose probability distribution is the same as P ( x k | z 1 : k ) , whose probability distribution is known and easily sampled. The importance weight of the particle is [172]:
ω k i P ( x k | z 1 : k ) q ( x k | z 1 : k )
The method mainly uses recursive ideas to calculate the weight of particle importance [173]:
ω k i = ω k 1 i P ( z k | x k i ) P ( x k i | x k 1 i ) q ( x k i | x k 1 i , z k )
Which q ( x k i | x k 1 i , z k ) is the importance sampling function, P ( z k | x k i ) is the likelihood function, and P ( x k i | x k 1 i ) is the prior probability distribution function. In order to calculate simply, we usually choose to render q ( x k i | x k 1 i , z k ) = P ( x k i | x k 1 i ) , there are [174]:
ω k i = ω k 1 i P ( z k | x k i )
Therefore, the updated particle weights can be as follows [5]:
ω k i = ω k i i = 1 N ω k i
Through this processing, the estimation of the final state variable is the weighted average of all the state values of the particles [172,173,174].
x ^ k = i = 1 N ω k i x k i
In agricultural research, PF has been widely applied in various CGMs (e.g., WOFOST, CERES-Wheat, APSIM, etc.) and the assimilation of key parameters such as LAI, significantly improving the accuracy of simulating crop yield, biomass, SM, and hydraulic parameters [175,176,177,178,179,180,181], as depicted in Figure 9, which shows a framework for assimilation based on PF. Meanwhile, research efforts have focused on the improvement and optimization of PF technology, including residual resampling, residual remapping, sensitivity analysis, PF-MCMC methods, and utilizing rigorous evaluation metrics such as RMSE and normalized RMSE [179,182,183,184].
It is important to note that several assimilation methods mentioned above either rely on analytical equations, cost function minimization, or solving covariance matrices, and often depend on a series of assumptions, such as the tangent linearity assumption, initial values following normal distributions, and Gaussian error structures [186]. When confronted with explicit non-Gaussian and nonlinear processes, adhering to Gaussian assumptions can not only result in suboptimal performance but can also lead to misleading results [187]. Against this backdrop, the importance of MCMC-based stochastic sampling methods becomes evident, offering a more flexible and effective solution. MCMC methods excel at drawing samples from complex posterior distributions, particularly demonstrating their strengths when dealing with problems involving non-normality and multimodality that are difficult to sample directly. By constructing a Markov chain and iteratively updating it according to predefined transition rules (such as the Metropolis–Hastings algorithm or Gibbs sampling), these methods ensure that the stationary distribution of the chain aligns with the target posterior probability distribution [188]. In theory, once the Markov chain reaches its stationary state, the sampling comprehensively covers the support region of the posterior distribution. However, in high-dimensional spaces, the Markov chain might require a long “burn-in period” and numerous iterations to converge, resulting in decreased sampling efficiency, commonly referred to as the “curse of dimensionality” [189]. To verify convergence to the target distribution, diagnostic indicators such as Gelman–Rubin statistic (R-hat) and effective sample size are usually employed [190]. In practice, combining MCMC with other methods or algorithmic improvements might be necessary to enhance efficiency and practicality. As shown in Figure 10, the basic steps of the Metropolis–Hastings MCMC method alongside the PF algorithm illustrate the operational framework of this methodology.
Consider that if the PF is placed in a high-dimensional and long-running scene, most of the particles may have very little weight, and only a few particles have significant weights, resulting in the decrease of the effective coverage of the filter to the state space and the loss of particle diversity, which is the so-called “particle degeneration” problem [192,193]. To deal with this problem, researchers have proposed a series of strategies, such as increasing the number of particles, introducing dynamic particle proliferation or an elimination mechanism, and using adaptive particle number adjustment in order to maintain the diversity of particle systems [194]. However, although these methods can improve the uncertainty estimation, they also increase the computational burden accordingly. To reduce the computational overhead while maintaining the diversity of particles, the researchers also tried to dynamically adjust the resampling threshold, resampling method [195], or migration distribution according to the current particle distribution characteristics (such as entropy and effective number of particles). It is worth noting that the traditional resampling techniques (such as system sampling, importance sampling, etc.) may fail in high-dimensional complex environments [30,124]. It is therefore necessary to employ more complex and computationally expensive strategies such as multivariate kernel density estimation and stratified sampling [196].
On the other hand, the performance of PF depends largely on the choice of initial particle distribution [197]. To avoid particle degradation or falling into local optimization, a priori knowledge, historical data, or model simulation results can be used to intelligently select or generate the initial particle distribution. In addition, PF involves multiple parameter adjustments, and the development of adaptive parameter adjustment algorithms, such as joint probabilistic data association-particle filter, can help to better meet these challenges [171,198]. Although these optimization methods do improve the performance of PF, they also increase the complexity and computational requirements of the algorithm. However, the core advantage of PF in dealing with non-Gaussian processes is always the key to gain a foothold in practical applications [199]. Therefore, in practical applications, we should comprehensively weigh various factors according to the specific situation and select the most appropriate algorithm and optimization strategy.
Hierarchical Bayesian Methods (HBMs), pioneered by Wikle et al. and Berliner et al. [200], are particularly adept at handling complex agricultural datasets with hierarchical or nested structures. Their essence lies in decomposing problems into distinct levels and employing conditional probability distributions to precisely describe the inherent relationships between these levels. This approach transforms the daunting task of solving a complex joint probability distribution into a series of more manageable posterior probability calculations, significantly streamlining the process [201]. HBMs typically consist of three components: a lower-level data model, a mid-level dynamic model, and a top-level parameter model [202]. Its core mechanism is shown in Formulas (16)–(21): In order to ensure the observation error, ϵ is introduced to represent the error term [203].
y = H x + ϵ
In the formula, y represents the observation vector, which contains a series of observations for the real state x, which represents the actual state of the system at a certain time or space (temperature, wind speed), and H is the observation operator (or observation matrix). On the questions of how the observed data are generated from the real state, it can be a case of simple scalar multiplication (for example, a single observation point directly corresponds to a state variable), or it can be a more complex function (for example, RS observation is associated with the real state through the sensor response function). ϵ is the observation error vector, which represents the deviation between the observed data and the real state. In general, we assume that ϵ obeys multivariate normal distribution.
ϵ ~ N 0 , R
Among them, R is the observation error covariance matrix, which reflects the correlation between observation errors and their respective variances. Then we give two expressions of the middle dynamic layer, including as many different application scenarios as possible. The first one represents the dynamic model as a set of partial differential equations [53]. That is:
T t = D 2 T x 2 + 2 T y 2 + f x , y , t
In the formula, T x , y , t is expressed as a temperature field, D is a heat conduction coefficient, and f x , y , t represents an external heat source or boundary condition. This kind of equation is usually solved by finite difference and finite element method discretization. In the second way, we regard dynamic behavior as a random walk process (considering one-dimensional random walk model), and consider the trend and fluctuation of state variables, whose state evolution follows [204]:
d x t = μ d t + σ d W t
where x t is the state variable of t at the time point, μ is the drift rate (long-term trend), σ is the diffusivity (fluctuation amplitude), and W t is the standard Brownian motion (Wiener process), indicating random disturbance. The top-level parameter model is mainly used to describe the parameter uncertainty (the error term of the data model and/or the coefficient of the dynamic model). It usually satisfies a certain prior distribution, based on our preliminary understanding or empirical estimation of them. For example, the coefficient θ of the dynamic model satisfies [203,204]:
θ ~ p θ
Note: p θ is not only Poisson distribution, but may be uniform distribution, normal distribution, or Gamma distribution, which depends on prior knowledge. For the observation error covariance matrix R in the data model, if there is a correlation between its elements, the inverse Wishart distribution may be used as its a priori.
R ~ I W ν , S
where ν represents the degree of freedom parameter and S represents the status matrix.
In practical DA procedures, initial prior distributions of parameters, combined with observational data, are updated through a Bayesian theorem to yield posterior distributions, thereby obtaining optimal parameter estimates or their probability distributions [205]. Thanks to its flexibility, efficiency, strong explanatory power, and robust handling of uncertainties, HBM has found wide application in fields including environmental science, oceanography, meteorology, and others [206,207,208,209,210]. However, its adoption in precision agriculture remains nascent, with numerous challenges still to overcome. For instance, dealing with multi-level uncertainties and computing high-dimensional posterior distributions within HBMs often employs MCMC sampling techniques. Hadley et al., within an HBM framework for assimilating IMTA models, successfully estimated the posterior distributions of state variables and parameters along with their uncertainties using Particle MCMC [211]. Nonetheless, when models have numerous parameters and dense observation data, this can significantly escalate computational costs during assimilation [212,213]. Addressing this, Zhang et al. employed Hamiltonian Monte Carlo (HMC) sampling to effectively mitigate such high computational expenses [214]. Moreover, gradient-based inference methods like variational inference, approximate Bayesian computation, and rapid assimilation algorithms such as PF are also worth exploring and applying [215,216].
In practice, while simplifying assumptions are often made based on linear problems, the presence of nonlinearity, non-stationarity, and multi-scale effects can, in reality, lead to intricate prior distributions and covariance structures in parameter models, complicating model construction and solution [217,218]. Techniques such as mode averaging and Principal Component Analysis (PCA) for dimensionality reduction are commonly used to decrease the number of parameters and alleviate computational loads [219]. With advancements in computational power and deeper understanding of physical processes, emerging methods like Bayesian analysis based on DL and Approaches combining Visual Salience with Hierarchical Bayesian hold promise to further advance HBM assimilation techniques in precision agriculture, ushering in new application opportunities [220,221].

4. Discussion

4.1. Challenges Faced by DA in Agricultural Application

4.1.1. Deep Fusion and Uncertainty Management of Multi-Source Heterogeneous Data

The issue of assimilation accuracy, to some extent, stems from data challenges, and the integration of multi-source heterogeneous data (such as RS, meteorological data, field observations, drone inspections, model simulations, etc.) offers a potential solution to this challenge [222]. However, the discrepancies in RS data currently used in agriculture are among the primary obstacles limiting the deep fusion of such multi-source and heterogeneous datasets. RS data with varying spatial, temporal, and spectral resolutions, as well as scale differences, when used as inputs to drive CGMs, can steer the evolution of model simulation accuracy and reliability in different directions [223,224,225]. Figure 11 vividly illustrates the disparities among different satellite sensors in terms of spectral, spatial, temporal, and radiometric resolution and scale. Generally, for regional-scale applications like regional water-saving irrigation, global climate change monitoring, and earth resource surveys, medium to low-resolution data from sources like MODIS and Sentinel-3 is sufficient [226,227]. Conversely, for local, real-time scale applications such as environmental monitoring and disaster emergency response, the demand for high precision and real-time data is extremely high, necessitating high-resolution RS data [228,229]. It is noteworthy that the acquisition and utilization of RS data in actual agricultural applications are not proactive processes and often overlook the significance of the differences in RS data (resolutions, time steps, observation modes, error characteristics, etc.), thereby increasing the complexity of integrating multiple RS datasets [230,231].
On the one hand, errors in observational data are inevitable, and even when actual crop growth data are obtained, they often suffer from sparsity in both space and time, potentially leading to insufficient fused information and degradation of data precision and quality [232]. For instance, in large-scale SM monitoring, due to resource constraints, only a limited number of monitoring stations can be established, with considerable distances between them, making data points across the entire farmland appear sparse. Additionally, instrument inaccuracies, environmental factors (like wind direction and cloud thickness), and human operational errors all introduce biases [140,162]. On the other hand, during data fusion, there is a lack of consideration for how well CGMs adapt to the data, as there are no visual references, leading to the natural absorption of correlation errors introduced by the diversity in RS data. Models may fail to fully represent crop conditions, especially when ground practices and climate impacts are not reliably understood, focusing more on critical parameter inputs and direct replacements. If certain CGMs data are not suitable for the current study area or climatic conditions, they can introduce substantial errors, propagating uncertainties in multi-source integration, which highlights the importance of incorporating a development agenda aimed at refining imperfect models to align with actual crop development, thereby mitigating uncertainties and errors [233,234]. With the rapid advancement and broad application of precision agriculture assimilation, fully considering and implementing a development agenda to guide precise decision-making, cross-sectoral resource integration, and addressing new scenario challenges is of vital importance.
Moreover, data quality control and uncertainty quantification are crucial considerations in the multi-source integration process. The quality of RS data from diverse sources and structures may vary greatly, and if rigorous selection of RS data quality is not performed, the fused outcome could potentially be less accurate than the original datasets [235,236]. This necessitates emphasis on steps such as data cleaning, deduplication, transformation, and validation to eliminate or reduce noise, outliers, and errors in the data [237]. For example, when studying the impact of climate change on crop yields, sensitivity analysis of parameters and uncertainty analysis should follow modeling to ensure decisions are made within controllable ranges, enhancing the accuracy and reliability of the fused results [238,239].
Nonetheless, the deep integration of multi-source heterogeneous data and uncertainty management in CGMs assimilation also confronts numerous challenges [240,241,242]. Some studies strive to optimize data fusion algorithms, theoretically providing higher quality data for assimilation, but many existing methods overlook complementarity and potential conflicts among datasets, potentially leading to failed fusion in practical agricultural contexts [243,244]. Hence, the development of more sophisticated inter-source relational modeling algorithms, such as those based on ML and DL fusion algorithms (e.g., SWPanGA super-resolution, MSFF, DB-CNN, etc.), to deeply reveal and capture the complex interactions among datasets, holds promise for future DA [82,106,143,144,245,246]. Meanwhile, other studies focus on quantifying and managing uncertainties in both CGMs and the data fusion process, employing methods such as Bayesian Model-Data Fusion (MDF) and Semantic Segmentation Networks [247,248,249]. Furthermore, building an end-to-end uncertainty model based on systems engineering represents another major focal point for future DA research.

4.1.2. Refinement and Dynamic Updating of Complex Crop Growth Models

Current CGMs often face limitations when simulating complex nonlinear physicochemical processes, ultimately due to incomplete understanding of crop growth mechanisms and constraints imposed by data quality and availability. On the one hand, models frequently fail to accurately replicate crop growth processes, which are influenced by various factors such as soil properties, water supply, nutrient uptake, light conditions, and pest infestations [250,251]. On the other hand, studies may lack sufficient and precise input of model parameters, encounter difficulties in measurement, or be hampered by transmission constraints [252]. The input parameters themselves are inherently uncertain and are usually defined within ranges based on prior knowledge and historical calibrations. There is ongoing debate about whether the focus should be on tracing the sources of uncertainty or on qualitatively analyzing the impact of these uncertainties on DA. Regardless of the emphasis, Bayesian frameworks offer effective means for quantifying both the sources of uncertainty and their impacts, providing not only optimal parameter estimates but also a standardized method to evaluate the magnitude and variability of uncertainties [253]. However, the use of advanced sampling techniques like MCMC often requires substantial computational resources and prolonged run times, which remains a topic of discussion [254]. By contrast, the integration of Gibbs sampling with Hierarchical Bayesian DA is seen as an ideal strategy for addressing complexities in model uncertainties [213], leveraging Gibbs sampling’s efficiency in univariate conditional probability distributions and the HBA’s capacity to handle multilevel uncertainties, although considerations for computational resources in agricultural contexts remain crucial.
Model refinement and dynamic updating are enabled by deepening our understanding of physical processes, such as precisely modeling changes in SM, evaporation, infiltration, and crop water uptake in fields, which are vital for realistic representations of crop growth environments. Recent studies highlight the significant benefits of enhanced grid resolution in capturing spatial heterogeneity in farmland, facilitating a more nuanced understanding of crop growth distribution and dynamics [255]. These advancements bring simulations closer to actual crop growth. Furthermore, refined parameter estimation is another advantage, exemplified by improvements to the YOLOv5s algorithm that reduce errors from parameter uncertainties [256]. High-resolution observations and advanced inversion algorithms allow for uncovering local structures and nonlinear relationships in the parameter space, with dynamic updates and adjustments to parameter estimates directly influencing outcomes as new observations are incorporated [257]. Studies have shown that dynamically estimating observation error covariances is more accurate than traditional methods for quantifying observation uncertainties, and, where necessary, assigning weights can help mitigate the impact of observational errors to some extent [258]. These advancements all revolve around quantitative and qualitative analyses of parameter uncertainties to enhance CGMs simulation accuracy; exploring ensemble methods that combine refined models rather than relying on a single optimal model leverages the strengths of multiple models and mitigates biases [259]. Inspired by online learning in ML, frameworks that dynamically adjust model parameters based on new observations are being designed to improve models’ timeliness and adaptability to complex scenarios [260,261].
However, it must be acknowledged that model refinement and dynamic updating can increase the complexity of model structures, leading to more state variables and parameters. This poses challenges for optimizing state estimation and parameter calibration, as these processes can involve solving nonlinear and non-Gaussian problems or managing uncertainty and error quantification [262]. Technical challenges in implementing dynamic updates cannot be overlooked, necessitating models with robust learning capabilities and adaptability to evolving environmental conditions. Innovative algorithms, such as the highly adaptive DL-Crop algorithm, have been proposed to support dynamic updates [146,263]. As we advance CGMs’ sophistication and dynamism, integrating precision agriculture and exploring DL applications in parameter automation and dynamic updating are essential. DL methods like multi-variable time-attention networks and U-nets have proven advantageous in fitting complex nonlinear relations [264]. Bayesian frameworks, including HBMs, HMC, and Bayesian Neural Networks, continue to offer promising avenues for solving nonlinear problems and quantifying uncertainties [264,265]. Hybrid assimilation methods, such as integrated Kalman and Bayesian filters, should also be considered for their advantages in tackling similar challenges [266]. The synergistic application of these methods holds potential for delivering comprehensive and accurate solutions.

4.1.3. From Univariate Assimilation to Multivariable and Multiscale Assimilation

Current DA primarily focuses on the assimilation of individual key variables such as LAI and NDVI, largely due to mismatches in scales between observations and models [267]. Satellite or sensor observations typically provide information at the pixel level, reflecting vegetation status over larger spatial scales and encapsulating complex, mixed ground information. Conversely, models operate based on simplified assumptions and interpret single variables, struggling to precisely capture all relevant biophysical processes when dealing with interconnected micro-scale variables like SM, nutrient content, and pest pressure. These variables are not directly measurable by RS, and model simplifications may not adequately represent their collective influence. Consequently, DA often prioritizes parameters that are easily extracted and have a significant impact on model predictions, overlooking the synergistic effects of multiple interrelated variables during crop growth. While this approach may suffice for broader tasks such as land cover classification, vegetation monitoring, and crop area estimation on a macroscopic scale [268,269], it can prove disastrous for precision operations, leading to intensified drought and flooding, environmental pollution, yield impacts, delayed pest management, and economic losses [270,271,272]. Although simpler to implement, this approach lacks versatility, especially in refined agricultural practices.
A promising solution lies in developing multi-variable, multi-scale integrated assimilation techniques to tackle key issues related to data fusion, model coupling, and uncertainty handling. This necessitates addressing scale transformation problems across different spatial and temporal dimensions, particularly when non-linear transformations and landscape heterogeneity are involved. Adjusting the spatial scale of agricultural landscapes from satellite RS to field scale typically involves averaging satellite (LAI/NDVI) data over fields to represent overall vegetation health [273,274], which may not adequately capture landscape variability [275]. More sophisticated downscaling techniques, such as ML-based methods trained to predict high-resolution parameter values, can be employed. Researchers have linked high-resolution ground measurements with RS (LAI/NDVI) to establish relationships with field-scale crop growth parameters, subsequently applied to low-resolution satellite data to estimate crop conditions at the field scale [276]. Furthermore, CNNs have been used to enhance the resolution of satellite RS images, improving the estimation accuracy of vegetation parameters at the field scale [277].
Adjusting for the time scale requires careful consideration, especially when converting daily data to decadal or monthly averages to match model time steps. Simple averaging or statistical downscaling may not suffice given the inherent nonlinear dynamics of time series data [176,278]. Adopting assimilation strategies that couple space and time and incorporating dynamic models to capture nonlinear temporal dynamics presents a promising avenue. For instance, in predicting pest outbreaks, spatiotemporal assimilation techniques can be designed to integrate forecasting systems seamlessly, enhancing the accuracy and timeliness of disease warnings [279]. Notably, future research may concentrate on dynamic windowing and adaptive filtering techniques [280], as well as ML-based multi-scale assimilation models [281,282], marking a focal point for advancing multi-variable, multi-scale integrated assimilation.

4.2. The Future Development Prospect of DA

4.2.1. Intelligent and Integrated DA

DA applications in agriculture are trending towards greater intelligence and precision, with efficiency and accuracy continually enhanced by the rapid development of multi-sensor integration, cloud computing, Internet of Things (IoT), and artificial intelligence technologies—a consensus among agricultural practitioners [283,284]. In the context of smart greenhouse and field management studies, IoT technology facilitates real-time monitoring and collection of microscale data, including SM, temperature, crop growth status, and soil chemical properties, which are then efficiently consolidated via cloud platforms [285]. The Copernicus Sentinel satellite constellation, offering high-resolution optical and thermal sensing data, combined with high-resolution RS and in-field sensor data revolutionizes Earth observation, enabling agricultural managers to adjust irrigation and fertilization strategies in real-time and with precision through advanced DA algorithms processed in parallel on cloud servers. This results in enhanced accuracy of crop yield forecasts [286,287]. In addition, the assimilation of spatial scale and time dimension and the tradeoff and fusion application of their integration methods are important development directions in the field of data assimilation in the future [288,289,290]. For example, we could explore a more adaptive spatial scale assimilation method, which can deal with the differences of data characteristics and model requirements in different regions and different environmental conditions. In addition, the introduction of time series analysis and sliding window technology can realize the real-time tracking and assimilation of the dynamic process and improve the ability of the model to support short-term prediction and real-time decision-making. Simultaneously, AI technologies have transformative potential, optimizing model parameter estimation and reducing response times through historical data analysis, thereby enhancing agricultural adaptability and resilience under complex climatic and environmental changes [291]. Nevertheless, the integration of smart technologies with precision DA in agriculture must navigate the unique complexities and variability of agricultural systems [292]. For example, data limitations, such as heterogeneity, singularity, scale differences, and uncertainties, often pose significant barriers to progress [74]. Yet technological advancements, like the Planet-Scope, Planet Labs Dove satellite constellations, and UAVs, facilitate the integration of multi-source data, overcoming challenges related to data uniformity and scale [86,87]. High-resolution satellites like Landsat-9 significantly contribute to discerning subtle variations in agricultural landscapes, complementing ground observations [78].
For complex CGMs, whether simulating crop development, processing data, or tracking growth stages dynamically, DA incurs significant costs, especially at larger spatial and temporal resolutions for regional or national scale simulations. The trade-off between model accuracy and complexity is often skewed towards simplicity for predictive power, neglecting modular design. CGMs typically treat simulation units as independent entities, allowing different modules to be assigned to separate units for parallel processing in cloud servers, a strategy facilitated by advances in CUDA, distributed storage, and multi-GPU systems [293]. Considering computational time and cost, Genetic Algorithm-based Task Scheduling (GA-TP), GPU clusters, and computer clusters, alongside AI technologies, are leveraged for their ability to distribute computational loads and expedite processing [294,295].
Large-scale parallel infrastructures like Google Earth Engine (GEE) offer a platform for redesigning DA algorithms with its immense computational power and data storage, presenting new possibilities [296]. Future developments should account for interactions and dependencies among modules, incorporating physical modeling with ML methods like ML-4D-Var, environmental stress indices, and crop variety traits, leveraging ML’s capacity for nonlinear modeling to improve adaptability to varied conditions and crop types [297,298].
While complex CGMs offer accuracy and performance advantages, their computational complexity and demands can hinder DA research. Thus, approximated models or emulators are proposed to simplify computational requirements while preserving model accuracy. Emulators simulate key growth parameters like LAI, biomass, and transpiration rates, combining these with observations and utilizing variational approximation or Gaussian process estimation for model parameterization, leading to more accurate crop growth status predictions and yield estimates [299,300]. Such approaches propel DA towards greater intelligence and service-oriented applications.

4.2.2. DA Applicability of Decision Support for Precision Agriculture

DA, serving as a crucial link between observational data and CGMs, has been extensively demonstrated through cutting-edge research to hold immense potential and application value in supporting precision agriculture decision-making, such as precision irrigation, detailed yield estimation, targeted monitoring, and precision pest control [83,223,301]. Despite these advancements, there remain gaps to bridge before DA fully aligns with the practical requirements of precision agriculture. First and foremost, the practice of precision agriculture imposes stricter demands on the quality and integration of observational data, likely tied to the necessity for high-quality agricultural decisions. Persistent challenges include cloud coverage, sensor errors, resolution disparities, and scale integration. Researchers have thus focused on constructing advanced data fusion frameworks, such as those based on Bayesian theory, designed to simultaneously address complex nonlinear issues arising during the fusion process. These frameworks typically incorporate multi-source RS platforms like LiDAR, SAR, hyperspectral imagery, and high-resolution data from drones to enhance spatial-temporal resolution and data continuity [302,303]. The EU Horizon 2020 project “SEN4CAP” exemplifies this approach. Additionally, for cloud detection and gap filling, ML algorithms, notably CNNs and transfer learning techniques, have emerged as promising solutions, validated by tools like FMask on GEE [304] and the Sen2Cor cloud detection algorithm developed by the European Space Agency (ESA) [305].
Second, the suitability of CGMs in precision agriculture DA necessitates addressing issues related to model optimization and quantification of parameter uncertainties. Approaches range from rule-based expert systems to ML-driven black-box simplifications of models, such as the simplified version of FAO’s AquaCrop model, AquaCrop-OS [223], and the USDA ARS’s Excel-based simplified EPIC model, which greatly facilitate farmer and consultant use [306]. Deeper explorations might involve modular and component-based methodologies, with initiatives like the US DOE’s Agricultural Model Intercomparison and Improvement Project (AgMIP) [307] indicating a promising path forward. The discourse on parameter uncertainty quantification revolves around MCMC and Bayesian parameter estimation under a Bayesian framework, with future developments potentially leaning towards data-driven adaptive CGMs.
Lastly, the choice and adaptability of DA algorithms as core components of precision agriculture decision support systems are critical to ensuring the full exploitation of DA in precision agriculture scenarios. Different DA methods, including forcing methods, 4D-Var, EnKF, PF, and HBM, each have distinct applicability and performance characteristics in precision agriculture. Forcing methods are primarily suited for parameter optimization and observational correction [308], whereas variational and Bayesian assimilation methods are better equipped for large-scale, global, or complex uncertainty handling tasks, such as refined yield prediction [309,310]. HBM particularly excels in integrating multi-source RS data and assessing model uncertainties [119]. Conversely, EnKF and PF show advantages in addressing medium-sized, localized issues, such as precision irrigation decisions, SM forecasting, and decision management [173,311]. Studies also propose hybrid assimilation strategies, e.g., EnKF-Variational assimilation and EnKF-SSPE [312,313], and emphasize the potential of leveraging GPU, cloud computing, and parallel platforms like GEE to enhance assimilation efficiency through large-scale parallel computations, a highly prospective research direction [314].

5. Conclusions

Data assimilation has obvious advantages, great potential, and broad space for development in agricultural applications. However, the quality, resolution, and applicability of assimilation data sources are still the key bottlenecks affecting the development of data assimilation precision agriculture. In the future, the assimilation applications of emerging satellite constellation remote sensing, UAV, ground station data, high spatio-temporal resolution remote sensing, and multi-source remote sensing fusion technology should be vigorously promoted. It is worth noting that in light of the quantification of remote sensing integration errors and uncertainties, the introduction of the development agenda has an efficient and significant effect and is an effective means to promote the future development agenda of precision agriculture. The assimilation application of crop growth model is uneven. The SWAP model stands out in the application of agricultural data assimilation because of its wide applicability. The development trend of Aquacrop, WOFOST, and APSIM models is also quite positive, but the applications of Cropsyst and GLAM models are relatively few. The potential of multi-model fusion and assimilation mechanism is worthy of further study. LAI is the mainstream assimilation parameter selection, and the progress of multi-parameter data assimilation in precision agriculture is worth looking forward to. Model parameter optimization and uncertainty quantification mainly depend on the quantitative calibration method under the Bayesian framework. With the rise of ML technology, automatic parameter adjustment methods and integrated assimilation strategies based on ML, such as EnKF, variational assimilation, Bayesian filtering, and so on, show broad research prospects. In the application trend of precision agriculture assimilation model strategy, the data assimilation research based on EnKF is the most active, while the Bayesian system assimilation strategy shows a trend of rapid development. The development of multi-level assimilation or multi-model integrated assimilation is worth exploring.
The data assimilation model is developing in the direction of refinement, diversification, and integration. Its mechanism framework has been continuously deepened and improved, resulting in a significant improvement in the overall performance. In the future, advances in GPU technology, GEE platform, cloud computing technology, and artificial intelligence will make large-scale parallel assimilation possible and release the great application potential of precision agriculture. However, the current research literature shows that it needs to be strengthened in the field of agricultural data assimilation. In order to meet the growing demand of data assimilation in precision agriculture, future research should focus on the optimization of assimilation data sources, the refinement of crop growth models, and the upgrading of assimilation model strategies to improve adaptability, accuracy, and efficiency. At present, there is an imbalance in spatial and regional distribution in the research of agricultural data assimilation, so it is very important to strengthen the sharing and exchange of international data assimilation results. It is necessary to promote cooperative research between countries and regions, share advanced experience and technological innovations, and promote the innovative development of global agricultural data assimilation.

Author Contributions

Conceptualization, Z.Q. and Y.W.; methodology, Y.W.; software, Y.W.; validation, J.W. and Y.W.; formal analysis, Z.Q.; investigation, Z.Q.; resources, Y.W.; data curation, J.W.; writing—original draft preparation, J.W.; writing—review and editing, J.W.; visualization, Y.W.; supervision, Y.W.; project administration, J.W.; funding acquisition, J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Support Fund for Young Graduate Guidance Teachers of Gansu Agricultural University (GAU-QDFC-2022-18); Gansu Provincial Department of Education Industry Support Plan Project (2022CYZC-41); National Natural Science Foundation of China (32360438); The Central Government guides the Local Science and Technology Development Project (24ZYQA023).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

Leaf area index (LAI), soil moisture (SM), vegetation index (VIs), Crop Growth Models (CGMs), Remote Sensing (RS), Data Assimilation (DA), Land Use and Land Cover (LULC), machine learning (ML), unmanned aerial vehicle (UAV), Sun-Induced Chlorophyll Fluorescence (SIF), grain filling rate (GFR), leaf chlorophyll content (LCC), Dual-Branch Convolutional Neural Networks (DB-CNN), deep learning (DL), Spatio-Temporal Fusion (STF), Markov chain Monte Carlo (MCMC), Four-Dimensional Variational (4D-Var), manifold embedding variational method (ME-Var), Biomass (BM), four-dimensional variational filtering (4D-VARF), Ensemble Kalman Filter (EnKF), Particle Filtering (PF), Iterative Expectation-Maximization (EM), Ensemble Square Root Filter (EnSRF), Hierarchical Bayesian Methods (HBMs), Hamiltonian Monte Carlo (HMC), Principal Component Analysis (PCA), Internet of Things (IoT), Google Earth Engine (GEE), European Space Agency (ESA).

References

  1. Li, Z.; Fan, C.; Zhao, Y.; Jin, X.; Casa, R.; Huang, W.; Song, X.; Blasch, G.; Yang, G.; Taylor, J.; et al. Remote sensing of quality traits in cereal and arable production systems: A review. Crop J. 2024, 12, 45–57. [Google Scholar] [CrossRef]
  2. Lokosang, L.; Ramroop, S.; Hendriks, S.L. Establishing a Robust Technique for Monitoring and Early Warning of Food Insecurity in Post-Conflict South Sudan Using Ordinal Logistic Regression. Agrekon 2011, 50, 101–130. [Google Scholar] [CrossRef]
  3. Vannoppen, A.; Gobin, A.; Kotova, L.; Top, S.; De Cruz, L.; Vīksna, A.; Aniskevich, S.; Bobylev, L.; Buntemeyer, L.; Caluwaerts, S.; et al. Wheat Yield Estimation from NDVI and Regional Climate Models in Latvia. Remote Sens. 2020, 12, 2206. [Google Scholar] [CrossRef]
  4. Wang, S.; Huang, K.; Yan, H.; Yan, H.; Zhou, L.; Wang, H.; Zhang, J.; Yan, J.; Zhao, L.; Wang, Y.; et al. Improving the light use efficiency model for simulating terrestrial vegetation gross primary production by the inclusion of diffuse radiation across ecosystems in China. Ecol. Complex. 2015, 23, 1–13. [Google Scholar] [CrossRef]
  5. He, L.; Zhongxin, C.; Gaohuan, L.; Zhiwei, J.; Chong, H. Improving Winter Wheat Yield Estimation from the CERES-Wheat Model to Assimilate Leaf Area Index with Different Assimilation Methods and Spatio-Temporal Scales. Remote Sens. 2017, 9, 190. [Google Scholar] [CrossRef]
  6. Liu, F.; Li, C.; Dong, Y.; Wang, Q.; Wang, J.; Huang, W. Monitoring crop growth based on assimilation of remote sensing data and crop simulation model. Editor. Off. Trans. Chin. Soc. Agric. Eng. 2011, 27, 101–106. [Google Scholar]
  7. Qu, C.; Wang, Z.; Jin, X.; Wang, X.; Wang, D. A moisture content prediction model for deep bed peanut drying using support vector regression. J. Food Process Eng. 2020, 43, e13510. [Google Scholar] [CrossRef]
  8. Gu, Q.; Han, Y.; Xu, Y.; Yao, H.; Niu, H.; Huang, F. Laboratory Research on Polarized Optical Properties of Saline-Alkaline Soil Based on Semi-Empirical Models and Machine Learning Methods. Remote Sens. 2022, 14, 226. [Google Scholar] [CrossRef]
  9. Mito, C.O.; Boiyo, R.K.; Laneve, G. A simple algorithm to estimate sensible heat flux from remotely sensed MODIS data. Int. J. Remote Sens. 2012, 33, 6109–6121. [Google Scholar] [CrossRef]
  10. Mottaleb, K.A.; Gumma, M.K.; Mishra, A.K.; Mohanty, S. Quantifying Production Losses due to Drought and Submergence of Rainfed Rice at the Household Level Using Remotely Sensed MODIS Data. Agric. Syst. 2014, 137, 227–235. [Google Scholar] [CrossRef]
  11. Tan, C.-W.; Zhang, P.-P.; Zhou, X.-X.; Wang, Z.-X.; Xu, Z.-Q.; Mao, W.; Li, W.-X.; Huo, Z.-Y.; Guo, W.-S.; Yun, F. Quantitative monitoring of leaf area index in wheat of different plant types by integrating NDVI and Beer-Lambert law. Sci. Rep. 2020, 10, 929. [Google Scholar] [CrossRef] [PubMed]
  12. Sun, K.; Su, Q.; Ming, Y. Dust Storm Remote Sensing Monitoring Supported by MODIS Land Surface Reflectance Database. Remote Sens. 2019, 11, 1772. [Google Scholar] [CrossRef]
  13. Parisay, Z.; Sheikh, V.; Bahremand, A.; Komaki, C.B.; Abdollahi, K. An Approach for Estimating Monthly Curve Number Based on Remotely-Sensed MODIS Leaf Area Index Products. Water Resour. Manag. 2019, 33, 2955–2972. [Google Scholar] [CrossRef]
  14. Wang, J.B.; Zhao, X.L.; Ye, H.; Zhang, Z.J.; He, H.L. Integration of evapotranspiration remote sensing products based on Bayesian model averaging: An example from Three-River-Source National Park. J. Appl. Ecol. 2021, 32, 2119–2128. [Google Scholar]
  15. Costa, D.d.A.; Bayissa, Y.; Junior, J.L.; Yamasaki, E.N.; Kyriakides, I.; Neto, A.J.S. Cyprus Surface Water Area Variation Based on the 1984–2021 Time Series Built from Remote Sensing Products. Remote Sens. 2023, 15, 5288. [Google Scholar] [CrossRef]
  16. Wichitwechkarn, V.; Rohde, W.; Fox, C.; Choudhary, R. Data-driven definition and modelling of plant growth. Smart Agric. Technol. 2024, 8, 100495. [Google Scholar] [CrossRef]
  17. Liu, Y.; Wang, B.; Sheng, Q.; Li, J.; Zhao, H.; Wang, S.; Liu, X.; He, H. Dual-polarization SAR rice growth model: A modeling approach for monitoring plant height by combining crop growth patterns with spatiotemporal SAR data. Comput. Electron. Agric. 2023, 215, 108358. [Google Scholar] [CrossRef]
  18. Senko, H.; Kajić, S.; Huđ, A.; Palijan, G.; Petek, M.; Rajnović, I.; Šamec, D.; Udiković-Kolić, N.; Mešić, A.; Brkljačić, L.; et al. Will the beneficial properties of plant-growth promoting bacteria be affected by waterlogging predicted in the wake of climate change: A model study. Appl. Soil Ecol. 2024, 198, 105379. [Google Scholar] [CrossRef]
  19. Kinose, Y.; Fukamachi, Y.; Okabe, S.; Hiroshima, H.; Watanabe, M.; Izuta, T. Toward an impact assessment of ozone on plant carbon fixation using a process-based plant growth model: A case study of Fagus crenata grown under different soil nutrient levels. Sci. Total Environ. 2020, 716, 137008. [Google Scholar] [CrossRef]
  20. Prates-Clark, C.D.; Saatchi, S.S.; Agosti, D. Predicting geographical distribution models of high-value timber trees in the Amazon Basin using remotely sensed data. Ecol. Model. 2008, 211, 309–323. [Google Scholar] [CrossRef]
  21. Buermann, W.; Saatchi, S.; Smith, T.B.; Zutta, B.R.; Chaves, J.A.; Milá, B.; Graham, C.H. Predicting species distributions across the Amazonian and Andean regions using remote sensing data. J. Biogeogr. 2008, 35, 1160–1176. [Google Scholar] [CrossRef]
  22. Al-Bakri, J.; Suleiman, A.; Berg, A. A comparison of two models to predict soil moisture from remote sensing data of RADARSAT II. Arab. J. Geosci. 2014, 7, 4851–4860. [Google Scholar] [CrossRef]
  23. Barnes, B.B.; Hallock, P.; Hu, C.; Muller-Karger, F.; Palandro, D.; Walter, C.; Zepp, R. Prediction of coral bleaching in the Florida Keys using remotely sensed data. Coral Reefs 2015, 34, 491–503. [Google Scholar] [CrossRef]
  24. Wang, J.; Wang, Z.; Lu, B. Assimilation of HJ-LAI into the WOFOST model for estimating regional rice yield. In MIPPR 2019: Remote Sensing Image Processing, Geographic Information Systems, and Other Applications; SPIE: Paris, France, 2020; Volume 11432. [Google Scholar]
  25. Ma, Y.; Wang, S. A review on research of implement application for crop simulation model in regional scale by using remote sensing technology. Chin. J. Appl. Ecol. 2004, 15, 1655–1661. [Google Scholar]
  26. Bai, T.; Wang, S.; Meng, W.; Zhang, N.; Wang, T.; Chen, Y.; Mercatoris, B. Assimilation of Remotely-Sensed LAI into WOFOST Model with the SUBPLEX Algorithm for Improving the Field-Scale Jujube Yield Forecasts. Remote Sens. 2019, 11, 1945. [Google Scholar] [CrossRef]
  27. Mallick, J.; Ahmed, M.; Alqadhi, S.D.; Falqi, I.I.; Parayangat, M.; Singh, C.K.; Rahman, A.; Ijyas, T. Spatial stochastic model for predicting soil organic matter using remote sensing data. Geocarto Int. 2022, 37, 413–444. [Google Scholar] [CrossRef]
  28. Zhang, L.; Guo, C.; Zhao, L.; Zhu, Y.; Cao, W.; Tian, Y.; Cheng, T.; Wang, X. Estimating wheat yield by integrating the WheatGrow and PROSAIL models. Field Crops Res. 2016, 192, 55–66. [Google Scholar] [CrossRef]
  29. Kasampalis, D.A.; Alexandridis, T.K.; Deva, C.; Challinor, A.; Moshou, D.; Zalidis, G. Contribution of Remote Sensing on Crop Models: A Review. J. Imaging 2018, 4, 52. [Google Scholar] [CrossRef]
  30. Jin, X.; Kumar, L.; Li, Z.; Feng, H.; Xu, X.; Yang, G.; Wang, J. A review of data assimilation of remote sensing and crop models. Eur. J. Agron. 2018, 92, 141–152. [Google Scholar] [CrossRef]
  31. Chisanga, C.B.; Phiri, E.; Chinene, V.R.N.; Chabala, L.M. Projecting maize yield under local-scale climate change sce-narios using crop models: Sensitivity to sowing dates, cultivar, and nitrogen fertilizer rates. Food Energy Secur. 2020, 9, e231. [Google Scholar] [CrossRef]
  32. Martínez, M.M.; Nakaegawa, T.; Pinzón, R.; Kusunoki, S.; Gordón, R.; Sanchez-Galan, J.E. Using a Statistical Crop Model to Predict Maize Yield by the End-Of-Century for the Azuero Region in Panama. Atmosphere 2020, 11, 1097. [Google Scholar] [CrossRef]
  33. McCown, R.; Hammer, G.; Hargreaves, J.; Holzworth, D.; Freebairn, D. APSIM: A novel software system for model development, model testing and simulation in agricultural systems research. Agric. Syst. 1996, 50, 255–271. [Google Scholar] [CrossRef]
  34. Rasmussen, J.; Ntakos, G.; Nielsen, J.; Svensgaard, J.; Poulsen, R.N.; Christensen, S. Are vegetation indices derived from consumer-grade cameras mounted on UAVs sufficiently reliable for assessing experimental plots? Eur. J. Agron. 2016, 74, 75–92. [Google Scholar] [CrossRef]
  35. Hsiao, T.C.; Heng, L.; Steduto, P.; Rojas-Lara, B.; Raes, D.; Fereres, E. AquaCrop—The FAO Crop Model to Simulate Yield Response to Water: III. Parameterization and Testing for Maize. Agron. J. 2009, 101, 448. [Google Scholar] [CrossRef]
  36. Jing, Q.; van Keulen, H.; Hengsdijk, H. Modeling biomass, nitrogen and water dynamics in rice–wheat rotations. Agric. Syst. 2010, 103, 433–443. [Google Scholar] [CrossRef]
  37. Batchelor, W.D.; Basso, B.; Paz, J.O. Examples of strategies to analyze spatial and temporal yield variability using crop models. Eur. J. Agron. 2002, 18, 141–158. [Google Scholar] [CrossRef]
  38. Stöckle, C.O.; Donatelli, M.; Nelson, R. CropSyst, a cropping systems simulation model. Eur. J. Agron. 2003, 18, 289–307. [Google Scholar] [CrossRef]
  39. Palosuo, T.; Kersebaum, K.C.; Angulo, C.; Hlavinka, P.; Moriondo, M.; Olesen, J.E.; Patil, R.H.; Ruget, F.; Rumbaur, C.; Takáč, J.; et al. Simulation of winter wheat yield and its variability in different climates of Europe: A comparison of eight crop growth models. Eur. J. Agron. 2011, 35, 103–114. [Google Scholar] [CrossRef]
  40. Jones, J.W.; Hoogenboom, G.; Porter, C.H.; Boote, K.J.; Batchelor, W.D.; Hunt, L.A.; Wilkens, P.W.; Singh, U.; Gijsman, A.J.; Ritchie, J.T. The DSSAT cropping system model. Eur. J. Agron. 2003, 18, 235–265. [Google Scholar] [CrossRef]
  41. Challinor, A.J.; Wheeler, T.R.; Craufurd, P.Q.; Slingo, J.M.; Grimes, D.I. Design and optimisation of a large-area pro-cess-based model for annual crops. Agric. For. Meteorol. 2004, 124, 99–120. [Google Scholar] [CrossRef]
  42. Brisson, N.; Mary, B.; Ripoche, D.; Jeuffroy, M.H.; Ruget, F.; Nicoullaud, B.; Gate, P.; Devienne-Barret, F.; Antonioletti, R.; Durr, C.; et al. STICS: A generic model for the simulation of crops and their water and nitrogen balances. I. Theory and parameterization applied to wheat and corn. Agronomie 1998, 18, 311–346. [Google Scholar] [CrossRef]
  43. Bouman, B.A.M. Linking physical remote sensing models with crop growth simulation models, applied for sugar beet. Int. J. Remote Sens. 1992, 13, 2565–2581. [Google Scholar] [CrossRef]
  44. Huang, J.; Ma, H.; Su, W.; Zhang, X.; Huang, Y.; Fan, J.; Wu, W. Jointly Assimilating MODIS LAI and ET Products Into the SWAP Model for Winter Wheat Yield Estimation. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2015, 8, 4060–4071. [Google Scholar] [CrossRef]
  45. Bansouleh, B.F.; Sargordi, F.; Sharifi, M.A.; Van, H. Point-based simulation of barley production using the crop growth simulation model WOFOST. In Proceedings of the First International Conference on Plant, Water, Soil & Weather Modeling, Kerman, Iran, 14–15 November 2010. [Google Scholar]
  46. Lv, Z.; Liu, X.; Tang, L.; Liu, L.; Cao, W.; Zhu, Y. Estimation of ecotype-specific cultivar parameters in a wheat phenology model and uncertainty analysis. Agric. For. Meteorol. 2016, 221, 219–229. [Google Scholar] [CrossRef]
  47. Jin, N.; Tao, B.; Ren, W.; He, L.; Zhang, D.; Wang, D.; Yu, Q. Assimilating remote sensing data into a crop model improves winter wheat yield estimation based on regional irrigation data. Agric. Water Manag. 2022, 266, 107583. [Google Scholar] [CrossRef]
  48. Huang, G.D. Assimilation of remote sensing into crop growth models: Current status and perspectives. Agric. For. Meteorol. 2019, 276, 107609. [Google Scholar] [CrossRef]
  49. White, J.W.; Hunt, L.; Boote, K.J.; Jones, J.W.; Koo, J.; Kim, S.; Porter, C.H.; Wilkens, P.W.; Hoogenboom, G. Integrated description of agricultural field experiments and production: The ICASA Version 2.0 data standards. Comput. Electron. Agric. 2013, 96, 1–12. [Google Scholar] [CrossRef]
  50. Tian, J.; Qin, J.; Yang, K.; Zhao, L.; Chen, Y.; Lu, H.; Li, X.; Shi, J. Improving surface soil moisture retrievals through a novel assimilation algorithm to estimate both model and observation errors. Remote Sens. Environ. 2022, 269, 112802. [Google Scholar] [CrossRef]
  51. Lewis, P.; Gómez-Dans, J.; Kaminski, T.; Settle, J.; Quaife, T.; Gobron, N.; Styles, J.; Berger, M. An Earth Observation Land Data Assimilation System (EO-LDAS). Remote Sens. Environ. 2012, 120, 219–235. [Google Scholar] [CrossRef]
  52. Luo, L.; Sun, S.; Xue, J.; Gao, Z.; Zhao, J.; Yin, Y.; Gao, F.; Luan, X. Crop yield estimation based on assimilation of crop models and remote sensing data: A systematic evaluation. Agric. Syst. 2023, 210, 103711. [Google Scholar] [CrossRef]
  53. Tang, Y.; Zhou, R.; He, P.; Yu, M.; Zheng, H.; Yao, X.; Cheng, T.; Zhu, Y.; Cao, W.; Tian, Y. Estimating wheat grain yield by assimilating phenology and LAI with the WheatGrow model based on theoretical uncertainty of remotely sensed observation. Agric. For. Meteorol. 2023, 339, 109574. [Google Scholar] [CrossRef]
  54. Dente, L.; Satalino, G.; Mattia, F.; Rinaldi, M. Assimilation of leaf area index derived from ASAR and MERIS data into CERES-Wheat model to map wheat yield. Remote Sens. Environ. 2008, 112, 1395–1407. [Google Scholar] [CrossRef]
  55. Curnel, Y.; de Wit, A.J.W.; Duveiller, G.; Defourny, P. Potential performances of remotely sensed LAI assimilation in WOFOST model based on an OSS Experiment. Agric. For. Meteorol. 2011, 151, 1843–1855. [Google Scholar] [CrossRef]
  56. Zhang, H.; Qin, S.; Ma, J.; You, H. Using Residual Resampling and Sensitivity Analysis to Improve Particle Filter Data Assimilation Accuracy. IEEE Geosci. Remote Sens. Lett. 2013, 10, 1404–1408. [Google Scholar] [CrossRef]
  57. Kumar, A.S.; Reddy, A.M.; Srinivas, L.; Reddy, P.M. Assessment of Surface Water Quality in Hyderabad Lakes by Using Multivariate Statistical Techniques, Hyderabad-India. Environ. Pollut. 2015, 4, 14. [Google Scholar] [CrossRef]
  58. Odermatt, D.; Danne, O.; Philipson, P.; Brockmann, C. Diversity II water quality parameters from ENVISAT (2002–2012): A new global information source for lakes. Earth Syst. Sci. Data 2018, 10, 1527–1549. [Google Scholar] [CrossRef]
  59. Shang, P.; Shen, F. Atmospheric Correction of Satellite GF-1/WFV Imagery and Quantitative Estimation of Suspended Particulate Matter in the Yangtze Estuary. Sensors 2016, 16, 1997. [Google Scholar] [CrossRef]
  60. Huang, G. Improving the Consistency of Nonlinear Estimators: Analysis, Algorithms, and Applications. Ph.D. Thesis, University of Minnesota, Minneapolis, MN, USA, 2013. [Google Scholar]
  61. Wang, S.; Zhuang, Q.; Jin, X.; Yang, Z.; Liu, H. Predicting Soil Organic Carbon and Soil Nitrogen Stocks in Topsoil of Forest Ecosystems in Northeastern China Using Remote Sensing Data. Remote Sens. 2020, 12, 1115. [Google Scholar] [CrossRef]
  62. Li, H.; Chen, Z.; Wu, W.; Jiang, Z.; Liu, B.; Hasi, T. Crop model data assimilation with particle filter for yield prediction using leaf area index of different temporal scales. In Proceedings of the 2015 Fourth International Conference on Agro-Geoinformatics, Istanbul, Turkey, 20–24 July 2015. [Google Scholar]
  63. Araújo, S.R.; Demattê, J.A.M.; Bellinaso, H. Analysing the effects of applying agricultural lime to soils by VNIR spectral sensing: A quantitative and quick method. Int. J. Remote Sens. 2013, 34, 4570–4584. [Google Scholar] [CrossRef]
  64. Angelopoulou, T.; Dimitrakos, A.; Terzopoulou, E.; Zalidis, G.; Theocharis, J.; Stafilov, T.; Zouboulis, A. Reflectance Spectroscopy (Vis-NIR) for Assessing Soil Heavy Metals Concentrations Determined by two Different Analytical Protocols, Based on ISO 11466 and ISO 14869-1. Water Air Soil Pollut. 2017, 228, 436. [Google Scholar] [CrossRef]
  65. Dibi, W.G.; Bosson, J.; Zobi, I.C.; Tié, B.T.; Zoueu, J.T. Use of Fluorescence and Reflectance Spectra for Predicting Okra (Abelmoschus esculentus) Yield and Macronutrient Contents of Leaves. Open J. Appl. Sci. 2017, 7, 537–558. [Google Scholar]
  66. Skakun, S.; Kalecinski, N.I.; Brown, M.G.L.; Johnson, D.M.; Vermote, E.F.; Roger, J.-C.; Franch, B. Assessing within-Field Corn and Soybean Yield Variability from WorldView-3, Planet, Sentinel-2, and Landsat 8 Satellite Imagery. Remote Sens. 2021, 13, 872. [Google Scholar] [CrossRef]
  67. Antonarakis, A.S.; Saatchi, S.S.; Chazdon, R.L.; Moorcroft, P.R. Using Lidar and Radar measurements to constrain pre-dictions of forest ecosystem structure and function. Ecol. Appl. A Publ. Ecol. Soc. Am. 2011, 21, 1120–1137. [Google Scholar]
  68. Peters, S.; Liu, J.; Bruce, D.; Finn, A.; O’hehir, J. Research note: Cost-efficient estimates of Pinus radiata wood volumes using multitemporal LiDAR data. Aust. For. 2021, 84, 206–214. [Google Scholar] [CrossRef]
  69. Chen, Y.; Tao, F. Improving the practicability of remote sensing data-assimilation-based crop yield estimations over a large area using a spatial assimilation algorithm and ensemble assimilation strategies. Agric. For. Meteorol. 2020, 291, 108082. [Google Scholar] [CrossRef]
  70. Kadhim, N.; Mourshed, M.; Bray, M. Advances in remote sensing applications for urban sustainability. Euro-Mediterranean J. Environ. Integr. 2016, 1, 7. [Google Scholar] [CrossRef]
  71. Wijmer, T.; Al Bitar, A.; Arnaud, L.; Fieuzal, R.; Ceschia, E. AgriCarbon-EO v1.0.1: Large-scale and high-resolution simulation of carbon fluxes by assimilation of Sentinel-2 and Landsat-8 reflectances using a Bayesian approach. Geosci. Model Dev. 2024, 17, 997–1021. [Google Scholar] [CrossRef]
  72. Allies, A.; Roumiguié, A.; Dejoux, J.-F.; Fieuzal, R.; Jacquin, A.; Veloso, A.; Champolivier, L.; Baup, F. Evaluation of Multiorbital SAR and Multisensor Optical Data for Empirical Estimation of Rapeseed Biophysical Parameters. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2021, 14, 7268–7283. [Google Scholar] [CrossRef]
  73. Dlamini, L.; Crespo, O.; van Dam, J.; Kooistra, L. A Global Systematic Review of Improving Crop Model Estimations by Assimilating Remote Sensing Data: Implications for Small-Scale Agricultural Systems. Remote Sens. 2023, 15, 4066. [Google Scholar] [CrossRef]
  74. Bouras, E.H.; Olsson, P.-O.; Thapa, S.; Díaz, J.M.; Albertsson, J.; Eklundh, L. Wheat Yield Estimation at High Spatial Resolution through the Assimilation of Sentinel-2 Data into a Crop Growth Model. Remote Sens. 2023, 15, 4425. [Google Scholar] [CrossRef]
  75. Manivasagam, V.; Sadeh, Y.; Kaplan, G.; Bonfil, D.J.; Rozenstein, O. Studying the Feasibility of Assimilating Sentinel-2 and PlanetScope Imagery into the SAFY Crop Model to Predict Within-Field Wheat Yield. Remote Sens. 2021, 13, 2395. [Google Scholar] [CrossRef]
  76. Gilardelli, C.; Stella, T.; Confalonieri, R.; Ranghetti, L.; Campos-Taberner, M.; García-Haro, F.J.; Boschetti, M. Downscaling rice yield simulation at sub-field scale using remotely sensed LAI data. Eur. J. Agron. 2019, 103, 108–116. [Google Scholar] [CrossRef]
  77. Ogweno, V.W.; Moore, N.; Wanyama, D. Assessing cropland disagreement in Tanzania using machine learning methods with Sentinel-2 and Planet Scope imagery. Int. J. Remote. Sens. 2023, 44, 6716–6735. [Google Scholar] [CrossRef]
  78. Wulder, M.A.; Roy, D.P.; Radeloff, V.C.; Loveland, T.R.; Anderson, M.C.; Johnson, D.M.; Healey, S.; Zhu, Z.; Scambos, T.A.; Pahlevan, N.; et al. Fifty years of Landsat science and impacts. Remote Sens. Environ. 2022, 280. [Google Scholar] [CrossRef]
  79. Huang, Y.; Brand, H.J.; Sui, R.; Thomson, S.J.; Furukawa, T.; Ebelhar, M.W. Cotton Yield Estimation Using Very High-Resolution Digital Images Acquired with a Low-Cost Small Unmanned Aerial Vehicle. Trans. ASABE 2016, 59, 1563–1574. [Google Scholar] [CrossRef]
  80. Liao, Z.-Q.; Dai, Y.-L.; Wang, H.; Ketterings, Q.M.; Lu, J.-S.; Zhang, F.-C.; Li, Z.-J.; Fan, J.-L. A double-layer model for improving the estimation of wheat canopy nitrogen content from unmanned aerial vehicle multispectral imagery. J. Integr. Agric. 2023, 22, 2248–2270. [Google Scholar] [CrossRef]
  81. Yamagishi, Y.; Kato, Y.; Ninomiya, S.; Guo, W. Image-Based Phenotyping for Non-Destructive In Situ Rice (Oryza sativa L.) Tiller Counting Using Proximal Sensing. Sensors 2022, 22, 5547. [Google Scholar] [CrossRef] [PubMed]
  82. Guo, S.J.; Wu, W.F.; Shao, Z.F.; Teng, J.H.; Li, D.R. Extracting urban impervious surface based on optical and SAR images cross-modal multi-scale features fusion network. Int. J. Digit. Earth 2024, 17, 19. [Google Scholar] [CrossRef]
  83. Xie, Y.; Wang, P.; Liu, J.; Li, L. Winter wheat yield estimation based on assimilation method combined with 4DVAR and EnKF. Trans. Chin. Soc. Agric. Eng. 2015, 31, 187–195. [Google Scholar] [CrossRef]
  84. Vigneault, P.; Lafond-Lapalme, J.; Deshaies, A.; Khun, K.; de la Sablonnière, S.; Filion, M.; Longchamps, L.; Mimee, B. An integrated data-driven approach to monitor and estimate plant-scale growth using UAV. ISPRS Open J. Photogramm. Remote Sens. 2024, 11, 100052. [Google Scholar] [CrossRef]
  85. Zhang, B.Y.; Gu, L.M.; Dai, M.L.; Bao, X.Y.; Sun, Q.; Zhang, M.Z.; Qu, X.Z.; Li, Z.H.; Zhen, W.C.; Gu, X.H. Estimation of grain filling rate of winter wheat using leaf chlorophyll and LAI extracted from UAV images. Field Crop Res. 2024, 306, 109198. [Google Scholar] [CrossRef]
  86. Wang, L.L.; Huang, X.R.; Li, W.J.; Yan, K.T.; Han, Y.F.; Zhang, Y.L.; Pawlowski, L.; Lan, Y.B. Progress in Agricultural Unmanned Aerial Vehicles (UAVs) Applied in China and Prospects for Poland. Agriculture 2022, 12, 397. [Google Scholar] [CrossRef]
  87. Khan, A.; Malebary, S.J.; Dang, L.M.; Binzagr, F.; Song, H.-K.; Moon, H. AI-Enabled Crop Management Framework for Pest Detection Using Visual Sensor Data. Plants 2024, 13, 653. [Google Scholar] [CrossRef]
  88. Zhang, J.; Pan, B.; Shi, W.; Zhang, Y. Monitoring Waterlogging Damage of Winter Wheat Based on HYDRUS-1D and WOFOST Coupled Model and Assimilated Soil Moisture Data of Remote Sensing. Remote Sens. 2023, 15, 4133. [Google Scholar] [CrossRef]
  89. Jannoura, R.; Brinkmann, K.; Uteau, D.; Bruns, C.; Joergensen, R.G. Monitoring of crop biomass using true colour aerial photographs taken from a remote controlled hexacopter. Biosyst. Eng. 2015, 129, 341–351. [Google Scholar] [CrossRef]
  90. Adão, T.; Hruška, J.; Pádua, L.; Bessa, J.; Peres, E.; Morais, R.; Sousa, J.J. Hyperspectral Imaging: A Review on UAV-Based Sensors, Data Processing and Applications for Agriculture and Forestry. Remote Sens. 2017, 9, 1110. [Google Scholar] [CrossRef]
  91. Milella, A.; Reina, G.; Nielsen, M. A multi-sensor robotic platform for ground mapping and estimation beyond the visible spectrum. Precis. Agric. 2019, 20, 423–444. [Google Scholar] [CrossRef]
  92. Kurtser, P.; Ringdahl, O.; Rotstein, N.; Berenstein, R.; Edan, Y. In-Field Grape Cluster Size Assessment for Vine Yield Estimation Using a Mobile Robot and a Consumer Level RGB-D Camera. IEEE Robot. Autom. Lett. 2020, 5, 2031–2038. [Google Scholar] [CrossRef]
  93. Schultz, J.A.; Hartmann, M.; Heinemann, S.; Janke, J.; Jürgens, C.; Oertel, D.; Rücker, G.; Thonfeld, F.; Rienow, A. DIEGO: A Multispectral Thermal Mission for Earth Observation on the International Space Station. Eur. J. Remote Sens. 2020, 53, 28–38. [Google Scholar] [CrossRef]
  94. Zhang, K.; Zhao, L.; Cui, J.; Mao, P.; Yuan, B.; Liu, Y. Design and Implementation of Evaluation Method for Spraying Coverage Region of Plant Protection UAV. Agronomy 2023, 13, 1631. [Google Scholar] [CrossRef]
  95. Akumaga, U.; Gao, F.; Anderson, M.; Dulaney, W.P.; Houborg, R.; Russ, A.; Hively, W.D. Integration of Remote Sensing and Field Observations in Evaluating DSSAT Model for Estimating Maize and Soybean Growth and Yield in Maryland, USA. Agronomy 2023, 13, 1540. [Google Scholar] [CrossRef]
  96. Boas, T.; Bogena, H.R.; Ryu, D.; Vereecken, H.; Western, A.; Franssen, H.-J.H. Seasonal soil moisture and crop yield prediction with fifth-generation seasonal forecasting system (SEAS5) long-range meteorological forecasts in a land surface modelling approach. Hydrol. Earth Syst. Sci. 2023, 27, 3143–3167. [Google Scholar] [CrossRef]
  97. Sisheber, B.; Marshall, M.; Mengistu, D.; Nelson, A. Assimilation of Earth Observation Data for Crop Yield Estimation in Smallholder Agricultural Systems. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2024, 17, 557–572. [Google Scholar] [CrossRef]
  98. Huang, S.; Zhang, X.; Chen, N.; Ma, H.; Fu, P.; Dong, J.; Gu, X.; Nam, W.; Xu, L.; Rab, G.; et al. A Novel Fusion Method for Generating Surface Soil Moisture Data With High Accuracy, High Spatial Resolution, and High Spatio-Temporal Continuity. Water Resour. Res. 2022, 58, e2021WR030827. [Google Scholar] [CrossRef]
  99. Rostami, A.; Raeini-Sarjaz, M.; Chabokpour, J.; Azamathulla, H.M.; Kumar, S. Determination of rainfed wheat agri-culture potential through assimilation of remote sensing data with SWAT model case study: ZarrinehRoud Basin, Iran. Water Sci. Technol. Water Supply 2022, 22, 5331–5354. [Google Scholar] [CrossRef]
  100. Nadeem, A.A.; Zha, Y.; Shi, L.; Ali, S.; Wang, X.; Zafar, Z.; Afzal, Z.; Tariq, M.A.U.R. Spatial Downscaling and Gap-Filling of SMAP Soil Moisture to High Resolution Using MODIS Surface Variables and Machine Learning Approaches over ShanDian River Basin, China. Remote Sens. 2023, 15, 812. [Google Scholar] [CrossRef]
  101. Ma, H.; Sun, S.; Wang, Z.; Jiang, Y.; Liu, S. Downscaling CLDAS Soil Moisture Product by Integrating Sentinel-1 and Sentinel-2 Data over Agricultural Area. Can. J. Remote Sens. 2022, 48, 737–746. [Google Scholar] [CrossRef]
  102. Tomer, S.K.; Al Bitar, A.; Sekhar, M.; Zribi, M.; Bandyopadhyay, S.; Kerr, Y. MAPSM: A Spatio-Temporal Algorithm for Merging Soil Moisture from Active and Passive Microwave Remote Sensing. Remote Sens. 2016, 8, 990. [Google Scholar] [CrossRef]
  103. Chen, Q.; Jia, L.; Menenti, M.; Hu, G.; Wang, K.; Yi, Z.; Zhou, J.; Peng, F.; Ma, S.; You, Q.; et al. A data-driven high spatial resolution model of biomass accumulation and crop yield: Application to a fragmented desert-oasis agroecosystem. Ecol. Model. 2023, 475, 110182. [Google Scholar] [CrossRef]
  104. Yang, Z.; He, Q.; Miao, S.; Wei, F.; Yu, M. Surface Soil Moisture Retrieval of China Using Multi-Source Data and Ensemble Learning. Remote Sens. 2023, 15, 2786. [Google Scholar] [CrossRef]
  105. Ye, F.; Zhou, Z.; Wu, Y.; Enkhtur, B. Application of convolutional neural network in fusion and classification of multi-source remote sensing data. Front. Neurorobot. 2022, 16, 1095717. [Google Scholar] [CrossRef] [PubMed]
  106. Qi, C.; Liu, W. Tomato Production Prediction Based on Deep Learning Algorithm-Cascade-PSPNET and Bayes. Int. J. Pattern Recognit. Artif. Intell. 2023, 37, 23. [Google Scholar] [CrossRef]
  107. Li, Y.; Gao, W.; Jia, J.; Tao, S.; Ren, Y. Developing and evaluating the feasibility of a new spatiotemporal fusion framework to improve remote sensing reflectance and dynamic LAI monitoring. Comput. Electron. Agric. 2022, 198, 107037. [Google Scholar] [CrossRef]
  108. Zhuo, W.; Huang, J.; Li, L.; Zhang, X.; Ma, H.; Gao, X.; Huang, H.; Xu, B.; Xiao, X. Assimilating Soil Moisture Retrieved from Sentinel-1 and Sentinel-2 Data into WOFOST Model to Improve Winter Wheat Yield Estimation. Remote Sens. 2019, 11, 1618. [Google Scholar] [CrossRef]
  109. Mathieu, P.-P.; Oneill, A. Data assimilation: From photon counts to Earth System forecasts. Remote Sens. Environ. 2008, 112, 1258–1267. [Google Scholar] [CrossRef]
  110. de Wit, A.; Duveiller, G.; Defourny, P. Estimating regional winter wheat yield with WOFOST through the assimilation of green area index retrieved from MODIS observations. Agric. For. Meteorol. 2012, 164, 39–52. [Google Scholar] [CrossRef]
  111. Zhuo, W.; Huang, H.; Gao, X.; Li, X.; Huang, J. An Improved Approach of Winter Wheat Yield Estimation by Jointly Assimilating Remotely Sensed Leaf Area Index and Soil Moisture into the WOFOST Model. Remote Sens. 2023, 15, 1825. [Google Scholar] [CrossRef]
  112. Wu, Y.; Xu, W.; Huang, H.; Huang, J. Bayesian Posterior-Based Winter Wheat Yield Estimation at the Field Scale through Assimilation of Sentinel-2 Data into WOFOST Model. Remote Sens. 2022, 14, 3727. [Google Scholar] [CrossRef]
  113. Huang, J.; Tian, L.; Liang, S.; Ma, H.; Becker-Reshef, I.; Huang, Y.; Su, W.; Zhang, X.; Zhu, D.; Wu, W. Improving winter wheat yield estimation by assimilation of the leaf area index from Landsat TM and MODIS data into the WOFOST model. Agric. For. Meteorol. 2015, 204, 106–121. [Google Scholar] [CrossRef]
  114. Yang, B.; Zhu, W.; Rezaei, E.E.; Li, J.; Sun, Z.; Zhang, J. The Optimal Phenological Phase of Maize for Yield Prediction with High-Frequency UAV Remote Sensing. Remote Sens. 2022, 14, 1559. [Google Scholar] [CrossRef]
  115. Carbone, G.J.; Narumalani, S.; King, M. Application of remote sensing and GIS technologies with physiological crop models. Photogramm. Eng. Remote Sens. 1996, 62, 171–179. [Google Scholar]
  116. Dorigo, W.A.; Zurita-Milla, R.; De Wit, A.J.W.; Brazile, J.; Singh, R.; Schaepman, M.E. A review on reflective remote sensing and data assimilation techniques for enhanced agroecosystem modeling. Int. J. Appl. Earth Observ. Geoinform. 2007, 9, 165–193. [Google Scholar] [CrossRef]
  117. Maas, S.J. Use of remotely-sensed information in agricultural crop growth models. Ecol. Model. 1988, 41, 247–268. [Google Scholar] [CrossRef]
  118. Delécolle, R.; Maas, S.; Guérif, M.; Baret, F. Remote sensing and crop production models: Present trends. ISPRS J. Photogramm. Remote Sens. 1992, 47, 145–161. [Google Scholar] [CrossRef]
  119. Lee, T.-R.; Chang, Y.-S.; Choi, J.-B.; Kim, Y.-J. Numerical simulation of cylinder oscillation by using a direct forcing technique. Nucl. Eng. Des. 2010, 240, 1941–1948. [Google Scholar] [CrossRef]
  120. Mohamedou, C.; Kangas, A.; Hamedianfar, A.; Vauhkonen, J. Potential of Bayesian formalism for the fusion and assimilation of sequential forestry data in time and space. Can. J. For. Res. 2022, 52, 439–449. [Google Scholar] [CrossRef]
  121. Chang, H.; Cai, J.; Zhang, B.; Wei, Z.; Xu, D. Early Yield Forecasting of Maize by Combining Remote Sensing Images and Field Data with Logistic Models. Remote Sens. 2023, 15, 1025. [Google Scholar] [CrossRef]
  122. Talagrand, O.; Courtier, P. Variational Assimilation of Meteorological Observations With the Adjoint Vorticity Equation. I: Theory. Q. J. R. Meteorol. Soc. 1987, 113, 1311–1328. [Google Scholar] [CrossRef]
  123. Li, H.-W.; Wang, J.; Su, H.-T. Improved particle filter based on differential evolution. Electron. Lett. 2011, 47, 1078–1079. [Google Scholar] [CrossRef]
  124. Dhanya, M.; Chandrasekar, A. Impact of variational assimilation using multivariate background error covariances on the simulation of monsoon depressions over India. Ann. Geophys. 2016, 34, 187–201. [Google Scholar] [CrossRef]
  125. Le Dimet, F.X.; Talagrand, O. Variational algorithms for analysis and assimilation of meteorological observations: Theoretical aspects. Tellus A Dyn. Meteorol. Oceanogr. 2016, 45, 449–461. [Google Scholar] [CrossRef]
  126. Chouinard, C.; Halle, J. Use of moisture sensitive satellite radiances in the Canadian Meteorological Centre Unified 3D-var system. Proc. SPIE Int. Soc. Opt. Eng. 2003, 4895, 63–74. [Google Scholar] [CrossRef]
  127. Li, J.; Chen, X.; Tian, L.; Huang, J.; Feng, L. Improved capabilities of the Chinese high-resolution remote sensing satellite GF-1 for monitoring suspended particulate matter (SPM) in inland waters: Radiometric and spatial considerations. ISPRS J. Photogramm. Remote Sens. 2015, 106, 145–156. [Google Scholar] [CrossRef]
  128. Rao, V.; Sandu, A. A time-parallel approach to strong-constraint four-dimensional variational data assimilation. J. Comput. Phys. 2016, 313, 583–593. [Google Scholar] [CrossRef]
  129. Jiang, Z.; Chen, Z.; Chen, J.; Ren, J.; Li, Z.; Sun, L. The Estimation of Regional Crop Yield Using Ensemble-Based Four-Dimensional Variational Data Assimilation. Remote Sens. 2014, 6, 2664–2681. [Google Scholar] [CrossRef]
  130. Wagner, M.P.; Slawig, T.; Taravat, A.; Oppelt, N. Remote Sensing Data Assimilation in Dynamic Crop Models Using Particle Swarm Optimization. ISPRS Int. J. Geo-Inf. 2020, 9, 105. [Google Scholar] [CrossRef]
  131. Huang, X.-Y.; Xiao, Q.; Barker, D.M.; Zhang, X.; Michalakes, J.; Huang, W.; Henderson, T.; Bray, J.; Chen, Y.; Ma, Z.; et al. Four-Dimensional Variational Data Assimilation for WRF: For-mulation and Preliminary Results. Mon. Weather Rev. 2009, 137, 299–302,305,307–314. [Google Scholar] [CrossRef]
  132. Alavi, M.; Albaji, M.; Golabi, M.; Naseri, A.A.; Homayouni, S. Estimation of sugarcane evapotranspiration from remote sensing and limited meteorological variables using machine learning models. J. Hydrol. 2024, 629, 130605. [Google Scholar] [CrossRef]
  133. Gauthier, P.; Tanguay, M.; Laroche, S.; Pellerin, S.; Morneau, J. Extension of 3DVAR to 4DVAR: Implementation of 4DVAR at the Meteorological Service of Canada. Mon. Weather Rev. 2007, 135, 2339–2354. [Google Scholar] [CrossRef]
  134. Jin, Z.; Tian, X.; Duan, M.; Han, R. An Efficient Algorithm for Retrieving CO2 in the Atmosphere From Hyperspectral Measurements of Satellites: Application of NLS-4DVar Data Assimilation Method. Front. Earth Sci. 2021, 9, 688542. [Google Scholar] [CrossRef]
  135. Zhang, Y.; Walker, J.P.; Pauwels, V.R.N.; Sadeh, Y. Assimilation of Wheat and Soil States into the APSIM-Wheat Crop Model: A Case Study. Remote Sens. 2022, 14, 65. [Google Scholar] [CrossRef]
  136. Römer, U.; Narayanamurthi, M.; Sandu, A. Solving parameter estimation problems with discrete adjoint exponential integrators. Optim. Methods Softw. 2018, 33, 750–770. [Google Scholar] [CrossRef]
  137. Qin, J.; Liang, S.; Liu, R.; Zhang, H.; Hu, B. A Weak-Constraint-Based Data Assimilation Scheme for Estimating Surface Turbulent Fluxes. IEEE Geosci. Remote Sens. Lett. 2007, 4, 649–653. [Google Scholar] [CrossRef]
  138. Dong, Y.; Wang, J.; Li, C.; Yang, G.; Wang, Q.; Liu, F.; Zhao, J.; Wang, H.; Huang, W. Comparison and Analysis of Data Assimilation Algorithms for Predicting the Leaf Area Index of Crop Canopies. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2013, 6, 188–201. [Google Scholar] [CrossRef]
  139. Jin, H.; Li, A.; Wang, J.; Bo, Y. Improvement of spatially and temporally continuous crop leaf area index by integration of CERES-Maize model and MODIS data. Eur. J. Agron. 2016, 78, 1–12. [Google Scholar] [CrossRef]
  140. Fang, H.; Liang, S.; Hoogenboom, G.; Teasdale, J.; Cavigelli, M. Corn-field estimation through assimilation of remotely sensed data into the CSM-CERES-Maize model. Int. J. Remote Sens. 2008, 29, 3011–3032. [Google Scholar] [CrossRef]
  141. Wu, S.; Yang, P.; Chen, Z.; Ren, J.; Li, H.; Sun, L. Estimating winter wheat yield by assimilation of remote sensing data with a four-dimensional variation algorithm considering anisotropic background error and time window. Agric. For. Meteorol. 2021, 301–302, 108345. [Google Scholar] [CrossRef]
  142. Minami, T.; Nakano, S.; Lesur, V.; Takahashi, F.; Matsushima, M.; Shimizu, H.; Nakashima, R.; Taniguchi, H.; Toh, H. A candidate secular variation model for IGRF-13 based on MHD dynamo simulation and 4DEnVar data assimilation. Earth Planets Space 2020, 72, 136. [Google Scholar] [CrossRef]
  143. Liu, W.; Mõttus, M.; Gastellu-Etchegorry, J.-P.; Fang, H.; Atherton, J. Seasonal and vertical variation in canopy structure and leaf spectral properties determine the canopy reflectance of a rice field. Agric. For. Meteorol. 2024, 355, 110132. [Google Scholar] [CrossRef]
  144. Huang, G.; Guo, Y.; Tan, W.; Zhang, M.; Li, Z.; Zhou, Y.; Duan, L. Enhancing maize radiation use efficiency under high planting density by shaping canopy architecture with a plant growth regulator. Crop Environ. 2024, 3, 51–63. [Google Scholar] [CrossRef]
  145. Shu, M.; Li, Q.; Ghafoor, A.; Zhu, J.; Li, B.; Ma, Y. Using the plant height and canopy coverage to estimation maize aboveground biomass with UAV digital images. Eur. J. Agron. 2023, 151, 126957. [Google Scholar] [CrossRef]
  146. Blessing, S.; Giering, R.; van der Tol, C. OptiSAIL: A system for the simultaneous retrieval of soil, leaf, and canopy parameters and its application to Sentinel-3 Synergy (OLCI + SLSTR) top-of-canopy reflectances. Sci. Remote Sens. 2024, 10, 100148. [Google Scholar] [CrossRef]
  147. Evensen, G. Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics. J. Geophys. Res. Oceans 1994, 99, 10143–10162. [Google Scholar] [CrossRef]
  148. Burgers, G.J.H.; Leeuwen, P.J.V.; Evensen, G. Analysis Scheme in the Ensemble Kalman Filter. Mon. Weather Rev. 1998, 126, 1719–1724. [Google Scholar] [CrossRef]
  149. Evensen, G. The Ensemble Kalman Filter: Theoretical formulation and practical implementation. Ocean Dyn. 2003, 53, 343–367. [Google Scholar] [CrossRef]
  150. Gao, J.; Xue, M.; Stensrud, D.J. The Development of a Hybrid EnKF-3DVAR Algorithm for Storm-Scale Data Assimilation. Adv. Meteorol. 2013, 2013, 179–182. [Google Scholar] [CrossRef]
  151. Montaldo, N.; Gaspa, A.; Corona, R. Multiscale Assimilation of Sentinel and Landsat Data for Soil Moisture and Leaf Area Index Predictions Using an Ensemble-Kalman-Filter-Based Assimilation Approach in a Heterogeneous Ecosystem. Remote Sens. 2022, 14, 3458. [Google Scholar] [CrossRef]
  152. Ren, S.; Chen, H.; Hou, J.; Zhao, P.; Dong, Q.; Feng, H. Based on historical weather data to predict summer field-scale maize yield: Assimilation of remote sensing data to WOFOST model by ensemble Kalman filter algorithm. Comput. Electron. Agric. 2024, 219, 108822. [Google Scholar] [CrossRef]
  153. Pauwels, V.R.N.; Verhoest, N.E.C.; Lannoy, G.J.M.D.; Guissard, V.; Lucau, C.; Defourny, P. Optimization of a coupled hydrology-crop growth model through the assimilation of observed soil moisture and leaf area index values using an ensemble Kalman filter. Water Resour. Res. 2007, 43, 4417–4421. [Google Scholar] [CrossRef]
  154. Kalnay, E.; Li, H.; Miyoshi, T.; Yang, S.-C.; Ballabrera-Poy, J. 4-D-Var or ensemble Kalman filter? Tellus A 2007, 59, 758–773. [Google Scholar] [CrossRef]
  155. Dhakar, R.; Sehgal, V.K.; Chakraborty, D.; Sahoo, R.N.; Mukherjee, J.; Ines, A.V.; Kumar, S.N.; Shirsath, P.B.; Roy, S.B. Field scale spatial wheat yield forecasting system under limited field data availability by integrating crop simulation model with weather forecast and satellite remote sensing. Agric. Syst. 2022, 195, 103299. [Google Scholar] [CrossRef]
  156. Silvestro, P.C.; Casa, R.; Hanuš, J.; Koetz, B.; Rascher, U.; Schuettemeyer, D.; Siegmann, B.; Skokovic, D.; Sobrino, J.; Tudoroiu, M. Synergistic Use of Multispectral Data and Crop Growth Modelling for Spatial and Temporal Evapotranspiration Estimations. Remote Sens. 2021, 13, 2138. [Google Scholar] [CrossRef]
  157. Kang, Y.; Özdoğan, M. Field-level crop yield mapping with Landsat using a hierarchical data assimilation approach. Remote Sens. Environ. 2019, 228, 144–163. [Google Scholar] [CrossRef]
  158. Ines, A.V.; Das, N.N.; Hansen, J.W.; Njoku, E.G. Assimilation of remotely sensed soil moisture and vegetation with a crop simulation model for maize yield prediction. Remote Sens. Environ. 2013, 138, 149–164. [Google Scholar] [CrossRef]
  159. Dreano, D.; Tandeo, P.; Pulido, M.; Ait-El-Fquih, B.; Chonavel, T.; Hoteit, I. Estimating model-error covariances in nonlinear state-space models using Kalman smoothing and the expectation–maximization algorithm. Q. J. R. Meteorol. Soc. 2017, 143, 1877–1885. [Google Scholar] [CrossRef]
  160. Jun, M.; Szunyogh, I.; Genton, M.G.; Zhang, F.Q.; Bishop, C.H. A Statistical Investigation of the Sensitivity of Ensemble-Based Kalman Filters to Covariance Filtering. Mon. Weather Rev. 2011, 139, 3036–3051. [Google Scholar] [CrossRef]
  161. Zhao, Y.; Chen, S.; Shen, S. Assimilating remote sensing information with crop model using Ensemble Kalman Filter for improving LAI monitoring and yield estimation. Ecol. Model. 2013, 270, 30–42. [Google Scholar] [CrossRef]
  162. Casa, R.; Varella, H.; Buis, S.; Guerif, M.; Solan, B.D.; Baret, F. Forcing a wheat crop model with LAI data to access ag-ronomic variables: Evaluation of the impact of model and LAI uncertainties and comparison with an empirical approach. Eur. J. Agron. 2012, 37, 1–10. [Google Scholar] [CrossRef]
  163. Gilpin, S.; Matsuo, T.; Cohn, S.E. A generalized, compactly supported correlation function for data assimilation applications. Q. J. R. Meteorol. Soc. 2023, 149, 1953–1989. [Google Scholar] [CrossRef]
  164. Chakrabarti, S.; Bongiovanni, T.; Judge, J.; Zotarelli, L.; Bayer, C. Assimilation of SMOS Soil Moisture for Quantifying Drought Impacts on Crop Yield in Agricultural Regions. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2017, 7, 3867–3879. [Google Scholar] [CrossRef]
  165. Liao, Q. Uncertainty Quantification and Data Assimilation via Transform Process for Strongly Nonlinear Problems. Ph.D. Thesis, University of Southern California, Los Angeles, CA, USA, 2014. [Google Scholar]
  166. Aftab, S.; Moghadam, R.H. Robust data smoothing algorithms and wavelet filter for denoising sonic log signals. J. Appl. Geophys. 2022, 206, 104836. [Google Scholar] [CrossRef]
  167. Yu, X.K.; Li, J.X. Optimal Filtering and a Smoothing Algorithm for a Singular System With a Complex Stochastic Uncertain Parameter Matrix. IEEE Trans. Circuits Syst. II Express Briefs 2020, 67, 780–784. [Google Scholar] [CrossRef]
  168. Xiao, D.; Du, J.; Fang, F.; Pain, C.C.; Li, J. Parameterised non-intrusive reduced order methods for ensemble Kalman filter data assimilation. Comput. Fluids 2018, 177, 69–77. [Google Scholar] [CrossRef]
  169. Lin, X.; Yang, X.H.; Hu, C.H.; Li, W. Improved forward and backward adaptive smoothing algorithm. GPS Solut. 2022, 26, 13. [Google Scholar] [CrossRef]
  170. Tian, M.C.; Bo, Y.M.; Chen, Z.M.; Wu, P.L.; Yue, C. Multi-target tracking method based on improved firefly algorithm optimized particle filter. Neurocomputing 2019, 359, 438–448. [Google Scholar] [CrossRef]
  171. Arulampalam, M.; Maskell, S.; Gordon, N.; Clapp, T. A tutorial on particle filters for online nonlinear/non-Gaussian Bayesian tracking. IEEE Trans. Signal Process. 2002, 50, 174–188. [Google Scholar] [CrossRef]
  172. Ouaadi, N.; Jarlan, L.; Khabba, S.; Ezzahar, J.; Le Page, M.; Merlin, O. Irrigation Amounts and Timing Retrieval through Data Assimilation of Surface Soil Moisture into the FAO-56 Approach in the South Mediterranean Region. Remote Sens. 2021, 13, 2667. [Google Scholar] [CrossRef]
  173. Ju, F.; An, R.; Yang, Z.; Huang, L.; Sun, Y. Assimilating SMOS Brightness Temperature for Hydrologic Model Parameters and Soil Moisture Estimation with an Immune Evolutionary Strategy. Remote Sens. 2020, 12, 1556. [Google Scholar] [CrossRef]
  174. Lei, F.; Crow, W.T.; Kustas, W.P.; Dong, J.; Yang, Y.; Knipper, K.R.; Anderson, M.C.; Gao, F.; Notarnicola, C.; Greifeneder, F.; et al. Data assimilation of high-resolution thermal and radar remote sensing retrievals for soil moisture monitoring in a drip-irrigated vineyard. Remote Sens. Environ. 2020, 239, 111622. [Google Scholar] [CrossRef]
  175. Mao, F.; Li, X.; Du, H.; Zhou, G.; Han, N.; Xu, X.; Liu, Y.; Chen, L.; Cui, L. Comparison of Two Data Assimilation Methods for Improving MODIS LAI Time Series for Bamboo Forests. Remote Sens. 2017, 9, 401. [Google Scholar] [CrossRef]
  176. Liu, J.; Huang, J.; Tian, L.; Su, W. Particle Filter-Based Assimilation Algorithm for Improving Regional Winter Wheat Yield Estimation. Sens. Lett. 2014, 12, 763–769. [Google Scholar] [CrossRef]
  177. Jiang, Z.; Chen, Z.; Chen, J.; Liu, J.; Ren, J.; Li, Z.; Sun, L.; Li, H. Application of Crop Model Data Assimilation With a Particle Filter for Estimating Regional Winter Wheat Yields. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2017, 7, 4422–4431. [Google Scholar] [CrossRef]
  178. Orlova, Y.; Linker, R. Data assimilation with sensitivity-based particle filter: A simulation study with AquaCrop. Comput. Electron. Agric. 2023, 204, 107538. [Google Scholar] [CrossRef]
  179. Haiyun, B.I.; Jianwen, M.A.; Sixian, Q.; Hongjuan, Z.; Earth, S.D.; Sciences, C.A.O.; Sciences, U.O.A.O. Simultaneous estimation of soil moisture and hydraulic parameters using residual resampling particle filter. Sci. China Earth Sci. 2014, 57, 824–838. [Google Scholar]
  180. Mansouri, M.; Dumont, B.; Leemans, V.; Destain, M.-F. Bayesian methods for predicting LAI and soil water content. Precis. Agric. 2014, 15, 184–201. [Google Scholar] [CrossRef]
  181. Machwitz, M.; Giustarini, L.; Bossung, C.; Frantz, D.; Schlerf, M.; Lilienthal, H.; Wandera, L.; Matgen, P.; Hoffmann, L.; Udelhoven, T. Enhanced biomass prediction by assimilating satellite data into a crop growth model. Environ. Model. Softw. 2014, 62, 437–453. [Google Scholar] [CrossRef]
  182. Jamal, A.; Linker, R. Genetic Operator-Based Particle Filter Combined with Markov Chain Monte Carlo for Data Assimilation in a Crop Growth Model. Agriculture 2020, 10, 606. [Google Scholar] [CrossRef]
  183. Jafari, M.; Keshavarz, A. Improving CERES-Wheat Yield Forecasts by Assimilating Dynamic Landsat-Based Leaf Area Index: A Case Study in Iran. J. Indian Soc. Remote Sens. 2021, 50, 285–298. [Google Scholar] [CrossRef]
  184. Ziliani, M.G.; Altaf, M.U.; Aragon, B.; Houborg, R.; Franz, T.E.; Lu, Y.; Sheffield, J.; Hoteit, I.; McCabe, M.F. Intra-Field Crop Yield Variability by As-Similating Cubesat Lai in the Apsim Crop Model. Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 2022, XLIII-B3-2022, 1045–1052. [Google Scholar] [CrossRef]
  185. Ito, S.; Oda, K.; Koizumi, K. Seepage analysis model based on field measurement data for estimation of posterior pa-rameter distribution using merging particle filter. Soils Found. 2024, 64, 101442. [Google Scholar] [CrossRef]
  186. Ahmadi, K.; Salari, E. Social-spider optimised particle filtering for tracking of targets with discontinuous meas-urement data. IET Comput. Vis. 2017, 11, 246–254. [Google Scholar] [CrossRef]
  187. Dowd, M. Bayesian statistical data assimilation for ecosystem models using Markov Chain Monte Carlo. J. Mar. Syst. 2007, 68, 439–456. [Google Scholar] [CrossRef]
  188. Nguyen, H.D. An Introduction to Approximate Bayesian Computation. In Proceedings of the Statistics and Data Science, Melbourne, Australia, 24–26 July 2019; pp. 96–108. [Google Scholar]
  189. Craiu, R.V.; Gustafson, P.; Rosenthal, J.S. Reflections on Bayesian inference and Markov chain Monte Carlo. Can. J. Stat. 2022, 50, 1213–1227. [Google Scholar] [CrossRef]
  190. Robert, C.P.; Rousseau, J. A special issue on Bayesian inference: Challenges, perspectives and prospects. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2023, 381, 20220155. [Google Scholar] [CrossRef]
  191. Hartig, F.; Calabrese, J.M.; Reineking, B.; Wiegand, T.; Huth, A. Statistical inference for stochastic simulation models—Theory and application. Ecol. Lett. 2011, 14, 816–827. [Google Scholar] [CrossRef]
  192. Doucet, A.; Godsill, S.; Andrieu, C. On sequential Monte Carlo sampling methods for Bayesian filtering. Stat. Comput. 2000, 10, 197–208. [Google Scholar] [CrossRef]
  193. Han, X.; Li, X. An evaluation of the nonlinear/non-Gaussian filters for the sequential data assimilation. Remote Sens. Environ. 2007, 112, 1434–1449. [Google Scholar] [CrossRef]
  194. Guo, D.; Wang, X. Quasi-Monte Carlo Filtering in Nonlinear Dynamic Systems. IEEE Trans. Signal Process. A Publ. IEEE Signal Process. Soc. 2006, 54, 2087–2098. [Google Scholar]
  195. Gordon, N.; Salmond, D.J. Novel approach to nonlinear/non-Gaussian Bayesian state estimation. IEE Proc. F Radar Signal Process. 1993, 140, 107–113. [Google Scholar] [CrossRef]
  196. Banerjee, A.; Burlina, P. Efficient Particle Filtering via Sparse Kernel Density Estimation. IEEE Trans. Image Process. 2010, 19, 2480–2490. [Google Scholar] [CrossRef]
  197. Andrieu, C.; Doucet, A.; Holenstein, R. Particle Markov Chain Monte Carlo Methods. J. R. Stat. Soc. Ser. B Stat. Methodol. 2010, 72, 269–342. [Google Scholar] [CrossRef]
  198. Dumont, B.; Leemans, V.; Mansouri, M.; Bodson, B.; Destain, J.-P.; Destain, M.-F. Parameter identification of the STICS crop model, using an accelerated formal MCMC approach. Environ. Model. Softw. 2014, 52, 121–135. [Google Scholar] [CrossRef]
  199. Poterjoy, J.; Sobash, R.A.; Anderson, J.L. Convective-Scale Data Assimilation for the Weather Research and Fore-casting Model Using the Local Particle Filter. Mon. Weather Rev. 2017, 145, 1897. [Google Scholar] [CrossRef]
  200. Berliner, L.M. Hierarchical Bayesian Time Series Models. In Maximum Entropy and Bayesian Methods; Hanson, K.M., Silver, R.N., Eds.; Springer: Dordrecht, The Netherlands, 1996; pp. 15–22. [Google Scholar]
  201. Sansó, B. Noel Cressie, Chris Wikle: Statistics for Spatio-Temporal Data. Math. Geosci. 2013, 45, 253–254. [Google Scholar] [CrossRef]
  202. Yan, H.X.; Moradkhani, H. A regional Bayesian hierarchical model for flood frequency analysis. Stoch. Environ. Res. Risk Assess. 2015, 29, 1019–1036. [Google Scholar] [CrossRef]
  203. Kim, H.; Crow, W.T.; Wagner, W.; Li, X.; Lakshmi, V. A Bayesian machine learning method to explain the error characteristics of global-scale soil moisture products. Remote Sens. Environ. 2023, 296, 113718. [Google Scholar] [CrossRef]
  204. Alexandre, D.A.; Chaudhuri, C.; Gill-Fortin, J. Continental Scale Regional Flood Frequency Analysis: Combining Enhanced Datasets and a Bayesian Framework. Hydrology 2024, 11, 119. [Google Scholar] [CrossRef]
  205. Zujev, M.; Elken, J. Testing marine data assimilation in the northeastern Baltic using satellite SST products from the Copernicus Marine Environment Monitoring Service. Proc. Estonian Acad. Sci. 2018, 67, 217–230. [Google Scholar] [CrossRef]
  206. Sarigiannis, D.A.; Karakitsios, S.P.; Handakas, E.; Simou, K.; Solomou, E.; Gotti, A. Inte-grated exposure and risk characterization of bisphenol-A in Europe. Food Chem. Toxicol. Int. J. Publ. Br. Ind. Biol. Res. 2016, 98, 134–147. [Google Scholar]
  207. Sarigiannis, D.; Karakitsios, S. Advancing Chemical Risk Assessment through Human Physiology-Based Biochemical Process Modeling. Fluids 2019, 4, 4. [Google Scholar] [CrossRef]
  208. Daniel, G.; Esteban, R.; Oksana, G.; Rebeca, L.; Reza, A.; Thomaidis, N.S.; Montse, M.; Pablo, G.; Rubén, G. Comprehensive profiling and semi-quantification of exogenous chemicals in human urine using HRMS-based strategies. Anal. Bioanal. Chem. 2023, 415, 7297–7313. [Google Scholar]
  209. Fuentes, M. A formal test for nonstationarity of spatial stochastic processes. J. Multivar. Anal. 2004, 96, 30–54. [Google Scholar] [CrossRef]
  210. McMillan, N.J.; Holland, D.M.; Morara, M.; Feng, J. Combining numerical model output and particulate data using Bayesian space–time modeling. Environmetrics 2010, 21, 48–65. [Google Scholar] [CrossRef]
  211. Hadley, S.; Jones, E.; Johnson, C.; Wild-Allen, K.; Macleod, C. A Bayesian inference approach to account for multiple sources of uncertainty in a macroalgae based integrated multi-trophic aquaculture model. Environ. Model. Softw. 2016, 78, 120–133. [Google Scholar] [CrossRef]
  212. Garcia-Fernandez, A.F.; Grajal, J.; Morelande, M.R. Two-Layer Particle Filter for Multiple Target Detection and Tracking. IEEE Trans. Aerosp. Electron. Syst. 2013, 49, 1569–1588. [Google Scholar] [CrossRef]
  213. Beck, J.; Dia, B.M.; Espath, L.; Tempone, R. Multilevel double loop Monte Carlo and stochastic collocation methods with importance sampling for Bayesian optimal experimental design. Int. J. Numer. Methods Eng. 2020, 121, 3482–3503. [Google Scholar] [CrossRef]
  214. Zhang, C.; Shahbaba, B.; Zhao, H. Variational Hamiltonian Monte Carlo via Score Matching. Bayesian Anal. 2018, 13, 485–506. [Google Scholar] [CrossRef]
  215. Gunapati, G.; Jain, A.; Srijith, P.K.; Desai, S. Variational inference as an alternative to MCMC for parameter esti-mation and model selection. Publ. Astron. Soc. Aust. 2022, 39, 12. [Google Scholar] [CrossRef]
  216. du Toit, J.; du Preez, J.; Wolhuter, R. Low-density parity-check codes: Tracking non-stationary channel noise using sequential variational Bayesian estimates. Telecommun. Syst. 2024, 85, 247–262. [Google Scholar] [CrossRef]
  217. Emzir, M.F.; Lasanen, S.J.; Purisha, Z.; Roininen, L.; Särkkä, S. Non-stationary multi-layered Gaussian priors for Bayesian inversion. Inverse Probl. 2021, 37, 15002. [Google Scholar] [CrossRef]
  218. Marques-Carvalho, R.; de Almeida, C.M.; Escobar-Silva, E.V.; Alves, R.B.D.; Lacerda, C.S.D. Simulation and Pre-diction of Urban Land Use Change Considering Multiple Classes and Transitions by Means of Random Change Allocation Algorithms. Remote Sens. 2023, 15, 29. [Google Scholar]
  219. Ribeiro Lima, C.H. Hierarchical Bayesian and Machine Learning Models for Multiscale Hydroclimatic Analysis and Prediction for Brazil; Columbia University: New York, NY, USA, 2009. [Google Scholar]
  220. Kan, S.; He, Z.; Cen, Y.; Li, Y.; Mladenovic, V.; He, Z. Contrastive Bayesian Analysis for Deep Metric Learning. IEEE Trans. Pattern Anal. Mach. Intell. 2023, 45, 7220–7238. [Google Scholar] [CrossRef] [PubMed]
  221. Fu, S.; Zheng, R.; Chen, X. Adaptive Infrared and Visible Image Fusion Based on Visual Saliency and Hierarchical Bayesian. IEEE Trans. Instrum. Meas. 2022, 71, 16. [Google Scholar] [CrossRef]
  222. Ma, R.; Zhang, N.; Zhang, X.; Bai, T.; Yuan, X.; Bao, H.; He, D.; Sun, W.; He, Y. Cotton Verticillium wilt monitoring based on UAV multispectral-visible multi-source feature fusion. Comput. Electron. Agric. 2024, 217, 17. [Google Scholar] [CrossRef]
  223. Upreti, D.; Pignatti, S.; Pascucci, S.; Tolomio, M.; Huang, W.; Casa, R. Bayesian Calibration of the Aquacrop-OS Model for Durum Wheat by Assimilation of Canopy Cover Retrieved from VENµS Satellite Data. Remote Sens. 2020, 12, 2666. [Google Scholar] [CrossRef]
  224. Balzter, H. Remote sensing and global environmental change, by Samuel Purkis and Victor Klemas. Int. J. Geogr. Inf. Sci. 2013, 27, 1688–1689. [Google Scholar] [CrossRef]
  225. Wang, S.; Rao, Y.; Chen, J.; Liu, L.; Wang, W. Adopting “Difference-in-Differences” Method to Monitor Crop Response to Agrometeorological Hazards with Satellite Data: A Case Study of Dry-Hot Wind. Remote Sens. 2021, 13, 482. [Google Scholar] [CrossRef]
  226. Alexandridis, T.K.; Gitas, I.Z.; Silleos, N.G. An estimation of the optimum temporal resolution for monitoring vege-tation condition on a nationwide scale using MODIS/Terra data. Int. J. Remote Sens. 2008, 29, 3589–3607. [Google Scholar] [CrossRef]
  227. Alexandridis, T.K.; Katagis, T.; Gitas, I.Z.; Silleos, N.G.; Eskridge, K.M.; Gritzas, G. Investigation of aggregation effects in vegetation condition monitoring at a national scale. Int. J. Geogr. Inf. Sci. 2010, 24, 507–521. [Google Scholar] [CrossRef]
  228. Fei, S.; Xiao, S.; Zhu, J.; Xiao, Y.; Ma, Y. Dual sampling linear regression ensemble to predict wheat yield across growing seasons with hyperspectral sensing. Comput. Electron. Agric. 2024, 216, 108514. [Google Scholar] [CrossRef]
  229. Hovi, A.; Schraik, D.; Kuusinen, N.; Fabiánek, T.; Hanuš, J.; Homolová, L.; Juola, J.; Lukeš, P.; Rautiainen, M. Synergistic use of multi- and hyperspectral remote sensing data and airborne LiDAR to retrieve forest floor reflectance. Remote Sens. Environ. 2023, 293, 113610. [Google Scholar] [CrossRef]
  230. Szpakowski, D.M.; Jensen, J.L.R. A Review of the Applications of Remote Sensing in Fire Ecology. Remote Sens. 2019, 11, 2638. [Google Scholar] [CrossRef]
  231. Kulk, G.; Sathyendranath, S.; Platt, T.; George, G.; Suresan, A.K.; Menon, N.; King, H.E.; Abdulaziz, A. Using Multi-Spectral Remote Sensing for Flood Mapping: A Case Study in Lake Vembanad, India. Remote Sens. 2023, 15, 5139. [Google Scholar] [CrossRef]
  232. Xiong, Y.; Zhang, Q.; Chen, X.; Bao, A.; Zhang, J.; Wang, Y. Large Scale Agricultural Plastic Mulch Detecting and Monitoring with Multi-Source Remote Sensing Data: A Case Study in Xinjiang, China. Remote Sens. 2019, 11, 2088. [Google Scholar] [CrossRef]
  233. Wang, Z.; Shi, Y.; Hou, C.; Zhang, P. Sensitivity analysis of simulated Lycium barbarum L. yield in the WOFOST model under different climate conditions. Ecol. Model. 2024, 488, 110602. [Google Scholar] [CrossRef]
  234. Nie, W.; Kumar, S.V.; Peters-Lidard, C.D.; Zaitchik, B.F.; Arsenault, K.R.; Bindlish, R.; Liu, P. Assimilation of Remotely Sensed Leaf Area Index Enhances the Estimation of Anthropogenic Irrigation Water Use. J. Adv. Model. Earth Syst. 2022, 14, e2022MS003040. [Google Scholar] [CrossRef]
  235. Li, X.; Zhang, J.; Cai, X.; Huo, Z.; Zhang, C. Simulation-optimization based real-time irrigation scheduling: A human-machine interactive method enhanced by data assimilation. Agric. Water Manag. 2023, 276, 108059. [Google Scholar] [CrossRef]
  236. Wei, Z.; Ainong, L. A Downscaling Method for Improving the Spatial Resolution of AMSR-E Derived Soil Moisture Product Based on MSG-SEVIRI Data. Remote Sens. 2013, 5, 6790–6811. [Google Scholar] [CrossRef]
  237. Gao, H.; Wang, Q.; Gu, X.; Yang, J.; Liu, Q.; Tao, Z.; Qiu, X.; Zhang, W.; Shi, X.; Zhao, X. Study on Surface Reflectance Sampling Method and Uncertainty Based on Airborne Hyperspectral Images. Remote Sens. 2023, 15, 5090. [Google Scholar] [CrossRef]
  238. Mao, J.; Wang, Y.; Shi, E.; Hu, X.; Wang, Q.; Wang, J. Pre-Launch Multi-Energy Radiance Calibration of the OMS-N. Remote Sens. 2023, 16, 119. [Google Scholar] [CrossRef]
  239. Khalifani, S.; Darvishzadeh, R.; Azad, N.; Shayesteh, M.G.; Kalbkhani, H.; Akbari, N. Comparison of influential input variables in the deep learning modeling of sunflower grain yields under normal and drought stress conditions. Field Crops Res. 2023, 303, 109145. [Google Scholar] [CrossRef]
  240. Hao, S.; Ryu, D.; Western, A.W.; Perry, E.; Bogena, H.; Franssen, H.J.H. Global sensitivity analysis of APSIM-wheat yield predictions to model parameters and inputs. Ecol. Model. 2024, 487, 110551. [Google Scholar] [CrossRef]
  241. He, C.; Sun, J.; Chen, Y.; Wang, L.; Shi, S.; Qiu, F.; Wang, S.; Yang, J.; Tagesson, T. PROSPECT-GPR: Exploring spectral associations among vegetation traits in wavelength selection for leaf mass per area and water contents. Sci. Remote Sens. 2023, 8, 100100. [Google Scholar] [CrossRef]
  242. Jiang, Y.; Tang, B.-H.; Zhang, H. Estimation of downwelling surface longwave radiation for cloudy skies by considering the radiation effect from the entire cloud layers. Remote Sens. Environ. 2023, 298, 113829. [Google Scholar] [CrossRef]
  243. Lao, Z.; Fu, B.; Wei, Y.; Deng, T.; He, W.; Yang, Y.; He, H.; Gao, E. Retrieval of chlorophyll content for vegetation communities under different inundation frequencies using UAV images and field measurements. Ecol. Indic. 2024, 158, 111329. [Google Scholar] [CrossRef]
  244. Chen, P.; Ma, X.; Yang, G. An assimilation method for wheat failure detection at the seedling stage. Eur. J. Agron. 2022, 141, 11. [Google Scholar] [CrossRef]
  245. Liu, L.; Wan, X.; Li, J.; Wang, W.; Gao, Z. An Improved Entropy-Weighted Topsis Method for Decision-Level Fusion Evaluation System of Multi-Source Data. Sensors 2022, 22, 6391. [Google Scholar] [CrossRef]
  246. Han, Y.; Guo, J.; Ma, Z.; Wang, J.; Zhou, R.; Zhang, Y.; Hong, Z.; Pan, H. Habitat Prediction of Northwest Pacific Saury Based on Multi-Source Heterogeneous Remote Sensing Data Fusion. Remote Sens. 2022, 14, 5061. [Google Scholar] [CrossRef]
  247. Huang, B.; Li, X.; Zhang, X. SWPanGAN: A hybrid generative adversarial network for pansharpening. IET Image Process. 2024, 18, 1950–1966. [Google Scholar] [CrossRef]
  248. Myrgiotis, V.; Harris, P.; Revill, A.; Sint, H.; Williams, M. Inferring management and predicting sub-field scale C dynamics in UK grasslands using biogeochemical modelling and satellite-derived leaf area data. Agric. For. Meteorol. 2021, 307, 15. [Google Scholar] [CrossRef]
  249. Wang, Z.; Zhao, W.; Fu, J.; Xie, H.; Zhang, Y.; Chen, M. V-Shaped Toothed Roller Cotton Stalk Puller: Numerical Modeling and Field-Test Validation. Agriculture 2023, 13, 1157. [Google Scholar] [CrossRef]
  250. Gavahi, K.; Foroumandi, E.; Moradkhani, H. A deep learning-based framework for multi-source precipitation fusion. Remote Sens. Environ. 2023, 295, 17. [Google Scholar] [CrossRef]
  251. Li, W.; Chen, T.; Han, J.; Wang, L.; Huang, J. A High-Precision Map Matching Algorithm Based on Nonlinear Filtering Optimization for IOT Industrial Park. IEEE Sensors J. 2024, 24, 2209–2215. [Google Scholar] [CrossRef]
  252. Li, W.; Song, R.; Awais, M.; Ji, L.; Li, S.; Liu, M.; Lang, T.; Qi, H. Global Sensitivity Analysis of Crop Parameters Based on AquaCrop Model. Water Resour. Manag. 2024, 38, 2039–2058. [Google Scholar] [CrossRef]
  253. Kganyago, M.; Adjorlolo, C.; Mhangara, P.; Tsoeleng, L. Optical remote sensing of crop biophysical and biochemical parameters: An overview of advances in sensor technologies and machine learning algorithms for precision agriculture. Comput. Electron. Agric. 2024, 218, 15. [Google Scholar] [CrossRef]
  254. Barnes, C.; Brierley, C.M.; Chandler, R.E. New approaches to postprocessing of multi-model ensemble forecasts. Q. J. R. Meteorol. Soc. 2019, 145, 3479–3498. [Google Scholar] [CrossRef]
  255. Wang, Q.; Moreno-Martinez, A.; Munoz-Mari, J.; Campos-Taberner, M.; Camps-Valls, G. Estimation of vegetation traits with kernel NDVI. ISPRS J. Photogramm. Remote Sens. 2023, 195, 408–417. [Google Scholar] [CrossRef]
  256. Araghi, A.; Martinez, C.J. Evaluation of CRU-JRA gridded meteorological dataset for modeling of wheat production systems in Iran. Int. J. Biometeorol. 2024, 68, 1201–1211. [Google Scholar] [CrossRef]
  257. Mikhailenko, I.M. Estimation of Parameters of Biomass State of Sowing Spring Wheat. Remote Sens. 2022, 14, 1388. [Google Scholar] [CrossRef]
  258. El-Hendawy, S.; Al-Suhaibani, N.; Alotaibi, M.; Hassan, W.; Elsayed, S.; Tahir, M.U.; Mohamed, A.I.; Schmidhalter, U. Estimating growth and photosynthetic properties of wheat grown in simulated saline field con-ditions using hyperspectral reflectance sensing and multivariate analysis. Sci. Rep. 2019, 9, 15. [Google Scholar] [CrossRef] [PubMed]
  259. Zheng, C.; Wang, S.; Chen, J.M.; Xiao, J.; Chen, J.; Zhu, K.; Sun, L. Modeling transpiration using solar-induced chlorophyll fluorescence and photochemical reflectance index synergistically in a closed-canopy winter wheat ecosystem. Remote Sens. Environ. 2024, 302, 20. [Google Scholar] [CrossRef]
  260. Asani, E.O.; Osadeyi, Y.P.; Adegun, A.A.; Viriri, S.; Ayoola, J.A.; Kolawole, E.A. mPD-APP: A mobile-enabled plant diseases diagnosis application using convolutional neural network toward the attainment of a food secure world. Front. Artif. Intell. 2023, 6, 1227950. [Google Scholar]
  261. Yin, Z.; Peng, M.; Guo, Z.; Zhao, Y.; Li, Y.; Zhang, W.; Li, F.; Guo, X. Lightweight Pig Face Feature Learning Evaluation and Application Based on Attention Mechanism and Two-Stage Transfer Learning. Agriculture 2024, 14, 156. [Google Scholar] [CrossRef]
  262. Zhu, G.Y.; Raghavan, G.S.V.; Xu, W.X.; Pei, Y.S.; Li, Z.F. Online Machine Vision-Based Modeling during Cantaloupe Microwave Drying Utilizing Extreme Learning Machine and Artificial Neural Network. Foods 2023, 12, 1372. [Google Scholar] [CrossRef]
  263. Nichols, V.A. Anything but Simple: Using Models and Field Studies to Explore the Effects of Cropping System Diversification in the Upper Midwestern United States. Ph.D. Thesis, Iowa State University, Ames, IA, USA, 2021. [Google Scholar]
  264. Moon, T.; Kim, D.; Kwon, S.; Son, J.E. Process-Based Crop Modeling for High Applicability with Attention Mech-anism and Multitask Decoders. Plant Phenomics 2023, 5, 13. [Google Scholar] [CrossRef]
  265. Li, Z.X.; Sun, N.N.; Gao, H.T.; Qin, N.; Li, Z.C. Adaptive Subtraction Based on U-Net for Removing Seismic Multiples. IEEE Trans. Geosci. Remote Sens. 2021, 59, 9796–9812. [Google Scholar] [CrossRef]
  266. Antil, H.; Elman, H.C.; Onwunta, A.; Verma, D. A deep neural network approach for parameterized PDEs and Bayesian inverse problems. Mach. Learn. Sci. Technol. 2023, 4, 35015. [Google Scholar] [CrossRef]
  267. Pagani, S.; Manzoni, A.; Quarteroni, A. Efficient State/Parameter Estimation in Nonlinear Unsteady PDEs by a Reduced Basis Ensemble Kalman Filter. SIAM/ASA J. Uncertain. Quantif. 2017, 5, 890–921. [Google Scholar] [CrossRef]
  268. Xu, T.; Chen, F.; He, X.; Barlage, M.; He, X. Improve the Performance of the Noah-MP-Crop Model by Jointly Assimilating Soil Moisture and Vegetation Phenology Data. J. Adv. Model. Earth Syst. 2021, 13, e2020MS002394. [Google Scholar] [CrossRef]
  269. Ma, Y.; Hou, P.; Zhang, L.; Cao, G.; Sun, L.; Pang, S.; Bai, J. High-Resolution Quantitative Retrieval of Soil Moisture Based on Multisource Data Fusion with Random Forests: A Case Study in the Zoige Region of the Tibetan Plateau. Remote Sens. 2023, 15, 1531. [Google Scholar] [CrossRef]
  270. Zandler, H.; Faryabi, S.P.; Ostrowski, S. Contributions to Satellite-Based Land Cover Classification, Vegetation Quantification and Grassland Monitoring in Central Asian Highlands Using Sentinel-2 and MODIS Data. Front. Environ. Sci. 2022, 10, 19. [Google Scholar] [CrossRef]
  271. Tsakiris, G. Flood risk assessment: Concepts, modelling, applications. Nat. Hazards Earth Syst. Sci. 2014, 14, 1361–1369. [Google Scholar] [CrossRef]
  272. Shtull-Trauring, E.; Cohen, A.; Ben-Hur, M.; Israeli, M.; Bernstein, N. NPK in treated wastewater irrigation: Regional scale indices to minimize environmental pollution and optimize crop nutritional supply. Sci. Total Environ. 2022, 806, 150387. [Google Scholar] [CrossRef]
  273. Wolfe, D.W.; DeGaetano, A.T.; Peck, G.M.; Carey, M.; Ziska, L.H.; Lea-Cox, J.; Kemanian, A.R.; Hoffmann, M.P.; Hollinger, D.Y. Unique challenges and opportunities for northeastern US crop production in a changing climate. Clim. Chang. 2018, 146, 231–245. [Google Scholar] [CrossRef]
  274. Zare, H.; Weber, T.K.D.; Ingwersen, J.; Nowak, W.; Gayler, S.; Streck, T. Combining Crop Modeling with Remote Sensing Data Using a Particle Filtering Technique to Produce Real-Time Forecasts of Winter Wheat Yields under Uncertain Boundary Conditions. Remote Sens. 2022, 14, 1360. [Google Scholar] [CrossRef]
  275. Pan, H.; Chen, Z.; de Wit, A.; Ren, J. Joint Assimilation of Leaf Area Index and Soil Moisture from Sentinel-1 and Sentinel-2 Data into the WOFOST Model for Winter Wheat Yield Estimation. Sensors 2019, 19, 3161. [Google Scholar] [CrossRef] [PubMed]
  276. Wu, L.; Liu, X.-N.; Zheng, X.-P.; Qin, Q.-M.; Ren, H.-Z.; Sun, Y.-J. Spatial scaling transformation modeling based on fractal theory for the leaf area index retrieved from remote sensing imagery. J. Appl. Remote. Sens. 2015, 9, 96015. [Google Scholar] [CrossRef]
  277. Duveiller, G.; Lopez-Lozano, R.; Cescatti, A. Exploiting the multi-angularity of the MODIS temporal signal to identify spatially homogeneous vegetation cover: A demonstration for agricultural monitoring applications. Remote Sens. Environ. 2015, 166, 61–77. [Google Scholar] [CrossRef]
  278. Pashaei, M.; Starek, M.J.; Kamangir, H.; Berryhill, J. Deep Learning-Based Single Image Super-Resolution: An In-vestigation for Dense Scene Reconstruction with UAS Photogrammetry. Remote Sens. 2020, 12, 1757. [Google Scholar] [CrossRef]
  279. Fattori Junior, I.M.; dos Santos Vianna, M.; Marin, F.R. Assimilating leaf area index data into a sugarcane pro-cess-based crop model for improving yield estimation. Eur. J. Agron. 2022, 136, 126501. [Google Scholar] [CrossRef]
  280. Aleksovska, I.; Raynaud, L.; Faivre, R.; Brun, F.; Raynal, M. Design and Evaluation of Calibrated and Seamless Ensemble Weather Forecasts for Crop Protection Applications. Weather Forecast. 2021, 36, 1329–1342. [Google Scholar] [CrossRef]
  281. Huang, H.; Liu, Q.; Jiang, Y.; Yang, Q.; Zhu, X.; Li, Y. Deep Spatio-Temporal Attention-Based Recurrent Network From Dynamic Adaptive Functional Connectivity for MCI Identification. IEEE Trans. Neural Syst. Rehabil. Eng. 2022, 30, 2600–2612. [Google Scholar] [CrossRef]
  282. Gottwald, G.A.; Reich, S. Supervised learning from noisy observations: Combining machine-learning techniques with data assimilation. Phys. D Nonlinear Phenom. 2021, 423, 15. [Google Scholar] [CrossRef]
  283. Wei, J.; Luo, X.; Huang, H.; Liao, W.; Lei, X.; Zhao, J.; Wang, H. Enable high-resolution, real-time ensemble simulation and data assimilation of flood inundation using distributed GPU parallelization. J. Hydrol. 2023, 619, 15. [Google Scholar] [CrossRef]
  284. Chao, Z.; Fang, X.; Na, J.; Che, M. A Collaborative Sensing System for Farmland Water Conservancy Project Maintenance through Integrating Satellite, Aerial, and Ground Observations. Water 2021, 13, 2163. [Google Scholar] [CrossRef]
  285. Azua-Barron, M.; Arteaga-Ramirez, R.; Vázquez-Peña, M.A.; Quevedo-Nolasco, A. Calibración y evaluación de modelos matemáticos para calcular evapotranspiración de referencia en invernaderos. Rev. Mex. Cienc. Agríc. 2020, 11, 125–137. [Google Scholar] [CrossRef]
  286. Zare, H.; Weber, T.K.; Ingwersen, J.; Nowak, W.; Gayler, S.; Streck, T. Within-season crop yield prediction by a multi-model ensemble with integrated data assimilation. Field Crops Res. 2024, 308, 16. [Google Scholar] [CrossRef]
  287. Sánchez-Fernández, L.P.; Flores-Carrillo, D.A.; Sánchez-Pérez, L.A. Computer Model for an Intelligent Adjustment of Weather Conditions Based on Spatial Features for Soil Moisture Estimation. Mathematics 2024, 12, 152. [Google Scholar] [CrossRef]
  288. Sisheber, B.; Marshall, M.; Mengistu, D.; Nelson, A. The influence of temporal resolution on crop yield estimation with Earth Observation data assimilation. Remote Sens. Appl. Soc. Environ. 2024, 36, 101272. [Google Scholar] [CrossRef]
  289. Gu, F. On-demand data assimilation of large-scale spatial temporal systems using sequential Monte Carlo methods. Simul. Model. Pract. Theory 2018, 85, 1–14. [Google Scholar] [CrossRef]
  290. Li, H.; Jiang, Z.-W.; Chen, Z.-X.; Ren, J.-Q.; Liu, B.; Hasituya. Assimilation of temporal-spatial leaf area index into the CERES-Wheat model with ensemble Kalman filter and uncertainty assessment for improving winter wheat yield estimation. J. Integr. Agric. 2017, 16, 2283–2299. [Google Scholar] [CrossRef]
  291. Zhao, H. A Data-Driven Intelligent Decision-Making Model for Irrigation Scheduling. Ph.D. Thesis, George Mason University, Fairfax, VA, USA, 2023. [Google Scholar]
  292. Abhishek, A. Dynamics of Seasonal Crop Yield Prediction Under Weather and Climate Extremes; Michigan State University: East Lansing, MI, USA, 2023. [Google Scholar]
  293. Wei, Y.Y.; Law, A.W.-K.; Yang, C. Probabilistic optimal interpolation for data assimilation between machine learning model predictions and real time observations. J. Comput. Sci. 2023, 67, 9. [Google Scholar] [CrossRef]
  294. Muñoz, F.; Asenjo, R.; Navarro, A.; Cabaleiro, J.C. CPU and GPU oriented optimizations for LiDAR data pro-cessing. J. Comput. Sci. 2024, 79, 102317. [Google Scholar] [CrossRef]
  295. Fu, Y.; Zhou, W. A heterogeneous parallel implementation of the Markov clustering algorithm for large-scale biological networks on distributed CPU–GPU clusters. J. Supercomput. 2022, 78, 9017–9037. [Google Scholar] [CrossRef]
  296. Singh, V.K.; Pandey, H.; Singh, S.K. Groundwater storage change estimation using GRACE data and Google Earth Engine: A basin scale study. Phys. Chem. Earth 2023, 129, 12. [Google Scholar] [CrossRef]
  297. Buizza, C.; Casas, C.Q.; Nadler, P.; Mack, J.; Marrone, S.; Titus, Z.; Le Cornec, C.; Heylen, E.; Dur, T.; Ruiz, L.B.; et al. Data Learning: Integrating Data Assimilation and Machine Learning. J. Comput. Sci. 2022, 58, 16. [Google Scholar] [CrossRef]
  298. Dong, R.Z.; Leng, H.Z.; Zhao, J.; Song, J.Q.; Liang, S.T. A Framework for Four-Dimensional Variational Data As-similation Based on Machine Learning. Entropy 2022, 24, 264. [Google Scholar] [CrossRef]
  299. Vakayil, A.; Joseph, V.R. A Global-Local Approximation Framework for Large-Scale Gaussian Process Modeling. Technometrics 2024, 11, 295–305. [Google Scholar] [CrossRef]
  300. Zhang, Y.; Yang, Y. Bayesian model selection via mean-field variational approximation. J. R. Stat. Soc. Ser. B Stat. Methodol. 2024, 29, 742–770. [Google Scholar] [CrossRef]
  301. Gobin, A.; Sallah, A.-H.M.; Curnel, Y.; Delvoye, C.; Weiss, M.; Wellens, J.; Piccard, I.; Planchon, V.; Tychon, B.; Goffart, J.-P.; et al. Crop Phenology Modelling Using Proximal and Satellite Sensor Data. Remote Sens. 2023, 15, 2090. [Google Scholar] [CrossRef]
  302. Bahrami, M.; Talebbeydokhti, N.; Rakhshandehroo, G.; Nikoo, M.R.; Adamowski, J.F. A fusion-based data assimi-lation framework for runoff prediction considering multiple sources of precipitation. Hydrol. Sci. J. 2023, 68, 614–629. [Google Scholar] [CrossRef]
  303. Ma, Y.; Shao, D.; Wang, J.; Li, H.; Zhao, H.; Ji, W. Estimating Fractional Snow Cover in the Pan-Arctic Region Using Added Vegetation Extraction Algorithm. Remote Sens. 2023, 15, 775. [Google Scholar] [CrossRef]
  304. Yang, X.; Sun, L.; Tang, X.; Ai, B.; Xu, H.; Wen, Z. An Improved Fmask Method for Cloud Detection in GF-6 WFV Based on Spectral-Contextual Information. Remote Sens. 2021, 13, 4936. [Google Scholar] [CrossRef]
  305. Li, W.; Huang, Y.; Shen, Q.; Yao, Y.; Xu, W.; Shi, J.; Zhou, Y.; Li, J.; Zhang, Y.; Gao, H. Assessment of Seven Atmospheric Correction Processors for the Sentinel-2 Multi-Spectral Imager over Lakes in Qinghai Province. Remote Sens. 2023, 15, 5370. [Google Scholar] [CrossRef]
  306. Novelli, F.; Spiegel, H.; Sandén, T.; Vuolo, F. Assimilation of Sentinel-2 Leaf Area Index Data into a Physically-Based Crop Growth Model for Yield Estimation. Agronomy 2019, 9, 255. [Google Scholar] [CrossRef]
  307. Kimball, B.A.; Thorp, K.R.; Boote, K.J.; Stockle, C.; Suyker, A.E.; Evett, S.R.; Brauer, D.K.; Coyle, G.G.; Copeland, K.S.; Marek, G.W.; et al. Simulation of evapotranspiration and yield of maize: An inter-comparison among 41 maize models. Agric. For. Meteorol. 2023, 333, 22. [Google Scholar] [CrossRef]
  308. Thorp, K.R.; Hunsaker, D.J.; French, A.N. Assimilating Leaf Area Index Estimates from Remote Sensing into the Simulations of a Cropping Systems Model. Trans. ASABE 2010, 53, 251–262. [Google Scholar] [CrossRef]
  309. Zhuo, W.; Fang, S.; Gao, X.; Wang, L.; Wu, D.; Fu, S.; Wu, Q.; Huang, J. Crop yield prediction using MODIS LAI, TIGGE weather forecasts and WOFOST model: A case study for winter wheat in Hebei, China during 2009–2013. Int. J. Appl. Earth Obs. Geoinf. 2022, 106, 102668. [Google Scholar] [CrossRef]
  310. Huang, X.; Huang, G.R.; Yu, C.Q.; Ni, S.Q.; Yu, L. A multiple crop model ensemble for improving broad-scale yield prediction using Bayesian model averaging. Field Crops Res. 2017, 211, 114–124. [Google Scholar] [CrossRef]
  311. Zhang, S.; Deng, Y.; Niu, Q.; Zhang, Z.; Che, Z.; Jia, S.; Mu, L. Multivariate Temporal Self-Attention Network for Subsurface Thermohaline Structure Reconstruction. IEEE Trans. Geosci. Remote Sens. 2023, 61, 16. [Google Scholar] [CrossRef]
  312. Houtekamere, P.L.; Buehner, M.; De la Chevrotiere, M. Using the hybrid gain algorithm to sample data assimilation uncertainty. Q. J. R. Meteorol. Soc. 2019, 145, 35–56. [Google Scholar] [CrossRef]
  313. Yang, C.; Lei, H. Evaluation of data assimilation strategies on improving the performance of crop modeling based on a novel evapotranspiration assimilation framework. Agric. For. Meteorol. 2024, 346, 18. [Google Scholar] [CrossRef]
  314. Zhao, B.; Liu, M.; Wu, J.; Liu, X.; Liu, M.; Wu, L. Parallel Computing for Obtaining Regional Scale Rice Growth Conditions Based on WOFOST and Satellite Images. IEEE Access 2020, 8, 223675–223685. [Google Scholar] [CrossRef]
Figure 1. Application of crop model data assimilation in Precision Agriculture Distribution Statistics of Frequency variation with time.
Figure 1. Application of crop model data assimilation in Precision Agriculture Distribution Statistics of Frequency variation with time.
Agronomy 14 01920 g001
Figure 2. (a) changes in peer-reviewed publications over time, (b) the spatial geographical distribution of data assimilation studies. The color depth directly reflects the level of research activities and output in different regions.
Figure 2. (a) changes in peer-reviewed publications over time, (b) the spatial geographical distribution of data assimilation studies. The color depth directly reflects the level of research activities and output in different regions.
Agronomy 14 01920 g002
Figure 3. The development trends of different satellite sensors in detail, feature recognition, and scale requirements are compared. The red square represents the spatial resolution of adjacent RS images, modified from [70].
Figure 3. The development trends of different satellite sensors in detail, feature recognition, and scale requirements are compared. The red square represents the spatial resolution of adjacent RS images, modified from [70].
Agronomy 14 01920 g003
Figure 4. Sketch of crop model framework for remote sensing data assimilation. Note: The figure on the left is composed of remote sensing observations, crop models, and assimilation systems, and the figure on the right is the assimilation comparison between the regeneration assimilation process and the whole growth period. The green line indicates the addition of remote sensing observations, and the blue line indicates the addition of no remote sensing observations.
Figure 4. Sketch of crop model framework for remote sensing data assimilation. Note: The figure on the left is composed of remote sensing observations, crop models, and assimilation systems, and the figure on the right is the assimilation comparison between the regeneration assimilation process and the whole growth period. The green line indicates the addition of remote sensing observations, and the blue line indicates the addition of no remote sensing observations.
Agronomy 14 01920 g004
Figure 5. The development trend and evolution process of DA model. The upper left part represents the Earth observation data (EO), including various types of sensors, and the output is the observed values of physical and chemical parameters or the results of pre-processing and inversion. The lower right part represents the crop growth model simulation process (CGM), including initializing various types of parameter inputs, model simulation, and output as crop parameter simulation values. The middle part is the data assimilation model assimilation process (DA), including the core mechanism of different models, assimilation process, and evolution trend.
Figure 5. The development trend and evolution process of DA model. The upper left part represents the Earth observation data (EO), including various types of sensors, and the output is the observed values of physical and chemical parameters or the results of pre-processing and inversion. The lower right part represents the crop growth model simulation process (CGM), including initializing various types of parameter inputs, model simulation, and output as crop parameter simulation values. The middle part is the data assimilation model assimilation process (DA), including the core mechanism of different models, assimilation process, and evolution trend.
Agronomy 14 01920 g005
Figure 7. Strategy Framework of assimilation Model based on 4D-Var method. The left figure illustrates the process of simulating Grassland Aboveground Biomass (BM) and Leaf Area Index (LAI) with the ModVege grassland model. The right figure displays the distribution of BM and LAI after assimilation [135].
Figure 7. Strategy Framework of assimilation Model based on 4D-Var method. The left figure illustrates the process of simulating Grassland Aboveground Biomass (BM) and Leaf Area Index (LAI) with the ModVege grassland model. The right figure displays the distribution of BM and LAI after assimilation [135].
Agronomy 14 01920 g007
Figure 8. Crop model framework of wheat LAI assimilation SAFY based on sentinel-2 remote sensing images [74].
Figure 8. Crop model framework of wheat LAI assimilation SAFY based on sentinel-2 remote sensing images [74].
Agronomy 14 01920 g008
Figure 9. The PF assimilation method based on improved particle degradation consists of four steps: prediction, filtering, resampling, and merging [185].
Figure 9. The PF assimilation method based on improved particle degradation consists of four steps: prediction, filtering, resampling, and merging [185].
Agronomy 14 01920 g009
Figure 10. Diagrams of Metropolis–Hastings MCMC and PF. Light gray shows the outline of the target distribution (rear). The circle represents the combination of parameters in the algorithm. (a) the Metropolis–Hastings MCMC sampler proposes a new candidate value based on the last sampling value, and then accepts (green) or rejects (red) according to the ratio of the likelihood approximation of the reference point. (b) the sequential Monte Carlo sampler weights the point-by-point likelihood values from the initial set of parameter values and selects new candidates from the current set based on weight [191].
Figure 10. Diagrams of Metropolis–Hastings MCMC and PF. Light gray shows the outline of the target distribution (rear). The circle represents the combination of parameters in the algorithm. (a) the Metropolis–Hastings MCMC sampler proposes a new candidate value based on the last sampling value, and then accepts (green) or rejects (red) according to the ratio of the likelihood approximation of the reference point. (b) the sequential Monte Carlo sampler weights the point-by-point likelihood values from the initial set of parameter values and selects new candidates from the current set based on weight [191].
Agronomy 14 01920 g010
Figure 11. Variation and comparison of spatial, temporal, and spectral resolution of different satellite images [70].
Figure 11. Variation and comparison of spatial, temporal, and spectral resolution of different satellite images [70].
Agronomy 14 01920 g011
Table 1. Crop growth models commonly used for data assimilation and list of precision agriculture applications.
Table 1. Crop growth models commonly used for data assimilation and list of precision agriculture applications.
Model NameApplication of Agricultural AssimilationReferences
APSIMSimulate crop growth, soil water, and nutrient dynamics and crop management practices (suitable for crops such as wheat, corn, and soybeans)[33]
Agromet ShellSimulated crop growth and yield (Prediction of crop growth and yield changes based on meteorological data)[34]
AquacropResponse of simulated crops to water stress and yield prediction (Suitable for arid and semi-arid areas)[35]
CERES-wheatSimulate the growth and yield of wheat (Taking into account a variety of environmental and regulatory factors)[36]
CROPGRO-SoybeanSimulate crop growth and yield, such as soybeans (The physiological and ecological processes of crops, soil conditions, and management measures were considered.)[37]
CropsystSimulate the growth, development, and yield of a variety of crops (Suitable for different climate and soil conditions)[38]
DAISYSimulate the interaction between soil moisture, nutrients, and crop growth (Suitable for farmland and paddy field management)[39]
DSSATSimulate the growth, yield, and soil dynamics of many crops (Taking into account the impact of climate change and management practices on crops)[40]
EPICSimulate crop growth, soil erosion, nutrient cycling (Sustainable management of agro-ecosystem)[34]
GLAMSimulated crop growth and yield are affected by global climate change (The effects of climate change on crop physiological and ecological processes were considered)[41]
HERMESSimulation of crop growth and water balance (It is especially suitable for crop management in the Mediterranean climate region)[39]
STICSSimulate crop growth, soil moisture, and nutrient dynamics (Suitable for wheat, corn, and other crops and different soil types)[42]
SUCROSSimulate the growth and yield of sugar crops, such as sugarcane (The physiological and ecological processes of crops and climatic factors were considered)[43]
SWAPSimulation of water cycle and crop growth in farmland system (Suitable for arid and semi-arid areas)[44]
WOFOSTSimulation of crop growth, yield, and soil moisture dynamics (Crop physiological and ecological processes and climate change are considered)[45]
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

Wang, J.; Wang, Y.; Qi, Z. Remote Sensing Data Assimilation in Crop Growth Modeling from an Agricultural Perspective: New Insights on Challenges and Prospects. Agronomy 2024, 14, 1920. https://doi.org/10.3390/agronomy14091920

AMA Style

Wang J, Wang Y, Qi Z. Remote Sensing Data Assimilation in Crop Growth Modeling from an Agricultural Perspective: New Insights on Challenges and Prospects. Agronomy. 2024; 14(9):1920. https://doi.org/10.3390/agronomy14091920

Chicago/Turabian Style

Wang, Jun, Yanlong Wang, and Zhengyuan Qi. 2024. "Remote Sensing Data Assimilation in Crop Growth Modeling from an Agricultural Perspective: New Insights on Challenges and Prospects" Agronomy 14, no. 9: 1920. https://doi.org/10.3390/agronomy14091920

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