Next Article in Journal
Tannin-Based Strategies for Mitigating Greenhouse Gas Emissions Through Nitrogen and Carbon Metabolism in Ruminants
Previous Article in Journal
Research on a Coordinated Control Method of Tractor Electro-Hydraulic Hitch Tillage Depth and Travel Speed Based on Optimal Overall Efficiency and Economic Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Monitoring Agricultural Land Use Intensity with Remote Sensing and Traits

by
Angela Lausch
1,2,3,*,
Jan Bumberger
4,5,6,
András Jung
7,
Marion Pause
3,
Peter Selsam
4,5,
Tao Zhou
2,8 and
Felix Herzog
9
1
Department of Computational Landscape Ecology, Helmholtz Centre for Environmental Research—UFZ, Permoserstr. 15, D-04318 Leipzig, Germany
2
Landscape Ecology Lab, Geography Department, Humboldt-Universität zu Berlin, Unter den Linden 6, D-10099 Berlin, Germany
3
Department of Architecture, Facility Management and Geoinformation, Institute for Geo-Information and Land Surveying, Anhalt University of Applied Sciences, Seminarplatz 2a, D-06846 Dessau, Germany
4
Department of Monitoring and Exploration Technologies, Helmholtz Centre for Environmental Research—UFZ, Permoserstr. 15, D-04318 Leipzig, Germany
5
Research Data Management—RDM, Helmholtz Centre for Environmental Research—UFZ, Permoserstraße 15, D-04318 Leipzig, Germany
6
German Centre for Integrative Biodiversity Research (iDiv) Halle-Jena-Leipzig, Puschstraße 4, D-04103 Leipzig, Germany
7
Faculty of Informatics, Institute of Cartography and Geoinformatics, Eötvös Loránd University, Pázmány Péter sétány 1/A, H-1117 Budapest, Hungary
8
School of Resources and Environmental Engineering, Ludong University, Middle Hongqi Road 186, Yantai 264025, China
9
Agroecology and Environment, Agroscope, 8046 Zürich, Switzerland
*
Author to whom correspondence should be addressed.
Agriculture 2025, 15(21), 2233; https://doi.org/10.3390/agriculture15212233
Submission received: 22 July 2025 / Revised: 22 September 2025 / Accepted: 20 October 2025 / Published: 26 October 2025
(This article belongs to the Section Artificial Intelligence and Digital Agriculture)

Abstract

The intensification of agricultural land use (A-LUI) is a central driver of global environmental change, affecting soil health, water quality, biodiversity, and greenhouse gas balances. Monitoring A-LUI remains challenging because it is shaped by multiple management practices, ecological processes, and spatio-temporal dynamics. This review provides a comprehensive synthesis of existing definitions and standards of A-LUI at national and international levels (FAO, OECD, World Bank, EUROSTAT) and evaluates in situ methods alongside the rapidly expanding potential of remote sensing (RS). We introduce a novel RS-based taxonomy of A-LUI indicators, structured into five complementary categories: trait, genesis, structural, taxonomic, and functional indicators. Numerous examples illustrate how traits and management practices can be translated into RS proxies and linked to intensity signals, while highlighting key challenges such as sensor limitations, cultivar variability, and confounding environmental factors. We further propose an integrative framework that connects management practices, plant and soil traits, RS observables, validation needs, and policy relevance. Emerging technologies—such as hyperspectral imaging, solar-induced fluorescence, radar, artificial intelligence, and semantic data integration—are discussed as promising pathways to advance the monitoring of A-LUI across scales. By compiling and structuring RS-derived indicators, this review establishes a conceptual and methodological foundation for transparent, standardised, and globally comparable assessments of agricultural land use intensity, thereby supporting both scientific progress and evidence-based agricultural policy.

1. Introduction

Agricultural intensification represents a major economic development in recent decades on a global scale. However, this phenomenon is concomitant with significant environmental and economic changes, disruptions, and challenges. Agricultural intensification, otherwise termed land use intensity (A-LUI), is defined here as the augmentation in production output per unit of land through increased management intensity (utilisation of high yielding crops and livestock, inputs such as fertilisers, pesticides, drainage or irrigation, mechanisation) and/or the adaptation of landscape structure (increased field size through, e.g., land consolidation, removal of structural elements) [1]. While increasing A-LUI has facilitated the procurement of sustenance for an expanding global population, it goes along with substantial ecological concerns, including soil degradation, alterations in water quality and resources, biodiversity loss, and augmented greenhouse gas emissions, in addition to health hazards. For instance, the ongoing utilisation of synthetic fertilisers has resulted in soil acidification, thereby impacting the availability of nutrients to plants and the health of soil microbiota [2]. Furthermore, the excessive application of fertilisers can lead to significant nitrogen leaching and runoff of phosphorus, impacting water resources and soil fertility [3].
Use of heavy agricultural machinery leads to soil compaction, resulting in a reduction in both water and air permeability. This, in turn, has the potential to precipitate the occurrence of erosion and desertification over time [4]. The intensive use of water resources, which accounts for approximately 70% of total water consumption in agriculture worldwide [5], increases pressure on surface and groundwater, especially in regions where water is scarce. The quality of water is diminished by the mobilisation of salts due to low water tables and the introduction of fertilisers and pesticides into the underlying aquifers, which can threaten drinking water supplies [6,7]. Another pertinent issue is the escalating eutrophication of water bodies due to excessive nutrient inputs, which culminates in oxygen depletion and the demise of aquatic organisms [8]. Land use intensification exerts a profound influence on biodiversity [9,10,11,12]. The phenomenon of biodiversity loss [13] and the alteration in networks between biodiversity and ecosystem functions and services [14] are also impacted by land use intensification. The establishment of monocultures has resulted in the displacement of species-rich ecosystems, which in turn has been shown to lead to a decline in biodiversity, genetic impoverishment, and reduced resilience. Furthermore, the process of intensification has been shown to result in a multi-trophic homogenisation of grassland communities [15]. These developments have consequences for the resilience of ecosystems, resulting in the loss of essential ecosystem services such as pollination, pest control, and soil formation [16,17,18]. The expansion of agricultural land, frequently at the expense of forests, wetlands, and other semi-natural ecosystems, contributes to habitat fragmentation and destruction, biodiversity loss, and the release of greenhouse gases, which in turn further exacerbates climate change [13,19,20,21,22]. Consequently, the agricultural sector is a substantial contributor to global warming. In addition to the ecological consequences, the intensification of land use poses a significant health risk. The presence of persistent pollutants from herbicides in food can result in health complications, including cancer and neurological disorders [23]. The overuse of antibiotics in intensive livestock production has been demonstrated to promote the development of antibiotic resistance, which poses a significant threat to public health [24,25].
As scientific debate has long emphasised, accurate recording and quantification of A-LUI [26] is essential for the assessment of the impact of intensification on agroecological systems and for the development of sustainable management strategies. In situ measurements are of central importance, as they provide detailed information directly in the field (e.g., [27,28,29]). The merits of in situ measurements are twofold. Firstly, they enable direct observation of complex ecological and agronomic processes. Secondly, they facilitate the capture of locally specific variability that is often not considered in large-scale modelling. This is particularly true in heterogeneous landscapes, where minor variations in soil quality or microclimate or management practices can have substantial consequences for A-LUI. Consequently, such measurements are imperative. However, in situ measurements are often time-consuming, costly, and have limited spatial coverage, making their large-scale application and continuous monitoring difficult. Moreover, the comparability of results between different regions and studies is problematic due to a lack of standardisation.
RS has emerged as a key approach to quantify and assess A-LUI indicators on a large scale, in a timely and standardised manner, and over long periods of time [30,31]. As demonstrated in the works of [32,33,34,35,36,37,38], RS technologies facilitate spectral, spatial, and temporal analyses, providing detailed information on vegetation structure, soil condition, and other key land cover parameters. Furthermore, RS-based indicators of A-LUI, including yield estimates, vegetation indices (e.g., NDVI), and soil moisture parameters, which are crucial for the assessment of agroecological processes, have been derived for some time, including the advent of unmanned aerial vehicles (UAVs) and autonomous robotic platforms, in conjunction with the freely available space-based RS data (Landsat mission [39,40], the Copernicus mission Sentinel [41], and the hyperspectral mission (EnMAP) [42]). As demonstrated in the 2015 Copernicus Hyperspectral Imaging Mission (CHIME) [43], the LiDAR mission (GEDI) [44], and in the planned future missions such as the Hyperspectral Infrared Imager Mission (HyspIRI) [45] and the Fluorescence Explorer (FLEX) sensor [46], the derivation of standardised and improved A-LUI indicators will be significantly improved. The substantial body of literature on the derivation of A-LUI indicators using RS is indicative of this phenomenon [47,48,49,50]. As demonstrated in the works of Segarra et al. [51,52,53,54,55,56] and Hank et al. [38], the subject has been extensively researched.
A promising approach to capture and quantify A-LUI is to understand traits and trait variation in land cover, vegetation, and geodiversity [57]. Traits manifest at all spatial and temporal scales, making them ideal for standardised monitoring and the derivation of A-LUI indicators from local to global levels. All RS technologies record traits and trait variation in vegetation (example [58,59]), soil (example [60]), terrain and geomorphology (example [61,62]), and water (example [63]). RS allows the monitoring of traits and their status, related processes, disturbances or resource limitations in both terrestrial and aquatic ecosystems, and their interactions in a timely and standardised manner. Furthermore, RS data that capture traits have the capacity to establish a correlation between the sensitivity of the analysed environmental unit and various globally relevant pressures, including climate change and LUI with its socio-ecological consequences [64]. In addition, novel indicators for quantifying urban LUI have already been developed using RS and the trait approach [65,66]. Yet, to ensure the comparability of data and derived A-LUI indicators at both local and global scales, it is crucial to develop standardised methods for data collection and analysis. In recent years, there has been an increasing focus at the international level on the establishment of measurement standards. International organisations such as the Food and Agriculture Organisation of the United Nations (FAO) and the Intergovernmental Panel on Climate Change (IPCC) promote the establishment of international standards for the measurement and assessment of agricultural intensification across local and global scales. These organisations are increasingly recognising the value of RS and incorporating RS-based indicators into their standards and guidelines.
Nevertheless, the full potential of RS for the development of A-LUI indicators remains to be unlocked. In order to understand RS-based A-LUI indicators, derive new ones, and assess the suitability of different RS techniques for developing and categorising new indicators, we first need to define and structure these indicators and discuss them in context. We still lack a compendium offering a comprehensive overview of A-LUI indicators that can be derived using RS, so the objectives of this paper are as follows: (I) to define and compile standardised indicators for monitoring A-LUI at national, European, and global levels (FAO, OECD, World Bank, EUROSTAT); (II) to review and synthesise in situ methods for assessing A-LUI; (III) to introduce an RS-based definition of A-LUI by structuring them into five indicator categories: trait, genesis, structural, taxonomic, and functional A-LUI indicators; (IV) to illustrate the operationalisation of this taxonomy through numerous RS-based examples; (V) to link management practices, traits, and RS proxies to A-LUI indicators, validation strategies, and policy relevance; and (VI) finally to present innovative approaches for quantifying and evaluating A-LUI using RS.

2. Definition, Standards, and Programmes for Monitoring the A-LUI

2.1. Definition of A-LUI

Despite the significance of quantifying the A-LUI, the definition remains elusive, as the monitoring of anthropogenic changes and pressures/impacts on agricultural ecosystems/landscapes is a complex and multidimensional phenomenon [67] that is challenging to quantify [33,68]. As Diogo et al. [69] emphasise, the direction of change (positive or negative) of the A-LUI is also difficult to assess, as it depends on highly context- and scale-dependent processes that vary regionally, which have direct and indirect effects on the whole system and can mutually influence each other (increase or decrease). Conversely, the utilisation of inadequate (one-dimensional) indicators to quantify the A-LUI has been observed [70]. This is primarily due to the restricted availability of readily available local in situ data, such as pesticide, fertiliser, or machinery use, often due to data protection constraints and being frequently available only in aggregated form within reports. (1) The FAO reference does not define “intensity” (the term is not even used). It describes datasets but is not about their interpretation. (2) Limiting A-LUI to only the use of inputs is too narrow, particularly in the context of RS. Landscape simplification is another aspect of intensification, and it can actually be well captured by RS. Therefore, we refer to Diogo et al. [69] for an important indicator of A-LUI, which includes the main indicators of management intensity, landscape structure, and agricultural productivity.

2.2. Programmes for Monitoring A-LUI at National, European and Global Scale

One of the main challenges in monitoring A-LUI is the need to standardise measurement methods and indicators. In order to achieve national and international comparability in the monitoring of A-LUI, standardised programmes and indicators for the monitoring of A-LUI have been introduced at the national (Germany), European, and global level. The most important programmes and responsibilities for the monitoring of agricultural LCI for Germany, Europe, and the world are listed below.
  • National scale
  • Land Register: The land register records the types of land and their use in Germany. It is maintained by the state surveying and land registry offices. Most countries have detailed land register records of land type and ownership, maintained by the state surveying and land registry offices.
  • Agricultural Structure Survey: Regular surveys of agricultural land use, yields, livestock, etc., by National Statistical Offices.
  • IACS (Integrated Administration and Control System for Management Aid): In agriculture, the IACS system plays a central role in monitoring and managing data such as information on the use of plant protection products, fertiliser data, soil and water data, and yield and production data, as well as environmental and health data. The monitoring and control of IACS data in agriculture is carried out by different institutions and authorities, mainly at regional, national, and European level.
  • Europe
  • World
  • Global Land Cover (GLC): Several international initiatives produce global land cover maps, including projects supported by FAO and the United Nations Environment Programme (UNEP).
  • MODIS (Moderate Resolution Imaging Spectroradiometer): An instrument on NASA’s Terra and Aqua satellites that provides global data on land cover and land use change.
  • Global Land Analysis and Discovery (GLAD): A University of Maryland project to monitor global land use using high-resolution satellite imagery.
  • FAO (Food and Agriculture Organisation of the United Nations), OECD (Organisation for Economic Co-operation and Development), and World Bank (World Bank) use indicators to monitor A-LUI worldwide.
Table A1 provides an overview of the main agricultural land use intensity (A-LUI) indicators reported by major international organisations (FAO, OECD, World Bank, EUROSTAT). The indicators span input dimensions (e.g., fertiliser and pesticide use), output dimensions (e.g., crop yields), and structural aspects (e.g., land use statistics).

3. Approaches to Monitoring A-LUI

The monitoring of indicators to measure and assess A-LUI relies on both methods: in situ approaches provide detailed local information, while RS approaches, as physically based systems, capture status and change over large areas, though the underlying causes of change may differ. Therefore, coupling both approaches is essential. Traits represent the crucial connecting element between in situ and RS methods, as they can be directly measured in the field or indirectly derived from RS data. However, RS systems capture only a subset of traits—those with spectrally observable properties—which are, therefore, referred to as “spectral traits.” Figure 1 illustrates how the trait approach helps to bridge the two methodologies and underlines the complementary role of in situ and RS data for a comprehensive assessment of A-LUI.

3.1. In Situ Approaches

The measurement and monitoring of land use intensity represent a pivotal facet of land use research, particularly in the context of sustainable resource utilisation and ecosystem conservation. In situ methods have been shown to be a valuable tool for the collection of detailed data and analysis of land use in different geographical and agricultural contexts.
The following observations were made during the course of field studies. One of the fundamental approaches to measuring land use intensity is through direct observation and measurement in situ. These methodological approaches provide direct insights into the environmental and agricultural conditions on the ground. (I) Direct field measurements entail detailed investigations at specific sites where scientists record land use patterns, plant species, soil conditions, and other relevant parameters. The methodology encompasses the measurement of plots, the collection of soil and plant samples, and the observation of agricultural practices. Direct measurements are imperative in order to generate accurate data on A-LUI and to understand the interactions between land use and environmental conditions. (II) Field mapping constitutes a complementary method in which researchers are tasked with the production of maps delineating land use types by traversing the study area on foot or by vehicle. The cartographic representations under consideration here were originally produced on paper or using early graphical systems. They provide a visual representation of the spatial distribution of land use. These data are of pivotal significance for subsequent analysis and interpretation of land use intensity.
Surveys and interviews: In addition to direct field measurements, surveys and interviews represent an integral component of the collection of land use intensity data, as they encompass the human and social aspects of land use. They also record information that only the farmer will know, such as the type and quantity of pesticides and fertilisers used, etc. Structured interviews and surveys with landowners, farmers, and other land users can be used to collect information on land use practices, crop cycles, and irrigation methods. The collection of qualitative data facilitates the development of a more profound comprehension of the decision-making processes employed by land users, which are frequently influenced by economic, cultural, and political factors. Cultural and historical studies: The utilisation of cultural and historical studies is instrumental in facilitating a more profound comprehension of the historical evolution of land use patterns. The analysis of historical maps, archival records, and government reports provides valuable information on the long-term use and change in land areas and helps in understanding trends and shifts in land use.
The disciplines of analogue and digital cartography, as well as Geographic Information Systems (GIS), are discussed herein. The utilisation of mapping technologies and Geographic Information Systems (GIS) is of pivotal significance in the processes of recording and analysing land use intensity. These methodologies provide a comprehensive visual representation of the physical and agricultural traits of an area. Topographic maps: Topographic maps, produced by surveying, provide a basic representation of physical features such as contour lines, land cover, and infrastructure. These maps constitute a valuable source of data for the spatial analysis of land use. Aerial mapping: Prior to the advent of contemporary satellite technologies, aerial photographs were captured from aircraft and utilised to generate detailed cartographic representations. The interpretation of these images, frequently facilitated by the use of stereoscopes for three-dimensional viewing, enable a precise analysis of land use patterns and changes. The third point of the categorisation is Geographical Information Systems (GIS) and vector data. Geographic Information Systems (GIS) utilise vector data to display and analyse geo-referenced information on land use types and distributions. These systems facilitate sophisticated spatial analysis and monitoring of A-LUI indicators at local, national, and global levels.
The collection and analysis of agricultural yield data, as well as the maintenance of administrative records: The analysis of land use intensity is facilitated by quantitative and administrative information, which is provided by agricultural yield data and legal documents. Yield measurements: Yield data, frequently supplied by local or national agricultural authorities, offer insights into the productivity and utilisation of agricultural land. This information is indispensable for drawing conclusions on the intensity and efficiency of land use. Cadastral data: Cadastral data, encompassing land registry records and associated legal documentation, contains information pertaining to land ownership, delineated parcel boundaries and land use rights. These data are of crucial importance for the comprehension of formal land use patterns and their legal framework. IACS data: The IACS system occupies a pivotal position in the aggregation and administration of agricultural data within the European Union. The database under consideration encompasses a wide range of data, including, but not limited to, information pertaining to plant protection products; fertilisers; soil and water data; yield data; and production data. The systematised nature of these data facilitates the monitoring and evaluation of A-LUI.
Phenotyping laboratories: Contemporary phenotyping laboratories (e.g., Danforth Plant Science Centre, USA; IPK Gatersleben, Germany; JPPC, Germany; International Plant Phenotyping Network) utilise technologies such as automated imaging, sensors, drones, and robots to collect substantial data on plant growth, developmental disorders, soil, climate, and their interactions under laboratory conditions. This high-throughput phenotyping approach enables researchers to analyse numerous plants expeditiously and efficiently. Phenotyping laboratories are of significant importance in the context of A-LUI monitoring, as they facilitate the analysis and comprehension of the repercussions that intensive agricultural practices have on both plants and soils. This analysis encompasses the assessment of the impact on plants, including the enhancement of yield and the cultivation of stress resistance, as well as the investigation of the sustainability of land use, encompassing issues such as soil degradation. The following aspects should be monitored: erosion and nutrient depletion; resource efficiency (reduced fertiliser use, water-saving irrigation techniques); biodiversity and ecosystem services (monitoring the genetic diversity of crops and analysing their interaction with the environment (changes in genotype, phenotype, epigenetics)). Phenotyping laboratories are particularly well-suited to the testing and development of new sensor systems in a range of realistic and controlled cultivation scenarios (e.g., the FLuorescence EXplorer (FLEX) [71]. The testing of sensor prototypes on different plant species under controlled conditions, such as varying light conditions, temperature, and humidity, is a further method of evaluation. For instance, the RS-based indicator of solar-induced chlorophyll fluorescence (SIF) has been the subject of study in phenotyping laboratories, with a focus on monitoring plant stress [72]. Moreover, these data are imperative for the validation of novel sensors and the assessment of their measurement accuracy and efficiency.
The implementation of in situ A-LUI monitoring techniques frequently necessitates a considerable investment of labour, often resulting in protracted monitoring processes. These methodologies are further constrained to specific geographical areas and temporal frames. Nevertheless, they furnish significant insights into land use and A-LUI, derived from highly accurate local information. These methodologies form the foundation for contemporary, technologically advanced RS technology and data analysis techniques. It is, therefore, evident that the combination of in situ and RS approaches is imperative for effective A-LUI monitoring.

3.2. RS Approach

3.2.1. Principles of Monitoring A-LUI Using RS

All RS technologies are non-contact and detect traits and trait variations in land cover from a few millimetres (close range) to thousands (air-spaceborne) of kilometres (see Figure 2) away. RS sensors are integrated on various RS platforms such as wireless sensor networks (WSN), laboratory and field platforms, lysimeters (soil), phenocameras, masts, drones, balloons, as well as air- and spaceborne platforms. Different RS technologies (RGB/photographic, multispectral, hyperspectral, TIR, laser, radio/RADAR, and LiDAR) are often used in combination on many platforms. As traits and trait variations exist locally and globally, RS allows objective and continuous monitoring and derivation of standardised A-LUI indicators from a local to global scale.
The collection of indicators that quantify A-LUI is a crucial RS application that began with the availability of spaceborne RS data in the 1970s [73]. The focus here was on land cover monitoring, LULC and crop classifications, land use change [73,74], and the determination of basic functional vegetation traits using indicators such as NDVI [75]. The free availability and opening up of RS missions (such as Landsat [76], the Copernicus missions [77], or the hyperspectral mission (EnMAP [42])) accelerated the use and development of further RS-based A-LUI indicators. RS approaches are certainly ideal for deriving A-LUI indicators, as RS is based on the following basic principle: RS captures traits and trait variations directly or indirectly in plants, vegetation diversity, geodiversity, geomorphology, terrain, and water diversity. The spectral reflectance and absorption of pixels are, thus, the result of interactions between light (the atmosphere), phylogenetic/genetic, biophysical, biochemical, physical, morphological, physiological, phenotypic, structural, taxonomic, and functional characteristics of the recorded traits of vegetation diversity, geodiversity [12,78], and anthropogenic changes and disturbances by A-LUI. RS-based monitoring can, thus, capture indicators of A-LUI, as A-LUI is subject to complex and multidimensional influences, which are characterised by the interaction of abiotic–biotic compartments and anthropogenic factors (e.g., pesticide use, fertilisation, management) and their interactions.

3.2.2. Challenges of Monitoring A-LUI Using RS

The recording of A-LUI through RS brings numerous advantages but also specific challenges associated with the particularities of agricultural practices and sensor characteristics (spectral, spatial, temporal). For example, Maudet et al. [79] clearly emphasised in a comparative study that there are significant differences between in situ indicators and land use data derived from RS. They demonstrated that land cover maps based on RS are not a reliable indicator of management intensity at the field level, as the classifications of these maps do not adequately capture the A-LUI caused by agricultural practices. In addition, the landscape structure described by the area diversity varies significantly depending on the classification systems used. These differences strongly depend on the number of intensity classes considered, which we analysed with regard to the sensitivity of a target variable [79]. The following challenges exist when deriving A-LUI from RS data:
(1)
Limited coverage of agricultural practices
RS can identify different agricultural crops, but differentiating between intensive and extensive cultivation (e.g., conventional vs. organic farming, monocultures vs. crop rotation) is still a challenge. Spectral indices such as the NDVI only provide information on vegetation density and health but not direct information on the intensity of use, such as the use of fertilisers, pesticides, or irrigation systems. In order to record the use of fertilisers, pesticides, or irrigation systems using RS, which is often performed using indirect indicators or a set of indicators
Recording management practices: The way agricultural land is managed, such as the frequency of ploughing, crop rotation, or the use of agrochemicals, is crucial for A-LUI. These management practices can only be derived from RS data with a high geometric resolution (<1 m).
(2)
Seasonal dynamics
Agricultural areas go through different phases within a year (sowing, growth, harvest, fallow), which lead to significant changes in the vegetation. These seasonal variations can lead to misjudgements of the A-LUI if sufficient high-resolution, temporally dense data is not available. The challenge is to distinguish between natural seasonal variations and actual intensity changes. Multiple harvests: In regions with several harvests per year (e.g., in tropical areas), repeated RS images are required to correctly record the number and intensity of harvests. However, the temporal coverage of satellite images is often insufficient to fully document such multiple harvests. The use of RADAR data (Sentinel-1) in combination with optical RS data is expedient here, as they are recorded independently of cloud cover and at a high temporal density.
(3)
Irrigation and water management
Irrigation is a central factor of A-LUI, but the detection of irrigation systems is only indirectly possible through RS, e.g., by quantifying soil moisture or vegetation health. Especially in regions with periodic rainfall, it is difficult to distinguish between naturally occurring moisture changes and human-induced irrigation. Recognising water stress: RS can indicate the condition of vegetation, but it is often difficult to distinguish between natural causes (e.g., drought, inadequate soil properties) and the effect of intensive irrigation practices or water stress.
(4)
Fertiliser and pesticide use
The use of fertilisers and pesticides is a key factor in the intensity of agricultural production, but these inputs are virtually invisible to RS. While it is possible to infer the impact of these inputs on vegetation health (e.g., via spectral indices), there is no direct evidence of the amount or type of chemicals used.
Long-term soil degradation: Intensive use of fertilisers can have long-term effects on the soil, such as salinisation or nutrient depletion, but these are difficult to detect by RS. These effects are not directly reflected in the vegetation indices.
(5)
Small-scale agricultural structures
In many parts of the world, particularly in developing countries, agriculture is small-scale and heterogeneous. Small farmers often cultivate very small plots of land with different utilisation intensities. As a result, there are numerous problems with the demarcation of field boundaries using RS. For example, different plant species or land use types can have similar spectral signatures, which makes differentiation difficult. Furthermore, natural field boundaries are often not sharp, e.g., due to transition zones or hedges, which makes precise demarcation difficult. The spatial resolution of many RS data is often not sufficient to reliably capture these small-scale differences. High-resolution RS data (<1 m) is required here, but this is often expensive or not regularly available. For example, Landsat or Sentinel-2 data cannot be used to determine roads, field paths, or small structures [80], which is crucial for deriving field structures. Furthermore, Figure 3 shows the problems of the spatial resolution of RS data in the detection of crop vegetation using the example of an oilseed rape plant, which was recorded at different flight altitudes (1–80 m). There are currently only a few RS-based sensors that are freely available and can quantify high-resolution landscape structures and patterns (e.g., detection of agricultural utilisation boundaries, small structures) with sufficient spatial accuracy (see Table A2). In order to record the small-scale nature and utilisation structure, aerial image data (spatial resolution of 20 cm) is, therefore, repeatedly used, which is subsequently recorded vectorially and/or manually [81,82,83,84].
(6)
Agroforestry and mixed cropping:
In agroforestry systems or mixed cropping, it is difficult to derive the intensity of agricultural use from RS, as the different plant species are intertwined and are often grown under trees. Tree canopies can obscure the underplanting, so that important information about the agricultural intensity is lost.
(7)
Limited spectral information of RS data
While standard satellite sensors such as Landsat or Sentinel provide useful spectral information, these are often insufficient to capture subtle differences in the type and intensity of agricultural use. Hyperspectral RS sensors (e.g., EnMAP, DESIS) could provide more detailed information, but in many cases they are not widely available and their spatial resolution is limited to at least 30 × 30 m. Vegetation indices are often insufficient: spectral indices such as the NDVI can capture general biomass and vegetation health, but they do not provide detailed information on the intensity of agricultural activities (e.g., distinction between intensive and extensive cultivation).
(8)
Climatic and topographical influences
Weather events such as drought or flooding influence vegetation development and can make it difficult to separate differences in A-LUI from natural or climate-related influences. Topography and land cover: In hilly or mountainous regions and in areas with widely varying land cover (e.g., grassland and arable land next to each other), RS data may have difficulty providing accurate A-LUI data, as topography or shading may affect the quality of the data.
Table A3 provides a bridging overview that assigns each major challenge to one or more A-LUI definitions (trait, genesis, structural, taxonomic, functional) and lists potential RS- and AI-based solutions. This connection illustrates how the proposed framework can serve as a structured response to the practical difficulties of monitoring A-LUI with RS.

3.2.3. Separating A-LUI Indicators from Productivity and Spectral Signals

A key conceptual challenge in monitoring A-LUI with RS is the risk of conflating management intensity with signals of productivity potential or land cover change. High crop yields, for example, may result from intensive management (e.g., irrigation, fertilisation), but they may also reflect favourable soil and climate conditions. Similarly, RS-based indicators can inadvertently capture land cover conversion (e.g., expansion or abandonment) rather than intensity per se. To address this, we distinguish three components (Figure 4):
  • Management intensity signals—captured by traits and functional indicators (e.g., leaf nitrogen content, irrigation proxies, yield per unit input).
  • Biophysical potential signals—separated through normalisation with soil and climate data (e.g., adjusting NDVI or yield proxies for rainfall and soil fertility), or through modelling and domain adaptation approaches.
  • Land cover change dynamics—treated as a separate dimension under genesis indicators, where RS time series are used to track expansion, abandonment, or rotations.
Explicitly separating these dimensions can help avoid misinterpretations and ensure conceptual clarity and operational robustness.

4. Definition of A-LUI Using RS

In order to understand RS-based A-LUI indicators, to derive new ones and to understand the suitability of different RS technologies with regard to the development and categorisation of new indicators, a definition of A-LUI using RS data is required. A-LUI can be defined through RS as the combined expression of five complementary dimensions: trait, genesis, structural, taxonomic, and functional indicators (see Figure 5 and Figure A1).
(I) 
Trait Indicators of A-LUI: “Trait indicators describe measurable biophysical and biochemical properties of plants, soils, or water that respond directly to management intensity”. Examples using RS include leaf chlorophyll or nitrogen content derived from hyperspectral sensors (e.g., EnMAP, Sentinel-2 red-edge indices), leaf area index (LAI), or biomass estimated from multispectral vegetation indices such as NDVI or EVI, water stress, or photosynthetic activity monitored through solar-induced chlorophyll fluorescence (SIF) from the FLEX mission.
(II) 
Genesis Indicators of A-LUI: “Genesis indicators capture the temporal development and history of agricultural management practices, i.e., how intensity evolves over time”. Examples using RS include detect crop rotations, tillage events, or multiple harvests (Sentinel-1/2 time series); long-term Landsat archives documenting intensification trends such as the expansion of irrigated areas or the transition to monocultures.
(III) 
Structural Indicators of A-LUI: “Structural indicators describe the spatial configuration and arrangement of agricultural land, including field geometry and landscape elements”. Examples using RS include field size and shape derived from high-resolution optical imagery (e.g., PlanetScope); hedgerows and boundary elements identified through LiDAR or UAV mapping; landscape diversity indices (e.g., number of crop types per hectare) based on classified RS data.
(IV) 
Taxonomic Indicators of A-LUI: “Taxonomic indicators refer to the diversity and composition of crop species or land use types within an agricultural landscape”. Examples using RS include crop type classification using spectral signatures (e.g., distinguishing wheat vs. maize with Sentinel-2); detection of mixed cropping or agroforestry systems with hyperspectral UAV imagery; regional crop mapping from multi-temporal Sentinel-2 and Landsat data.
(V) 
Functional Indicators of A-LUI: “Functional indicators represent the ecological processes and services affected by agricultural intensity”. Examples using RS include crop productivity (e.g., yield estimates per hectare) derived from vegetation indices and biomass models; soil moisture inferred from radar backscatter (Sentinel-1) as a proxy for irrigation intensity; carbon sequestration potential or emission estimates based on biomass and soil models combined with RS observations.
To provide a detailed understanding of these five key dimensions of A-LUI indicators, Section 4.1, Section 4.2, Section 4.3, Section 4.4 and Section 4.5 elaborate on each category individually, namely trait, genesis, structural, taxonomic, and functional indicators, illustrating their derivation and application using RS approaches.

4.1. Monitoring the Trait Indicators of A-LUI Using RS

The recording and monitoring of traits form the basis for monitoring the genetic, taxonomic, structural, and functional A-LUI indicators using RS [94,95]. The monitoring of traits and trait variations (vegetation, soil, geomorphology, water) is, therefore, an essential basis for the assessment and management of A-LUI using RS. Traits are plant, soil, and hydrological properties that represent indicators of agricultural processes and their intensity. The targeted monitoring of such traits makes it possible to use resource inputs such as fertilisers, water, and pesticides more efficiently and, thus, make agricultural production more sustainable. The A-LUI traits refer directly to the extent of technological progress, the precision of the control of the resources used, and increases in efficiency in agriculture. The more precisely plant- and soil-related traits such as growth, yield, resistance to stress factors, or nutrient uptake can be monitored, the more effectively land use intensity can be controlled and optimised. Table A4 contains numerous examples, sensors, and references.

4.1.1. Trait Indicators of A-LUI—Spectranometric Approach

A particularly suitable approach for recording A-LUI is the spectranometric approach according to Greg Asner [95]. This method utilises, e.g., hyperspectral and multispectral RS data, which enables a detailed and direct recording of biochemical and structural characteristics of the vegetation (see Figure 6). The approach is characterised by several specific strengths: The method allows a detailed biochemical, structural, and functional characterisation of vegetation traits. Chemical characteristics such as nitrogen and chlorophyll content as well as concentrations of lignin, cellulose, and water content are precisely quantified using RS. As intensive agricultural use is typically associated with increased use of nitrogen fertilisers and pesticides, the resulting biochemical changes in the vegetation can be precisely recorded and spatially mapped. The hyperspectral approach allows precise quantification of plant structural characteristics such as leaf area index (LAI), leaf angle distribution, plant height, and biomass. These parameters are directly dependent on the type and intensity of cultivation, so that direct conclusions can be drawn about the intensity of land use. This method monitors the early detection of functional characteristics such as plant stress, for example, caused by water scarcity, over-fertilisation, or pest infestation. The detailed spectral signatures make stress symptoms visible at an early stage so that management decisions can be adapted and optimised in good time. By using hyperspectral RS technologies, which capture hundreds of narrow spectral bands, changes in plant physiology and soil can be measured and quantified in a differentiated manner. This allows a precise characterisation of the intensity of use at both field and landscape level. Finally, the spectranometric approach integrates hyperspectral data with ecological and agronomic models as well as satellite data from missions such as FLEX or Sentinel-3, enabling validated, precise, and in-depth statements about vegetation processes and the intensity of land use. The scientific significance of Greg Asner’s approach lies particularly in making complex ecological relationships such as biodiversity, carbon storage, and the effects of human activities on ecosystems comprehensible in detail. In the agricultural context, this enables a better understanding of sustainability and the ecological effects of different land use strategies. To summarise, the spectranometric approach offers a comprehensive, high-resolution, and differentiated method for the precise recording of A-LUI and, thus, represents an important basis for sustainable agricultural practices. Specific examples of monitoring the trait A-LUI indicators are as follows.

4.1.2. Trait Indicators of A-LUI—Chlorophyll Content

The measurement of chlorophyll content (Cab) using RS technology is of central importance, as this parameter is closely correlated with photosynthetic performance and, thus, plant vitality and productivity [96]. Chlorophyll serves as an effective indicator of A-LUI, as it reflects the influence of agricultural practices, fertiliser use, and plant health. Higher anthropogenic interventions, for example, through intensive fertilisation or the use of pesticides and precision agriculture, are directly reflected in changes in chlorophyll levels. An increased chlorophyll content often signals improved plant vitality, while stress factors such as drought, disease or nutrient deficiency can lead to a reduction in chlorophyll content. However, intensive management methods, including targeted plant protection measures, can partially compensate for such stress factors, which in turn results in more stable chlorophyll levels [96]. The importance of chlorophyll content arises from its role as an essential ecophysiological variable, which is closely linked to photosynthetic activity and, thus, to the vitality and productivity of plants [97]. In particular, the chlorophyll content provides information about nitrogen uptake and the general nutritional status of the vegetation. Plants in intensive farming show higher chlorophyll levels due to a higher nitrogen supply, whereas extensive or less intensively farmed systems typically have lower chlorophyll concentrations [98].
Hyperspectral RS techniques, which are characterised by their high spectral resolution and sensitivity to biophysical parameters, are primarily used for RS of chlorophyll content [97] (see Figure 7). Current and future hyperspectral missions such as PRISMA [99], HISUI [100], SHALOM [101], CHIME [43], or EnMAP [42] and others enable the acquisition of detailed spectral signatures, which form the basis for a precise estimation of the chlorophyll content. The Copernicus Hyperspectral Imaging Mission (CHIME) of the European Space Agency (ESA) in particular, with a spatial resolution of 20 to 30 m and a temporal repetition cycle of around 10–12 days, opens up new perspectives for monitoring chlorophyll content in agricultural contexts [43,96]. There are two main traditional approaches to determine chlorophyll content by RS: empirical regression techniques and physically based modelling approaches. Empirical techniques usually use spectral indices calibrated to field measurements but often show site-specific and vegetation-dependent limited transferability [102]. Physically based models, on the other hand, which are based on radiative transfer models (RTMs), are more robust and transferable, but require complex calibration and are computationally intensive [98,103]. More recently, the hybrid approach has become established, which combines physical models with machine learning and, thus, unites the advantages of both methods: the robustness of physical models and the efficiency of machine learning methods. Especially in combination with active learning techniques, this approach shows promising results in chlorophyll estimation and other vegetation parameters [96,97]. Despite the progress, challenges remain, such as spectral saturation effects at high chlorophyll levels or interference from ground reflections in open vegetation stands. In addition, the relationship between chlorophyll and nitrogen content can vary from species to species, which makes it difficult to apply universal models [104]. Therefore, hybrid approaches combining physical and data-driven methods are currently the most promising way to improve chlorophyll estimation by RS and ensure more precise monitoring of plant condition and nitrogen uptake in agriculture.

4.1.3. Trait Indicators of A-LUI—Chlorophyll Fluorescence

The Fluorescence Explorer (FLEX) sensor of the European Space Agency (ESA) [106] offers outstanding potential for the precise measurement of A-LUI (see Figure 8). By directly measuring solar-induced chlorophyll fluorescence (SIF), FLEX provides profound insights into the photosynthetic activity, vegetation health, and productivity of agricultural land [72,107]. The methodological suitability of FLEX for the assessment of A-LUI is based on several crucial factors: Firstly, FLEX directly measures photosynthetic activity, as SIF directly correlates with the photosynthetic rate of vegetation. Intensively used agricultural areas, characterised by increased use of fertilisers, irrigation, and pesticides, typically have higher fluorescence values, making FLEX a reliable tool for assessing A-LUI [72]. Secondly, the FLEX sensor allows early detection of plant stress, for example, caused by drought, nutrient deficiency, or over-fertilisation [108]. This early detection makes it possible to initiate targeted management measures before visible damage or significant yield losses occur [72]. Thirdly, with the FLORIS instrument (Fluorescence Imaging Spectrometer), FLEX has a high spectral and spatial resolution, which means that subtle differences in photosynthetic performance between intensively farmed areas can be precisely recorded. The spatial resolution of around 300 m allows detailed analyses and differentiated interpretations of land use intensity at a regional level [106]. Another methodological advantage is the integration of FLEX with Sentinel-3 satellite data. The synergetic use of optical and thermal sensors significantly improves the accuracy of deriving vegetation-relevant parameters such as leaf area index (LAI) and chlorophyll content. These parameters are essential for the comprehensive assessment of vegetation health and enable a differentiated assessment of agricultural utilisation intensity [107]. In addition, FLEX contributes significantly to the quantification of plant carbon sequestration, as SIF is closely linked to carbon uptake and, thus, to the global carbon cycle. This information is not only relevant for agricultural issues but also provides important insights for global climate modelling and sustainable development concepts [109].

4.1.4. Trait Indicators of A-LUI—Leaf Nitrogen Content

The monitoring of leaf nitrogen (leaf nitrogen, LN; leaf nitrogen content, LNC) as an indicator of A-LUI, provides important insights into the relationship between agricultural practices and plant physiology. Leaf nitrogen is an essential component of plant protein metabolism and plays a central role in photosynthesis. Intensively farmed agricultural areas, which are often characterised by increased use of fertilisers, generally have higher leaf nitrogen concentrations. This increased nitrogen availability promotes plant growth and increases productivity. A study by Dong et al. [110] emphasises that the allocation of nitrogen in leaf structures, especially in cell walls, increases with leaf mass per area (LMA), which indicates the importance of structural and metabolic components of leaf nitrogen. The intensity of land use influences not only the leaf nitrogen content but also the biodiversity of agroecosystems.
RS technologies have proven to be effective tools to measure LNC non-invasively and over large areas. There are a number of review studies on the detection of leaf nitrogen using RS technologies on different platforms [111,112,113,114,115,116,117]. Hyperspectral RS captures reflectance spectra of vegetation over a broad wavelength spectrum, which enables detailed analysis of leaf biochemistry. A study by Berger et al. [98] developed a hybrid method for estimating the aboveground nitrogen content of plants that combines physically based models with machine learning. This method identified specific wavelengths in the shortwave infrared (SWIR) range that are particularly relevant for nitrogen detection [98]. The use of hyperspectral RS technology opens up enormous potential for detecting the biochemical constitution of plant traits like the leaf nutrient content. For example, studies use hyperspectral technologies (such as EnMap [118] or Prisma [119]) to record the leaf nitrogen content. The use of UAVs RS technologies [120] in combination with advanced machine learning algorithms has increased the precision of LNC estimation. Zhang et al. [121] developed a self-supervised spectral–spatial transformer network using UAV imagery to accurately predict the nitrogen status of wheat fields. This model achieved high accuracy (0.96) and showed good generalisability for nitrogen status estimation [121]. Vegetation indices, such as the Normalised Difference Vegetation Index (NDVI), have traditionally been used to estimate LNC. However, more recent studies have developed more specific indices that are more sensitive to nitrogen variation. A study on estimating leaf nitrogen content in rice using vegetation indices emphasised the role of UAV-based RS in accurately determining nitrogen status at the field level [122]. The combination of different RS platforms, such as satellite imagery and UAVs, enables scalable and flexible monitoring of LNC. A comprehensive analysis of RS monitoring of nitrogen levels in rice and wheat crops over the last 20 years highlighted the importance of integrating different platforms to improve the accuracy and efficiency of nitrogen monitoring [111]. Traditional RS methods to determine leaf nitrogen (leaf N) content are usually based on indirect indicators, such as vegetation indices or chlorophyll-a+-b (Cab) content. However, these approaches reach their limits as the relationship between Cab and leaf N saturates at higher values and they are not very sensitive to early nutrient deficiency. A study by Y. Wang et al. [120] used Sentinel-2 satellite images to estimate various plant biochemical traits in large almond orchards in a two-year study. The traits, including leaf dry mass, leaf water content, and leaf Cab, were derived using a radiative transfer model and were used to explain the observed variability in leaf N. The resulting Sentinel-2 model for leaf N prediction showed high accuracy with an r2 of 0.82 and an nRMSE of 13%. Both the model performance and the contributing traits proved to be stable over the entire two-year period. The integration of these plant biochemical traits, thus, provides a more reliable and stable basis for leaf N prediction than conventional approaches, opening up promising prospects for application in precision agriculture (see Figure 9).
Table A4 presents a structured overview of trait-based A-LUI indicators, including concrete examples, the corresponding RS sensors, and the representative literature references.

4.2. Monitoring the Genesis Indicators of A-LUI with RS

Genesis indicators capture the temporal dynamics and historical development of A-LUI. They describe how management practices such as crop rotations, multiple harvests, tillage events, or land conversions evolve over time. RS is particularly suited to monitor these processes through dense time series, enabling the detection of management cycles and long-term trajectories of intensification.

4.2.1. Genesis Indicators of A-LUI—Subsurface Drainage

Subsurface drainage (DS) systems play an essential role in modern agriculture by efficiently draining excess water, thereby improving soil quality and agricultural productivity. Accurately locating and analysing these systems is crucial for sustainable land management, as unmapped drainage systems can lead to water quality degradation and increased nutrient inputs into water bodies [123]. Over the centuries, various civilisations such as the Egyptians, Chinese, and Indians developed their own drainage systems. In Europe, the drainage of agricultural land was established in the 17th century [124]. With the advent of motorised machinery in the 20th century, underground drainage systems spread rapidly, expanding agricultural land and making previously wet areas suitable for arable farming [125]. It is estimated that between 54% and 87% of the world’s wetlands have been lost since 1700 AD [126]. In addition to their positive effects on agricultural production, drainage systems also have undesirable side effects. They can accelerate the release of nutrients, especially nitrogen and phosphorus, into water bodies and, thus increase the risk of eutrophication [123]. In addition, draining carbon-rich wetlands can lead to increased CO2 emissions [127].
RS offers an efficient alternative to time-consuming manual investigations using ground penetrating RADAR and electromagnetic induction and enable large-area detection of drainage systems [128,129] (see Figure 10). The first attempts to record underground drainage systems using airborne thermal infrared images were made as early as the 1970s [130]. Multispectral and hyperspectral imaging utilises near-infrared (NIR) and shortwave infrared radiation (SWIR) to detect soil moisture. Vegetation indices such as NDVI and NDWI help to identify wet areas where drainage systems may not be working effectively [131]. RADAR RS such as Sentinel-1 enable the detection of soil moisture differences and help to recognise drainage patterns, even under cloudy skies or at night [132]. High-resolution digital terrain models (DTM/DEM) based on LiDAR RS data help to analyse natural and artificial drainage paths. LIDAR can also detect microtopographies that indicate inadequate drainage [133]. Moist or water-saturated soils have different temperatures than dry soils. Thermal infrared images (TIR), for example from Landsat 8, can be used to recognise drainage, especially after precipitation or at night [134,135]. Studies have shown that the combination of optical and thermal images can significantly increase detection accuracy [136].

4.2.2. Genesis Indicators of A-LUI—Terrace Mapping

Terrace fields are an important indicator for the genesis of A-LUI because they reflect the long-term adaptation and transformation of the landscape by humans. Here are some key reasons. Terraces were built to intensify the cultivation of slopes and to minimise soil erosion. These cultivation terraces are often found in steep, mountainous regions.
In the study by Liu et al. [137], RS data (Sentinel-1/2) was used as an efficient alternative for recording terrace structures, as it enables large-scale monitoring. However, optical satellite images, especially in mountainous regions, are affected by high cloud cover and varying vegetation cover, which makes precise detection of terrace fields difficult. Previous studies on automated terrace mapping using high-resolution satellite imagery, such as the GF-2 satellite mission or WorldView-1/3, have focussed primarily on the Loess Plateau in China, a region with comparatively less topographical challenges [87,138,139] (see Figure 11). This work mainly utilised optical RS data and applied object-oriented or deep learning methods for classification [140]. The use of high-resolution satellite images and digital terrain models (DEM) with an accuracy of 1–2 m significantly improves the recognition accuracy of terrace structures. However, these methods are limited for large-scale analyses due to high costs and a considerable volume of data [137]. Especially in mountainous regions, such as the analysed landscape in Southwest China, there are still significant challenges in the RS of terraces. Complex planting patterns, including crop rotation and mixed cropping, make it difficult to clearly identify terraces due to spectral similarities between different land cover classes [141]. In addition, low- to medium-resolution satellite images have a limited ability to detect small-scale terrace structures, as these often only appear as mixed pixels in heterogeneous landscapes [142]. LiDAR (Light Detection and Ranging) and RADAR (Radio Detection and Ranging) are key RS technologies for the detailed detection of terrace structures. They provide precise topographical information that is essential for analysing and managing such landscapes. LiDAR, in particular, enables the creation of high-resolution, three-dimensional terrain models, which allow reliable mapping of terrace structures even in densely forested areas [143].
An example of the application of this technology is provided by the study by Le Vot et al. [144], which aimed to reconstruct the historical development of land use on terraces. The aim of this study was to test the hypothesis of the resilience of these landscapes in the period from the 17th to the 21st century. For this purpose, current and archived geodata sets as well as LiDAR-based digital terrain models with a resolution of 1 m were used. The analysis was carried out in an area that was recently affected by an extreme event and whose reconstruction was considered a challenge. The results showed that the optimal utilisation of the terraces corresponded to the demographic optimum in the mid-19th century. After the Second World War, there was a gradual abandonment of the terraces, with significant differences between mountain regions. Nevertheless, the terraces remained intact despite these developments and survived the extreme event under investigation. This confirms the hypothesis of resilience and provides important insights for future strategies to revitalise these landscapes in the context of climate change.
In the study by Garzón-Oechsle et al. [145], a mobile LiDAR-based mapping system (MMS) without the use of UAVs was used to map the terrain around the documented stone architecture of the Manteños (ca. 650–1700 AD). The study area covered 1.2 km2 in the cloud forests of Bola de Oro, Manabí, Ecuador. The resulting digital terrain models (DTMs), when combined with soil surveys and archaeological excavations, revealed a Manteño landscape that had been significantly altered by the construction of agricultural terraces, drainage channels, and water retention basins. These structures were designed to store and distribute water from seasonal rainfall and marine layers at higher altitudes. The extensive investment in this sophisticated landscape is likely due to the fact that the Chongón-Colonche Mountains were considered resilient areas to extreme climate changes associated with the El Niño–Southern Oscillation (ENSO) during the Medieval Climatic Anomaly (MCA, ca. 950–1250 AD) and the Little Ice Age (LIA, ca. 1400–1700 AD) [145].

4.2.3. Genesis Indicators of A-LUI—Allmenden

Allmenden refers to communally used areas that played a central role in pre-modern agricultural societies. The term originates from the medieval legal system and referred to areas that were not privately owned by individuals but were used jointly by several or all members of a village community. In Europe, commons were widespread and were an important addition to private farmland, particularly in the three-field economy. In England, Germany, and other parts of Europe, numerous commons were privatised in the 17th–19th centuries, which often caused social tensions. Remnants of historical commons have been preserved, for example, in alpine pastures, heathland, or traditional co-operative forests
Modern RS methods can be used to effectively record historical field systems and commons. The combination of different technologies, including LiDAR (Light Detection and Ranging) as well as multispectral and hyperspectral satellite images, is particularly powerful. LiDAR has the advantage of being able to penetrate vegetation and detect fine ground elevations and structures. This makes it possible to identify relics of earlier landforms, vaulted fields, hedge structures, and medieval paths. A practical application example is the discovery of former three-field farming areas and commons that are now covered by woodland or modern agriculture. Medieval plough tracks and plot structures, particularly in Great Britain, Germany, and France, can also be detected using this method. In addition, multispectral and hyperspectral satellite images make it possible to differentiate between different soil types and vegetation cover, allowing conclusions to be drawn about historical agricultural use. Deviating vegetation structures also help to identify historical field boundaries. Former agricultural areas often show characteristic vegetation patterns or soil features that can be visualised using these techniques. Hyperspectral analyses also offer the possibility of identifying differences in moisture content, soil chemistry, or erosion patterns, which provides additional insights into past land use practices.
Edisa Lozić [146] analysed the use of airborne LiDAR data to discover, document, and interpret agricultural land use systems in the early medieval microregion of Bled (Slovenia). By combining LiDAR data with archaeological, geological, and pedological analyses, significant environmental variations within a microregion were identified. These enabled a detailed reconstruction of early medieval settlements and their agricultural use. The study by Masini et al. [147] investigated the effectiveness of LiDAR data for reconstructing the urban form of a medieval village near Matera, southern Italy. The research shows how LiDAR data can be used to reconstruct the urban structure and architectural features of historical settlements, even in densely forested or difficult to access areas.

4.2.4. Genesis Indicators of A-LUI—Deforestation

The recording of deforestation to gain pasture or arable land is an essential indicator of A-LUI. It allows a detailed analysis of human interventions in the environment, especially with regard to changes in the carbon balance, biodiversity loss, resource utilisation, and soil changes. Modern RS technologies offer precise methods for measuring these environmental changes and assessing their ecological consequences over longer periods of time. Global deforestation shows significant losses of forest area in different regions of the world. The study “Forest Pulse: The Latest on the World’s Forests” describes the latest trends in forest loss and deforestation and provides an up-to-date assessment of the global state of forests (https://gfr.wri.org/latest-analysis-deforestation-trends accessed on 19 October 2025). According to Smith et al. [148], the global forest cover was around 4.06 billion hectares, with approximately 420 million hectares lost between 1990 and 2020, mainly in tropical regions.
Slash-and-burn agriculture plays a significant role in the deforestation process and causes serious climate effects, including temperature increases, changes in precipitation patterns, and loss of biodiversity [149]. The use of unmanned aerial vehicles (UAVs) to analyse land cover during slash-and-burn has shown that multispectral imagery enables rapid and accurate assessment of land use change. In the future, this technology could serve as a standard method for recording slash-and-burn events [150]. The use of satellite imagery has proven to be one of the most efficient methods for the comprehensive and regular recording of deforestation. Optical satellites such as Landsat or MODIS provide high-resolution images that can be used to detect forest loss [151]. However, they are limited by weather conditions and cloud cover. RADAR systems such as Sentinel-1, on the other hand, work independently of light conditions and atmospheric influences, which makes them a reliable alternative for forest monitoring [152,153]. In addition, high-resolution satellite images make it possible to identify smaller deforested areas that are often overlooked in large-scale analyses [154]. The combination of different RS technologies can, thus, provide a comprehensive analysis of global deforestation and contribute to the development of effective conservation measures (see Figure 12).
Table A4 presents a structured overview of genesis indicators of A-LUI indicators, including concrete examples, the corresponding RS sensors, and the representative literature references.

4.3. Monitoring the Structural Indicators of A-LUI with RS

Structural indicators describe the spatial configuration and arrangement of agricultural land, including field size, shape, boundaries, and the presence or loss of semi-natural elements such as hedgerows, buffer strips, or terraces. These structural properties are closely linked to management intensity, as land consolidation, removal of landscape elements, and increasing field sizes typically indicate intensification. RS offers powerful tools to capture such patterns, ranging from high-resolution optical imagery and LiDAR data to radar-based mapping of field boundaries and landscape complexity. By quantifying structural diversity and fragmentation, RS enables a systematic assessment of how land use intensity reshapes the agricultural landscape.

4.3.1. Structural A-LUI Indicators—Crop Composition and Configuration

The quantification of landscape structure and the derivation of structural indicators play a decisive role in the monitoring of A-LUI. For example, the extraction of farmland boundaries from RS data is a key A-LUI indicator and supports agricultural planning, resource conservation, and sustainable development. Field boundaries are defined by changes in the type of crops planted, which are visible in RS data as discontinuities in grey value, colour, or texture. Wang et al. [88] provide a comprehensive overview of Farmland Boundary Extraction using RS data. Spatially high-resolution satellite images (≤1 m) such as WorldView-2/-3 (0.3–0.5 m), QuickBird (0.61 m), Pleiades (0.5 m), or GeoEye-1 (0.41 m) are particularly suitable for capturing field boundaries, as they allow fine structures such as narrow field paths and small plots to be captured. Medium-resolution satellite data (1–5 m) such as Sentinel-2 (10 m, with super-resolution at 5 m), Landsat 8 and 9 (30 m, for large-scale land use analyses), GF-2 (1 m, Chinese satellite), or RapidEye (5 m, multispectral available) are also suitable for large-scale analyses [88] (see Figure 13).
RS data enable the quantification of field sizes and their spatial distributions, which allow conclusions to be drawn about the degree of A-LUI and its management practices. Large, contiguous areas on which a single plant species is cultivated are indicative of industrial agricultural practices [155]. The arrangement of such monocultures can be easily recognised by RS and is a structural characteristic of intensive use. High-resolution satellites (Sentinel-2, WorldView, RapidEye) show that these agricultural areas appear as numerous small, geometric fields that are often separated by paths or hedges. Here, the degree of A-LUI is shown by small, highly parcelled fields, which gives an indication of the maximum utilisation of the available land [67]. Kümmerle et al. [31] use the image texture of Landsat data to derive the patch size, whereby the texture explained up to 93% of the variability of the field sizes in the study area in the border region between Poland, Slovakia, and Ukraine. The patch size (field size) indicator also offers a unique opportunity of investigating changes in land use that have occurred due to post-socialist land reform strategies, as many large agricultural areas have been parcelled out through privatisation. For example, Figure 14 shows a Landsat RS dataset in the 1990s, which clearly shows the state border between Saxony-Anhalt and Lower Saxony north of the Harz Mountains due to the change in patch size and small-scale parcelling.
In the study by Roilo et al. [67], various A-LUI indicators (e.g., field size, LULC_homogeneity) are used to analyse their effects on biodiversity. To calculate the field size, they used the LULC classification (2020, at 20 × 20 m resolution) [156], which was subsequently converted into polygons. The problem here is that not all crops could be properly classified using Sentinel-2 RS data. Furthermore, no roads and field paths could be included in the classification, which meant that the actual field size and the agricultural pattern could only be insufficiently quantified. In the study by Martin et al. [84], which deals with the effects of farmland heterogeneity on biodiversity, field size is emphasised as an important indicator. In order to improve the accuracy of the derivation of field size, it is often derived vectorially from aerial image data [82]. In this study, Mohr et al. [83] used aerial image data in combination with in situ data and interviews to answer the following question: Why has farming in Europe changed since the 1960s? In the study by Baessler and Klotz [157], historical and temporal time series of aerial image data were used to analyse changes in agricultural land use on landscape structure and arable weed vegetation over the last 50 years. The Interspersion and Juxtaposition Index (IJI) quantifies the mixing of different land use types and reflects the heterogeneity of the landscape. Higher IJI values indicate a more complex, diversified landscape, which has potentially positive effects on biodiversity [158].
The Shape Index is also used to analyse differences in land use patterns and management practices between different regions, such as East and West Germany. Such analyses can provide information on the impact of different management practices on landscape structure and function [159]. Furthermore, shape indicators can be used, for example, to estimate operational efficiency, to justify the merging of two field plots or to facilitate land consolidation projects [160]. In his study, Oksanen [160] uses various shape indicators such as convexity, compactness, triangularity, rectangularity, ellipticity, the ratio of principal moments, the radius of the inscribed circle, and the kerb index to classify the real field plots in order to quantify the operational efficiency (time and distance of the necessary travelling distance). Griffel et al. [161] examine the relationship between field shape and size and empirically derived crop efficiency to support assumptions related to the prediction of crop costs, greenhouse gas emissions, labour requirements, and other factors that affect the willingness to grow energy crops. Salas and Subburayalu [162] used Airborne Hyperspectral AVIRIS and HYDICE datasets to assess the potential of an optimised shape index to discriminate between tillage types (maize-min and maize-notill) and between grass/pasture and grass/trees, tree, and grass.
The indicator homogeneity of agricultural areas is a very good indicator for quantifying the A-LUI. Areas with high A-LUI are characterised by high homogeneity in species distribution and homogeneous spectral characteristics in contrast to organically cultivated areas with increased diversity of species (no use of pesticides) [163]. Blüthgen et al. [163] were able to prove through in situ measurements at 150 grassland sites in the Biodiversity Exploratories in three regions in Germany (Alb, Hainich, Schorfheide) that the vascular plant diversity in grassland sites in two regions (Alb and Hainich) decreased significantly with the A-LUI. Important work on the assessment of homogeneity from RS data of landscapes can be found in Rocchini et al. [164], which provides an overview of the current state of RS-based techniques for deriving spectral heterogeneity as a proxy of species diversity. Based on these approaches, Rocchini et al. [165] developed the Rao’s Q diversity index, which is considered a remotely sensed spatial heterogeneity indicator for taxonomic and functional plant species diversity [166].

4.3.2. Structural A-LUI Indicators—Surface Roughness of the Vegetation

Closely related to homogeneity is the surface roughness of the vegetation, which describes the structural variability of the vegetation surface and provides valuable information on plant architecture, stand density, species distribution, cultivation methods, and thus, A-LUI. Intensively managed fields with monocultural cultivation generally have low roughness (homogeneous stands), while more extensive, more diverse forms of cultivation or agroforestry systems have higher roughness. Steele-Dunne et al. [167] provide an overview of RADAR RS of agricultural canopies (see Figure 15). RADAR RS technologies can be used for a variety of applications resulting from the detection of the surface roughness of vegetation in agricultural areas. These range from crop classification, vegetation dynamics, vegetation phenology, water stress, and soil moisture derivation. Much of our understanding of vegetation backscatter from agricultural vegetation plots comes from SAR field-scale classification and monitoring studies [167]. Howison et al. [168] used Sentinel-1 RADAR data to quantify the spatial dynamics of surface roughness of vegetation in agricultural landscapes. Herrero-Huerta et al. [169] used the roughness of plant features (soya beans) using UAV aerial image data to estimate biomass in agricultural systems. Alfieri et al. [170] used the roughness, canopy structure, and configuration of vineyards to estimate the evapotranspiration loss required for irrigation and effective utilisation of limited water resources.

4.3.3. Structural A-LUI Indicators—Soil Roughness

Soil roughness is a crucial indicator for A-LUI as it allows direct conclusions on tillage practices, water balance, erosion processes, and vegetation development. Soil roughness is an inhomogeneous medium consisting of different types of soil textures, different shapes and sizes of stones, clods, SM gradients, organic matter, etc. The microwave signal incident on this layer is modified, scattered, and attenuated due to the physical and structural properties of this medium [171]. Soil roughness, thus, reflects various physical and agronomic processes. For example, the intensity of soil cultivation (e.g., ploughing, harrowing) changes the soil roughness considerably. High roughness often indicates intensive mechanical interventions, while low roughness indicates minimal soil turnover or conservation agriculture (see Figure 16 and Figure 17). Different crops and management practices produce specific roughness patterns. During a vegetation cycle, a gradual smoothing of the soil can be observed due to natural processes (rain, wind, biological activity) or renewed roughness formation due to agricultural interventions. Furthermore, high soil roughness favours water infiltration, as depressions can store water. Too little roughness, on the other hand, favours surface runoff and increases the risk of erosion. Heavily tilled and, therefore, less rough soils are more susceptible to erosion, especially in dry areas. Roughness can, therefore, be used as an indicator for the risk of erosion and the sustainability of cultivation. Soil roughness influences the temperature and moisture distribution on the surface. High roughness can reduce soil warming and influence evaporation rates. By monitoring roughness, conclusions can be drawn about plant growth. Heavily cultivated soils with low roughness could, for example, indicate a high use of fertilisers and irrigation. It can be analysed very well using RS methods such as RADAR and LiDAR technologies as well as optical sensors [172].
Table A4 presents a structured overview of structural indicators of A-LUI indicators, including concrete examples, the corresponding RS sensors, and the representative literature references.

4.4. Monitoring the Taxonomic A-LUI Indicators with RS

Taxonomic indicators capture the land use types within agricultural landscapes. They reflect whether farming systems are dominated by monocultures or characterised by mixed cropping, rotations, or agroforestry practices. Such diversity strongly influences ecological resilience and is a central dimension of land use intensity. RS enables taxonomic differentiation by exploiting spectral signatures, multi-temporal observations, and classification algorithms to distinguish crop types, detect rotations, or identify mixed stands.

4.4.1. Taxonomic A-LUI Indicators—Cropping Patterns

The monitoring of cropping patterns using RS is a key indicator of A-LUI. They enable precise characterisation of cropping intensity, harvest frequency, diversity, and management strategies. With the help of RS such as multispectral, hyperspectral, and RADAR data, changes can be analysed on a large scale, and long-term trends in agriculture can be identified [155,173,174].
Extensive agriculture shows more variable patterns with longer fallow periods, especially in semi-arid or mountainous regions, and relies on crop rotation, mixed cropping, or agroforestry. Single cropping indicates low intensity, while double/multi-cropping indicates high A-LUI, often under irrigated conditions in tropical and subtropical areas. Intercropping increases vegetation variability and is often used in sustainable agricultural systems. High A-LUI is associated with short or no fallow periods, while low A-LUI has longer fallow periods for soil regeneration. Long-term changes in cropping patterns can be indicators of soil degradation, water scarcity, or climate change, which is why the adoption of diversification strategies such as agroforestry and mixed cropping as sustainable measures against overexploitation is essential. Mahlayeye et al. [173] give a very good overview of the detection of cropping patterns using RS. Optical sensors are most commonly used for mapping single cropping, especially those with high spatial resolution, such as UAVs. These sensors enable the precise identification of single crop fields but are often only suitable for smaller areas. For large-scale (regional/global) analyses, on the other hand, medium to coarse resolutions are usually used, such as Spot, Landsat 8, MODIS, or Sentinel-2 [175,176]. In addition to optical sensors, microwave sensors with high temporal resolution, such as RADARSAT-2 or Sentinel-1 [177], are also used, particularly for rice cultivation in Asia or maize in Africa. Some studies have combined microwave and optical sensors for more precise crop mapping [178]. In addition, hyperspectral or LIDAR sensors are increasingly being used [179,180,181]. Studies on the mapping of individual crops are based on phenology and the spatial distribution of crops. Mapping individual crops using single images may be insufficient, as plants change during the growing season. Continuous monitoring of plant development is, therefore, necessary. Overall, the analysis shows that single crop cultivation can be successfully mapped at both local and regional levels with high spatial and temporal resolution [173]. The mapping of multiple cropping and sequential cropping systems is carried out at different levels using optical sensors with high temporal resolution [173]. MODIS satellite data are frequently used [182,183], while Indian RS (IRS) satellites and the Wide Field Sensor (WiFS) [184] and Sentinel-2 [174] are also used in some studies. The detection of triple cropping patterns was also carried out [182]. Microwaves (Sentinel-1 C-band time series data) and optical sensors enable the creation of detailed temporal profiles of sequential crops [185,186]. Commonly mapped crops are maize, rice, wheat, and soybeans, with studies on sequential cropping patterns increasingly being conducted in tropical regions characterised by long rainy seasons.
Mapping sequential cropping patterns is more complex than single cropping, as different crops are planted in the same growing season, requiring a longer growing season and more continuous ground cover [173]. In particular, high-resolution multispectral and hyperspectral imaging provide valuable insights into the structure and dynamics of mixed crops. Vegetation indices such as the NDVI (Normalised Difference Vegetation Index) or the EVI (Enhanced Vegetation Index) help to differentiate between different plant species based on their spectral reflectance properties, while hyperspectral sensors enable even more precise differentiation by analysing specific wavelength ranges.

4.4.2. Taxonomic A-LUI Indicators—Crop Classifications

A study on high-resolution mapping of the German agricultural landscape using RS provides detailed insights into parcelling and field structures through crop classification. RS-based classifications of agricultural land use for the years 2017–2020 for the German agricultural landscape (grid of 10 m × 10 m) provide detailed insights into area size, distribution, and crop types cultivated (https://ows.geo.hu-berlin.de/webviewer/landwirtschaft/index.html accessed on 19 October 2025 [90], see Figure 18. Crop types such as rapeseed or sugar beet can be differentiated very well. However, species that are spectrally very similar in the course of the growth phases or in their appearance (e.g., winter wheat and triticale) or that differ solely in terms of their type of utilisation (e.g., silage maize and grain maize) cannot yet be recorded with sufficient accuracy using RS. Patterns of land use intensity, such as crop rotation or fallow periods, can be effectively captured by time-series RS data, providing insights into the sustainability of agricultural practices. Preidl et al. [156] used Sentinel-2A imagery data for crop classification on the national scale (Germany).
The global distribution of A-LUI is crucial for understanding agricultural land use. Previous studies used coarse-resolution data, which are unsuitable for heterogeneous landscapes. To fill this gap, Zhang et al. [187] developed the global, spatially continuous CI dataset GCI30 with 30 m resolution using Landsat 7 ETM+, Landsat 8 OLI, and Sentinel-2 MSI time series during 2016–2018. GCI30 captures global patterns and spatial details, with monocultures dominating 81.57% of cropland. Regional differences reflect natural and anthropogenic influences [187]. Howison et al. [168] developed a new RADAR-based RS technique for large-scale quantification of A-LUI. The method utilises the temporal stability of RADAR signals to capture differences in land use and provides more precise tracking of A-LUI at the landscape scale.

4.4.3. Taxonomic A-LUI Indicators—Intensification of Grassland

The intensification of grassland utilisation (e.g., more frequent mowing, increased grazing) significantly impairs biodiversity and ecosystem services. However, detailed information on utilisation intensity is usually locally limited. Numerous studies show [35,188] that mowing events can be mapped over large areas using satellite image time series. Time-series phenology can overcome limitations of classification-based mapping approaches, especially when characterising grassland use intensity, using the frequency and timing of mowing events as important indicators [189]. Lange et al. [188] developed a method for the RS-based derivation of grassland intensity for Germany (www.ufz.de/land-use-intensity accessed on 19 October 2025). Based on Sentinel-2 time series (spatial resolution of 20m) from 2017 to 2018, the NDVI time series data and available in situ indicators (grazing intensity, mowing frequency, and fertiliser use) of the Biodiversity Exploratories for Germany [190] were used to train and derive a continuous A-LUI index for grassland for Germany using Convolutional Neural Networks (CNN). An overall classification accuracy of up to 66% for grazing intensity, 68% for mowing, and 85% for fertilisation was achieved. Weber et al. [35] developed a rule-based algorithm for mapping mowing and grazing events in Switzerland (2018–2021) based on Sentinel-2 and Landsat-8 data. The validation was carried out with time-series data from public webcams. The review (2020–2021) showed that ≥78% of the recorded events reflect actual management, but up to 57%—especially grazing events at higher altitudes—were not recognised. Bartold et al. [91] present a comprehensive study on the classification of management intensity of grasslands in two different regions of Poland (see Figure 19). By using Sentinel-1 and Sentinel-2 data synergistically, different intensity types could be identified, allowing conclusions to be drawn about herbicide use.
Table A4 presents a structured overview of taxonomic indicators of A-LUI indicators, including concrete examples, the corresponding RS sensors, and the representative literature references.

4.5. Monitoring the Functional A-LUI Indicators with RS

Functional indicators describe the ecological processes and ecosystem services that are directly affected by A-LUI. They include aspects such as crop productivity, soil fertility, carbon and nutrient cycling, irrigation demand, and greenhouse gas emissions. These functions are critical for assessing sustainability, as they link agricultural practices to environmental impacts and resource efficiency. RS contributes to functional monitoring by providing proxies for biomass production, evapotranspiration, soil moisture, and photosynthetic activity. Combined with modelling approaches, RS-derived functional indicators allow for large-scale assessments of agricultural performance, efficiency, and environmental trade-offs.

4.5.1. Functional A-LUI Indicators—Plant Density and Biomass Production

RS for recording plant density and biomass production is essential for analysing vegetation structures and assessing the A-LUI, as it reflects the direct effects of management practices on vegetation. RS technologies enable the area-wide analysis of vegetation parameters using the Normalised Difference Vegetation Index (NDVI), the Soil-Adjusted Vegetation Index (SAVI) and the Enhanced Vegetation Index (EVI) to minimise soil and atmospheric influences in order to determine plant density and biomass production. Recent studies, for example, by Sousa Júnior et al. [191], demonstrate the successful use of Landsat 8 to estimate aboveground biomass in agricultural mosaics. The combination of different sensor systems, especially optical- and RADAR-based RS technologies, improves the accuracy of biomass estimates [192]. The use of unmanned aerial vehicles (UAVs) offers an efficient alternative due to high spatial resolution and flexibility [193,194]. Da et al. [194] combined UAV-derived spectral, textural, and structural features for biomass monitoring of soybean and achieved a model accuracy of R2 = 0.85. Spaceborne RS data are also used to assess plant biomass. Breunig et al. [195] used PlanetScope and Sentinel-1 SAR data to monitor intercrop biomass in Southern Brazil. Hagn et al. [196] analysed Sentinel-2 data to model crop-specific biomass yield potential in precision farming. Their results showed a strong correlation between relative biomass potential (r = 0.62–0.73) and soil properties such as soil organic carbon (SOC) and total nitrogen (TN). However, optical satellite systems such as Landsat and Sentinel-2 are weather-dependent and do not collect data under cloudy conditions. To overcome this limitation, Planet developed the Biomass Proxy product [197], which provides a daily, ready-to-analyse biomass estimate with 10 m spatial resolution. The Biomass Proxy algorithm fuses Sentinel-1 and Sentinel-2 data and enables continuous monitoring of vegetation. This facilitates the early detection of growth anomalies, the assessment of crop yields, and the identification of potential environmental hazards and supports informed agronomic decision-making. The difference map of Aboveground Biomass (AGB) estimates of 18 August 2017 and 26 August 2017 derived from PlanetScope (PS) optical, Sentinel-1 SAR, and hybrid (optical plus SAR) datasets is shown in Figure 20.

4.5.2. Functional A-LUI Indicators—Pesticide, Herbicide, and Fungicide

The use of pesticides and herbicides is an important indicator for assessing the A-LUI. The use of pesticides and herbicides causes vegetation-related changes and stress reactions in plant populations, which can be recorded using RS via vegetation anomalies. Herbicides specifically influence metabolic processes by disrupting biochemical reactions, e.g., triazines (atrazine) lead to the inhibition of photosynthesis, glyphosate to the blocking of amino acid synthesis, or auxin analogues (2,4-D) to the impairment of cell growth. These effects can be detected in the short term by spectral analyses of RS data.
The RS-based recording of pesticide intensity is a growing field of research with the aim of making crop protection more efficient and environmentally friendly, as well as being able to detect the use of pesticides and herbicides. The use of satellite images, drones, and hyperspectral sensors allows conclusions to be drawn about the use and distribution of pesticides. While current applications are primarily focussed on laboratory analyses with hyperspectral sensors (e.g., ASD, MSV-500) [198,199,200,201,202], space-based RS data such as Sentinel-2 are also being used [203]. Spectral reflectance data, particularly in the red and near-infrared range, enable the calculation of vegetation indices such as NDVI, whose changes indicate herbicide applications and associated stress reactions [203]. Hyperspectral RS captures detailed spectral signatures that can identify specific pesticide applications and their effects [199]. For example, hyperspectral imaging combined with machine learning has been used to detect herbicide stress early and identify new sites of action. Zhang et al. [200] extracted the Physiological Reflectance Index (PRI) and NDVI from hyperspectral images and classified glyphosate-induced plant damage using Support Vector Machine (SVM). Chu et al. [201] used neural networks to identify different herbicide damage to wheat, finding significant spectral differences in the wavelength ranges 518–531 nm, 637–675 nm, and at the red edge. Pon Arasan et al. [204] analysed UAV-based mapping methods to optimise herbicide use. Bartold et al. [91] combined Sentinel-1 and Sentinel-2 data to classify management intensities in Polish grasslands and identified herbicide applications. Bautista et al. [91] investigated the efficiency of drone applications with cyhalofop-butyl in Spanish rice fields using NDVI analyses with Sentinel-2. Sentinel-2 and Landsat-8/9 are suitable for general monitoring, while PRISMA and EnMAP enable more precise spectral analyses. WorldView-3 offers high spatial resolution for detailed field studies. The combination of these satellites allows the monitoring of pesticide and herbicide use. Exemplary application of SugarViT (Vision Transformer based model for disease severity) for disease severity prediction in sugar beet using UAV multispectral data is shown in Figure 21; each prediction is completely independent of its surrounding predictions [92].

4.5.3. Functional A-LUI Indicators—Fertilisation Intensity

Recording fertilisation intensity using RS is central to precision farming and allows conclusions to be drawn about A-LUI. For example, the intensive use of fertilisers and pesticides promotes a homogeneous and vital vegetation pattern by increasing plant growth and yield quality. Precise nutrient monitoring includes plant traits and nutrient information, with imaging spectroscopy as a key method to determine the nutrient status of crops and soil availability quickly and non-destructively. However, there are challenges as macro- and micronutrients, stress factors, and phenological development stages have similar spectral signatures, which favours confusion at different scales.
Vegetation indices such as the Normalised Difference Vegetation Index (NDVI) quantify plant health and density and provide information on fertilisation and management practices. NDVI is often used to measure plant vigour and derive fertiliser recommendations. Li et al. [205] demonstrated UAV-based hyperspectral imaging to optimise nitrogen stress indices in maize. The Normalised Difference Red Edge Index (NDRE) more precisely determines the chlorophyll and nitrogen content of plants, which Li et al. [205] confirmed for maize. The chlorophyll index also serves as an indicator for nutrient status, whereby hyperspectral data enable an exact determination of the chlorophyll content [205]. Yin et al. [206] used ensemble learning models and Sentinel-2 data to quantify the nitrogen concentration and aboveground biomass of potato plants with a coefficient of determination R2 = 0.74. Almawazreh et al. [207] used UAV to investigate the effects of nitrogen fertilisation on the canopy temperature of agricultural crops in Southern India. Increased nitrogen applications reduced the leaf temperature of maize by 2.1 °C and finger millet by 1.3 °C under sunny conditions. Hossen et al. [208] developed an AI-based, near real-time multispectral sensor solution for drones to accurately estimate the nitrogen content in the soil.

4.5.4. Functional A-LUI Indicators—Soil Organic Carbon (SOC)

Soil organic carbon (SOC) is a key component of soil quality and plays a crucial role in the global carbon cycle [209]. Higher A-LUI (e.g., heavy fertilisation, frequent tillage) generally leads to a decrease in soil organic carbon, as ploughing, erosion, and humus decomposition mineralise carbon more quickly and release it as CO2. Furthermore, A-LUI influences plant cover and biomass production, which in turn has an impact on carbon storage in the soil. Precise mapping and monitoring of SOC is necessary to develop sustainable agricultural practices and optimise carbon storage in soils.
RS enables efficient and cost-effective monitoring of large areas, provides data from regions that are difficult to access, and allows the continuous recording of SOC dynamics with high temporal resolution [210]. Research on satellite-based SOC mapping started in the 1990s with Landsat TM data, where first correlations between spectral signatures and SOC concentrations were found [211]. These early studies showed promising results, but the spatial resolution was limited to 30 m and correlations often only reached R2 values of around 0.5, indicating high uncertainties [212]. In the 2000s, high-resolution RS data was combined with ground-based measurements to better map the spatial variability of SOC. Initial attempts to couple soil chemical properties with the Normalised Difference Vegetation Index (NDVI) method from Landsat data demonstrated the importance of vegetation cover for SOC modelling [213]. Studies show that multispectral, hyperspectral, and RADAR sensors on satellite platforms can provide crucial data for SOC mapping [214]. However, optical RS is subject to certain limitations, particularly due to cloud cover. One possible solution is to combine optical- and RADAR-based data [215]. With the introduction of the Sentinel-1 and Sentinel-2 satellites in the 2010s, SOC mapping improved significantly. Sentinel-2 provides multispectral images with a resolution of up to 10 m, while Sentinel-1 provides RADAR images that can be used regardless of weather conditions [215,216]. RADAR data, in particular SAR, has potential for SOC mapping [214,217,218], but parameters such as polarisation, band frequency, orbit, and time window significantly influence the accuracy of the models [219,220]. For example, SAR signals interact differently with vegetation layers depending on wavelength, which means that C-band and L-band systems differ in their applicability. Nevertheless, comprehensive analyses comparing different optical- and RADAR-based sentinel satellites (Sentinel-1/2/3) for SOC mapping are still rare. In recent years, deep learning algorithms and hybrid models have proven to be particularly promising. Recent studies combine optical (Sentinel-2) and RADAR-based (Sentinel-1) RS data to further improve accuracy [221]. In addition, AI-based methods such as Random Forest, Light Gradient Boosting Machine (LGBM), and neural networks have been successfully used for SOC mapping [222,223]. Mean SOC content and C:N ratio maps predicted by 100 runs of BRT in Model V at a resolution of 100 m and their corresponding standard deviation maps (Model V: all available predictors, Sentinel-1-predictors, Sentinel-2 predictors, Landsat-8 predictors, climate predictors, topography predictors) [93] (see Figure 22).
Table A4 presents a structured overview of functional indicators of A-LUI indicators, including concrete examples, the corresponding RS sensors, and the representative literature references.

5. Examples of Trait–Sensor Linkages

RS offers the ability to monitor A-LUI through proxies that reflect underlying plant and soil traits. While our framework is primarily conceptual, Table A5 provides an illustrative subset of trait–sensor linkages to demonstrate how the taxonomy can be operationalised. The examples cover biochemical, structural, phenological, and genesis traits that respond sensitively to management intensity, and they highlight three key aspects: (1) Sensor characteristics. Different traits are best captured by specific RS technologies. For example, leaf nitrogen and chlorophyll are well resolved with red-edge indices or hyperspectral sensors, while canopy structure is better observed through LiDAR metrics or SAR backscatter. Thermal sensors and radar are essential for detecting water status and irrigation practices, whereas taxonomic composition and diversity require multi-temporal optical classification. (2) Expected directionality with intensity. Intensive management practices such as fertilisation, irrigation, or high sowing density are typically associated with increased canopy greenness, higher LAI and biomass, and more frequent disturbance signals from tillage or harvesting. Conversely, long-term intensive use often leads to declining soil organic matter or reduced crop diversity. These relationships provide measurable signatures of intensity, but their interpretation must be contextualised. (3) Confounding factors. Trait–intensity relationships are not deterministic. Cultivar differences, soil fertility gradients, and climatic variability can mimic or mask management effects. For example, high chlorophyll content may reflect either fertiliser application or inherently fertile soils; frequent harvest signals may stem from double-cropping systems or from regionally specific phenologies. A key research challenge is, therefore, to separate management-driven intensity signals from background biophysical potential and land cover dynamics. This Table A5 is not intended as a complete mapping, but as a demonstration of how conceptual trait categories can be translated into implementable RS indicators. Developing a systematic and validated trait–sensor–management matrix across crop types, agroecological zones, and sensor platforms represents a crucial agenda for subsequent research. Such work will require integration with farm records, independent proxy datasets, and uncertainty quantification to ensure policy-relevant and globally comparable intensity assessments.
While Table A5 focuses on traits and their observable RS proxies, a further step is needed to explicitly link these to management practices and their broader policy relevance. Table A6 extends the trait-based perspective by integrating concrete management actions (e.g., fertiliser application, irrigation, tillage, crop rotation, field consolidation, or hedgerow management) with the traits and processes they influence, the corresponding RS observables, and the resulting A-LUI indicator categories. In addition, the Table specifies validation needs—such as ground sampling, farm records, flux tower data, or biodiversity field surveys—and highlights the potential for direct policy applications, ranging from nutrient efficiency reporting to monitoring compliance with agri-environmental schemes or biodiversity conservation targets.
Together, Table A5 and Table A6 illustrate the pathway from plant and soil traits to RS observables, and from there to validated A-LUI indicators that are directly relevant for management and policy. This integrative perspective underscores the importance of trait-based frameworks not only for scientific monitoring but also for supporting evidence-based governance of agricultural intensification.

6. Linking Management, Traits, and RS to A-LUI Indicators, Validation, and Policy

The monitoring of A-LUI requires an integrative perspective that connects management practices with biophysical responses and policy-relevant indicators. Figure 23 provides such a schematic overview, linking management inputs (e.g., fertiliser, irrigation, crop protection, tillage, crop rotations) to plant and soil traits, which serve as the central mediators between human interventions and the biophysical signals recorded by RS. These traits—biophysical, biochemical, and structural properties—can either be measured in situ or derived indirectly from RS data. RS observables and indices such as vegetation indices (NDVI, EVI), solar-induced chlorophyll fluorescence (SIF), soil moisture, or canopy structural metrics derived from LiDAR represent the quantifiable expressions of these traits.
The framework illustrates how RS-derived observables feed into the five categories of A-LUI indicators: trait, genesis, structural, taxonomic, and functional indicators. Each category addresses a distinct dimension of A-LUI, ranging from biochemical leaf properties (trait indicators) to temporal dynamics (genesis indicators), landscape configuration (structural indicators), crop and species composition (taxonomic indicators), and ecosystem processes (functional indicators). However, RS alone cannot fully capture the underlying causes of change. Hence, the integration of validation datasets—such as ground-based measurements, experimental phenotyping, and farmer-reported management data—is essential to calibrate and verify RS products.

7. From Inputs–Outputs–Impacts to A-LUI Indicators: Advancing the Framework

To better position our proposed A-LUI taxonomy against prior syntheses, we provide a cross-walk between established frameworks of LUI and our five indicator categories. Earlier frameworks have typically emphasised the three-pillar structure of inputs, outputs, and system-level impacts, with some extensions towards land use change intensity, efficiency measures, and socioeconomic drivers. While these approaches provide a solid foundation, they often remain generic and do not explicitly capture the specific opportunities and challenges of RS-based monitoring.
Our new definition (trait, genesis, structure, taxonomic, function) builds on these established dimensions but introduces several novel contributions. In particular, the structural and taxonomic categories add an explicit consideration of field geometry and crop diversity, dimensions often overlooked in previous frameworks. The genesis category uniquely accounts for temporal trajectories of land use intensity, such as crop rotations, multiple harvests, or abandonment, which are central for RS time-series analysis. Moreover, the taxonomy operationalises agricultural management metrics (e.g., irrigation, fertilisation) through RS proxies, thereby making input intensity directly observable. Finally, by providing a structured indicator appendix and a bridging analysis between challenges, framework categories, and RS/AI solutions, the taxonomy explicitly addresses operational and policy readiness. Table 1 highlights both the conceptual overlaps and the distinct advances of our framework, clarifying how the proposed taxonomy extends beyond previous reviews and offers a more operational and policy-relevant structure for monitoring agricultural land use intensity.

8. New Approaches for the Quantification and Evaluation of A-LUI Using RS

8.1. RS and AI for Recording A-LUI

The precise recording of A-LUI is central to quantifying the impact of agricultural management on ecosystems and developing sustainable strategies. Recently, RS and artificial intelligence (AI) have established themselves as key technologies for recording agricultural utilisation intensities on a large scale, promptly and objectively [224]. RS data and its time series such as Sentinel-2, Landsat-8, MODIS, or WorldView-3 provide high-resolution information on vegetation, soil surface, and hydrology, allowing numerous indicators to be derived (see Table A7), which are closely related to agronomic interventions such as fertilisation, tillage, detection of crop rotations, harvest cycles, and tillage patterns or multiple harvests per year as characteristics of A-LUI [225,226].
A critical step in interpreting these data are the integration of AI methods, in particular machine learning and deep learning, which can recognise complex, non-linear patterns in large, heterogeneous datasets. AI models such as Convolutional Neural Networks (CNNs), Support Vector Machines (SVMs), or Random Forests (RF) have been successfully used many times in the literature to quantify characteristics such as nutrient availability, plant health, or management practices [226,227,228]. For example, Castaldi et al. [229] used Sentinel-2 data to derive soil organic carbon, indicating many years of intensive use. Shi et al. [230] combined RGB images with Backpropagation Neural Networks (BPNN) to estimate nitrogen accumulation and biomass in rice fields. This allows conclusions to be drawn about fertiliser intensity and growth potential. Sahabiev et al. [231] extended these approaches by incorporating soil characteristics (e.g., organic carbon, soil texture) into ML models for the spatial prediction of nutrient distributions. A particularly relevant example in the context of utilisation intensity is the use of CNNs to detect crop cycles, which is made possible by the time series of satellite images (e.g., Sentinel-2 or MODIS). The detection of multiple harvests or intensive crop rotations is possible by analysing NDVI time histories [232,233]. In this context, Wang et al. [226] demonstrated that a combined LSTM-CNN model, trained with weather and soil data, was able to provide very precise predictions of the winter wheat harvest in China—a direct measure of output intensity. Various methods have also been established for nutrient intensity. Jaihuni et al. [234] used deep learning to estimate the spatio-temporal distribution of nitrogen, potassium, and phosphorus.
Despite these successes, challenges remain: The technical complexity of RS data processing requires specialised expertise and high-performance infrastructures [224]. High-resolution data material, such as UAV-based hyperspectral images, is often only available locally. There is a lack of standardised definitions and indicators for deriving A-LUI, which makes comparability between regions difficult [235]. In addition, many deep learning models are difficult to interpret—a problem that recent work on Explainable AI (XAI) aims to counteract.
Nevertheless, future prospects are extremely promising. New architectures such as edge cloud computing or the edge cloud continuum make it possible to process large amounts of data in a decentralised manner on sensors and drones [224]. At the same time, methods such as transfer learning or few-shot learning allow models to be adapted for new regions with little training data [226,236]. This could make globally standardised, AI-supported maps of land use intensity a reality—a valuable tool for agricultural policy, climate protection, and sustainable land use worldwide [224].

8.2. Semantic Web and Linked Open Data for the Monitoring of A-LUI

Semantic web integration (SWI) and linked open data (LOD) approaches are designed to make heterogeneous datasets interoperable, machine-readable, and reusable across institutions and platforms. For A-LUI monitoring, this is highly relevant because RS-derived indicators, in situ measurements, farm management records, and policy data often exist in separate silos and use inconsistent terminologies. Semantic technologies provide a way to bridge these gaps by assigning shared vocabularies and ontologies to different data sources [237,238]. At their core, semantic technologies such as RDF (Resource Description Framework), OWL (Web Ontology Language), and SPARQL (a query language) create a common structure that allows datasets to be linked through concepts rather than file formats. For example, different databases may store information on “crop type,” “fertilisation,” or “irrigation.” Through a shared ontology, these concepts can be harmonised, enabling automated queries and reasoning across datasets. For A-LUI monitoring, the main advantage is the ability to connect three layers of information: (1) RS indicators (e.g., crop type maps, soil moisture, vegetation indices, yield proxies). (2) Farm and management data (e.g., fertiliser application, irrigation logs, crop rotations). (3) Policy and sustainability frameworks (e.g., SDG 2.4.1 on sustainable agriculture, FAO agri-environmental indicators). This integration allows complex questions to be addressed consistently, such as “Which wheat fields in Region X received irrigation in 2023 and show high biomass in Sentinel-2 indices?” or “How do RS-derived nitrogen proxies correspond to reported fertilisation levels in regional statistics?” A practical example illustrates this potential. RS-derived crop classification maps (e.g., from Sentinel-2) can be linked to farm irrigation and fertilisation records. Using an ontology that defines shared concepts like “crop type” and “management practice”, a query could identify all irrigated maize fields with high NDVI values in a given season. The semantic web layer then produces a harmonised map or table that combines RS and farm data into a single, policy-relevant product. For practitioners, this means that information which today is scattered across agencies (satellite data at space agencies, farm logs at agricultural offices, policy indicators at statistical bureaus) could be accessed in one place, with automated links ensuring consistency. For researchers, this enhances reproducibility and data transparency, as datasets can be cited and queried through open standards. At present, such applications remain in pilot and prototyping stages, with operational deployment still limited. Therefore, we present semantic web and linked open data approaches as a future prospect for A-LUI monitoring. Their adoption could greatly enhance the integration of RS and management data, improve cross-scale comparability, and facilitate alignment with international monitoring frameworks (Figure 24).

9. Conclusions and Further Research

This review synthesised current concepts, definitions, and methodological approaches for monitoring A-LUI with a particular focus on RS- and trait-based indicators. By proposing a structured taxonomy that distinguishes trait, genesis, structural, taxonomic, and functional indicators, we aimed to bring conceptual clarity and to align RS-derived measures more explicitly with established A-LUI dimensions. The review highlights how traits can serve as a common denominator between in situ measurements and RS proxies, and how linking traits to measurable observables helps to bridge methodological and disciplinary gaps.
At the same time, our synthesis demonstrates the limitations and uncertainties inherent in RS-based monitoring of A-LUI. Cultivar-specific differences, mixed pixels, sensor saturation, and phenological variation can bias indicator derivation and interpretation. While we have outlined representative examples and conceptual solutions, systematic validation strategies remain a major research need. Future work should, therefore, prioritise long-term, multi-scale validation efforts that combine ground measurements, phenotyping infrastructures, and RS observations.
Another frontier is the development of operational frameworks that can integrate existing datasets (e.g., FAO, OECD, Eurostat) with Earth Observation-derived indicators in a transparent and standardised manner. Progress in open-source data infrastructures and community repositories provides promising entry points, but dedicated projects are needed to ensure consistency, accessibility, and long-term comparability.
Looking ahead, further research should address the following:
  • The design of multi-scale validation protocols to quantify uncertainty and improve indicator robustness.
  • The integration of hyperspectral, thermal, and radar missions with AI-based approaches for trait retrieval and intensity mapping.
  • The differentiation of management intensity signals from biophysical potential and land cover dynamics through coupled RS–model frameworks.
  • The systematic assessment of smallholder and heterogeneous landscapes, where high-resolution data and advanced image analysis are crucial.
  • The establishment of specialised studies focusing on cultivar-specific effects, phenological corrections, and management practices that cannot yet be robustly inferred from RS alone.
In conclusion, while significant progress has been made in conceptualising and operationalising RS-based A-LUI monitoring, a globally consistent and validated framework is still in development. By clarifying definitions, structuring indicator categories, and highlighting limitations, this review provides a foundation for subsequent studies to address these open challenges and to move towards an integrated, transparent, and scalable monitoring of agricultural land use intensity.

Author Contributions

A.L., F.H.: conceptualisation, methodology, writing—original draft preparation, writing—review and editing, visualisation, J.B.: supervision; P.S., J.B., M.P., T.Z., A.J.: investigation, resources, visualisation, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No data were integrated in this paper.

Acknowledgments

Our special thanks go to the Helmholtz Centre for Environmental Research—UFZ and the TERENO project, funded by the Helmholtz Association and the Federal Ministry of Education and Research, for providing the remote sensing research. The authors gratefully acknowledge the German Helmholtz Association for supporting the activities of research data science approaches and advanced data technologies. András Jung was supported by project No. TKP2021-NVA-29, which was implemented with the support of the Hungarian Ministry of Culture and Innovation from the National Fund for Research, Development, and Innovation under the TKP2021-NVA funding programme.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

A-LUIAgricultural Land-Use Intensity
AGROVOCAgricultural Vocabulary (FAO controlled vocabulary)
AIArtificial Intelligence
CHIMECopernicus Hyperspectral Imaging Mission for the Environment
EnMAPEnvironmental Mapping and Analysis Programme
ETEvapotranspiration
EUROSTATStatistical Office of the European Union
EVIEnhanced Vegetation Index
FAOFood and Agriculture Organisation of the United Nations
FLEXFluorescence Explorer
GEDIGlobal Ecosystem Dynamics Investigation
GHGGreenhouse Gas
GISGeographic Information System
GLADGlobal Land Analysis and Discovery
GLCGlobal Land Cover
GPPGross Primary Productivity
HISUIHyperspectral Imager Suite
HyspIRIHyperspectral Infrared Imager
IACSIntegrated Administration and Control System
IPCCIntergovernmental Panel on Climate Change
LAILeaf Area Index
LandsatLand Satellite (USGS/NASA Earth observation programme)
LiDARLight Detection and Ranging
LUCASLand Use/Cover Area Frame Survey
MLMachine Learning
MODISModerate Resolution Imaging Spectroradiometer
NDVINormalised Difference Vegetation Index
OECDOrganisation for Economic Co-operation and Development
PlanetScopeHigh-resolution satellite constellation operated by Planet Labs
PRISMAPRecursore IperSpettrale della Missione Applicativa
RSRemote Sensing
SARSynthetic Aperture Radar
SDGSustainable Development Goal
Sentinel-1C-band Synthetic Aperture Radar mission (Copernicus)
Sentinel-2Multispectral optical imaging mission (Copernicus)
Sentinel-3Ocean and land monitoring mission (Copernicus)
Sentinel-5PTropospheric monitoring mission (Copernicus)
SHALOMSpaceborne Hyperspectral Applicative Land and Ocean Mission
SIFSolar-Induced Fluorescence
SOCSoil Organic Carbon
UAVUnmanned Aerial Vehicle
World BankWorld Bank (International Financial Institution)

Appendix A

Figure A1. Monitoring the five characteristics of A-LUI using RS. These are traits of A-LUI, genesis traits of A-LUI, structural traits of A-LUI, taxonomic traits of A-LUI, functional traits of A-LUI with examples.
Figure A1. Monitoring the five characteristics of A-LUI using RS. These are traits of A-LUI, genesis traits of A-LUI, structural traits of A-LUI, taxonomic traits of A-LUI, functional traits of A-LUI with examples.
Agriculture 15 02233 g0a1
Table A1. Geographical area of monitoring, temporal availability of indicators, link, and selected examples of indicators for measuring and monitoring A-LUI; carried out by the FAO, OECD, World Bank and EUROSTAT.
Table A1. Geographical area of monitoring, temporal availability of indicators, link, and selected examples of indicators for measuring and monitoring A-LUI; carried out by the FAO, OECD, World Bank and EUROSTAT.
FAOOECDWorld BankEUROSTAT
Geographical area of monitoring
  • Worldwide coverage, with a special focus on developing countries
  • Primarily OECD member countries, focus on highly developed industrialised nations
  • Developing countries and emerging markets
  • European Union (EU) and some enlargement countries
Time availability of the indicators
  • Indicators of land use intensity have been available since the 1960s,
  • Increased surveillance since the 1990s
  • Data and analyses on land use intensity since the 1980s,
  • Regular reports since the early 2000s.
  • Data on land use intensity since the 1990s,
  • Comprehensive database (WDI) since the 2000s.
  • Harmonised data on agriculture and land use since the 1990s,
  • Regular (every three to five years) surveys since the 1990s
Link
Indicators (selective examples)
IndicatorFAOOECDWorld BankEUROSTAT
Agricultural areaTotal area for agriculture (arable land, permanent grassland, permanent crops)Agricultural land, including arable land, permanent crops, and pasturesAgricultural land (sq. km)Utilised agricultural area (UAA)
Arable landLand for crops, including repeatedly cultivated soils and fallow landArable land, including temporary crops and fallow landArable land (hectares)Arable land
Permanent grasslandLand for perennial grasses and forage plantsPermanent pastures and meadowsPermanent meadows and pastures (hectares)Permanent grassland
Permanent cropsLand for perennial crops such as fruit trees and vineyardsPermanent crops, such as orchards and vineyardsPermanent crops (hectares)Permanent crops
Harvest yieldsAmount of crop per unit areaCrop yields, measured by specific crop outputs per hectareCereal yield (kg per hectare)Crop production per unit area
Use of fertilisersAmount of fertiliser per hectareFertiliser consumption (kg per hectare of arable land)Fertiliser consumption (kg per hectare of arable land)Consumption of fertilisers per unit area of agricultural land
Pesticide useAmount of pesticides per hectarePesticide sales and usagePesticide consumption (kg per hectare of arable land)Pesticide sales and consumption
Irrigated areaProportion of artificially irrigated agricultural landArea equipped for irrigation (hectares)Irrigated land (% of total agricultural land)Irrigated area
Machine inventoryNumber and type of machines per unit areaAgricultural machinery, such as tractors per hectareAgricultural machinery (tractors per 100 sq. km of arable land)Number of tractors and other agricultural machinery per unit area of agricultural land
Labour inputLabour hours per unit areaLabour input in agriculture, measured by hours worked per hectareEmployment in agriculture (% of total employment)Labour force in agriculture
Livestock densityNumber of animals per unit area of pasturelandLivestock density, measured as livestock units per hectare of pasture landLivestock production indexLivestock density per unit area of pasture land
Carbon sequestration in the soilAmount of carbon sequestered in the soilSoil organic carbon contentSoil organic carbon contentSoil organic carbon content
Ground coverType and extent of ground coverLand cover types and changesLand cover (% of land area)Land cover and land use
Erosion riskRisk of soil erosion due to water or windSoil erosion ratesSoil erosion ratesSoil erosion and degradation risk
BiodiversityDiversity of plant and animal species on farmland land (e.g., Farmland birds, pollinators, butterflies)Farmland biodiversity indices (e.g., Farmland birds, pollinators, butterflies)Agricultural biodiversity indices (e.g., Farmland birds, pollinators, butterflies)Biodiversity indicators in agricultural landscapes (e.g., Farmland birds, pollinators, butterflies)
Water consumption in agricultureAmount of water used for irrigationAgricultural water withdrawalAgricultural water withdrawal (% of total water withdrawal)Water use in agriculture
Agricultural production per unit of inputEfficiency of the means of production in agricultureTotal factor productivity in agricultureAgricultural value added per workerOutput per hectare of agricultural land
Energy consumption in agricultureEnergy consumption in agricultureEnergy use in agricultureEnergy use in agricultureEnergy consumption in agriculture
Sustainability indicatorsSustainability of agricultural practicesSustainable agriculture practices indicatorsSustainable land management indicatorsSustainable farming practices
Climate impact of agricultureGreenhouse gas emissions from agricultureGreenhouse gas emissions from agricultureAgricultural methane emissions (kt of CO2 equivalent)Greenhouse gas emissions from agriculture
Nutrient balance in the soilBalance of nitrogen and phosphorus in the soilNitrogen and phosphorus balanceSoil nutrient balanceNutrient balance in agricultural soils
BioproductivityProductivity of biological systems on agricultural landBiological productivity of agricultural systemsAgricultural productivity indexesBiological productivity of agricultural lands
Plant protection measuresMeasures to combat pests and diseasesPest and disease control practicesPest and disease control indicatorsPlant protection measures and their impact
Energy efficiency in agricultureEfficiency of energy consumption in agricultureEnergy efficiency in agricultural practicesEnergy productivity in agricultureEnergy efficiency indicators in farming
Utilisation of genetic resourcesUtilisation and conservation of genetic resources in agricultureUse and conservation of genetic resourcesGenetic resource management indicatorsConservation and use of agricultural genetic resources
Landscape diversityDiversity of landscapes and agroecosystemsLandscape diversity and heterogeneityLandscape diversity indicatorsLandscape heterogeneity and diversity in agricultural areas
Soil compactionDegree of soil compaction caused by agricultural machinerySoil compaction indicatorsSoil compaction riskSoil compaction due to agricultural practices
Waste management in agricultureHandling agricultural wasteAgricultural waste management practicesWaste management in agricultureManagement and recycling of agricultural waste
Soil moistureMoisture content of the soilSoil moisture levelsSoil moisture content indicatorsSoil moisture monitoring in agricultural lands
Landscape fragmentationFragmentation of natural and agricultural landscapesLandscape fragmentation and its impact on agricultureLandscape fragmentation indexesImpact of landscape fragmentation on agriculture
Sustainable land use practicesSpreading sustainable agricultural practicesAdoption of sustainable agricultural practicesSustainable land management practicesImplementation of sustainable farming practices
Water utilisation efficiencyEfficiency of water utilisation in agricultureWater use efficiency in agricultural practicesAgricultural water productivityWater use efficiency in irrigated agriculture
Agroecological indicatorsIndicators for the assessment of agroecological systemsAgroecological assessment indicatorsAgroecological practicesAssessment of agroecological systems
Erosion due to windLoss of topsoil due to wind erosionWind erosion ratesWind erosion indicatorsImpact of wind erosion on agricultural land
Soil fertilityLevel of soil fertility and its changesSoil fertility levelsSoil fertility indicatorsChanges in soil fertility
Land use changesChanges in the utilisation of agricultural landChanges in agricultural land useLand use change indicatorsAgricultural land use changes
Irrigation efficiencyEfficiency of irrigation methodsIrrigation efficiencyEfficiency of irrigation systemsEfficiency of water use in irrigation systems
Climate adaptation measuresMeasures to adapt to climate changeClimate adaptation practices in agricultureClimate resilience indicatorsImplementation of climate adaptation measures in agriculture
Resource utilisation efficiencyEfficient use of natural resourcesResource use efficiency in agricultureResource productivity indicatorsEfficiency of resource use in agriculture
Soil acidificationDegree of soil acidification and its causesSoil acidification levelsSoil pH indicatorsImpact of acidification on agricultural soils
Soil salinisationLevel of soil salinisation and its effectsSoil salinisation ratesSoil salinity indicatorsEffects of salinisation on agricultural productivity
Utilisation of renewable energiesShare of renewable energies in agricultureRenewable energy use in agricultural practicesShare of renewable energy in agricultureUse of renewable energy sources in farming
Environmentally friendly cultivation methodsSpreading environmentally friendly cultivation methodsAdoption of eco-friendly farming practicesEco-friendly agricultural practicesImplementation of environmentally friendly farming methods
Economic sustainabilityEconomic viability of farmsEconomic sustainability of agricultural holdingsEconomic viability indicatorsEconomic sustainability of farms
Social sustainabilitySocial aspects of agricultural practiceSocial sustainability in agricultureSocial indicators in rural areasSocial impacts of agricultural practices
Productivity per unit areaProductivity of agricultural landLand productivity indicatorsProductivity of agricultural landOutput per unit of agricultural area
Water quality indicatorsImpact of agriculture on water qualityImpact of agriculture on water qualityWater quality in agricultural areasEffects of agricultural runoff on water quality
Infrastructure for agricultureAvailability and quality of agricultural infrastructureAgricultural infrastructure developmentInfrastructure investment in agricultureQuality and accessibility of agricultural infrastructure
Innovation in agricultureImplementation of new technologies and processesAgricultural innovation and technology adoptionInnovation indicators in agricultureAdoption of new agricultural techn
Table A2. High spatial resolution satellite missions, sensor/type, spatial resolution, spectral bands/type, availability, launch date and operator.
Table A2. High spatial resolution satellite missions, sensor/type, spatial resolution, spectral bands/type, availability, launch date and operator.
Satellite/MissionSensor/TypeSpatial
Resolution
Spectral Bands/Sensor Type AvailabilityStart DateOperator of the Satellite Mission
WorldView-3Visible (PAN+MS+SWIR)0.31 m (PAN), 1.24 m (MS)Panchromatic
Multispectral SWIR
Commercial2014Maxar
WorldView-2Optically0.46 m (PAN), 1.84 m (MS)Panchromatic
Multispectral
Commercial2009Maxar
GeoEye-1Optically0.41 m (PAN), 1.65 m (MS)Panchromatic
Multispectral
Commercial2008Maxar
Pleiades NeoOptically0.3 m (PAN), 1.2 m (MS)Panchromatic
Multispectral
Commercial2021+Airbus
Pleiades 1A/1BOptically0.5 m (PAN), 2.0 m (MS)Panchromatic
Multispectral
Commercial2011/2012Airbus
SkySatOptically + Video0.5–0.8 m (PAN), 1–2 m (MS)RGB, NIR, VideoCommercial2013+Planet
BJ-3B (SuperView-2)Optically0.3 m (PAN), 1.2 m (MS)Panchromatic
Multispectral
Commercial202221AT (China)
Capella SpaceRADAR (X-Band SAR)0.3–0.5 m (Spotlight)SAR Commercial2018+Capella Space (USA)
ICEYERADAR (X-Band SAR)0.25–1 mSARCommercial2018+ICEYE (Finland)
TerraSAR-XRADAR (X-Band SAR)bis 1 m (Spotlight-Modus)SARCommercial/Scientifically free2007DLR/Airbus
PAZRADAR (SAR)1 mSAR (X-Band)Commercial2018Hisdesat (Spain)
Sentinel-1A/BRADAR (C-Band SAR)10 mSARFreely available2014/2016ESA/Copernicus
Drohnen/UAVOptically + Multispectral<0.1 mRGB, Multispectral,
Hyperspectral, LiDAR
Own operation User-based
Aerial photosOptically0.20 cmOrthophotos (DOP)
True Orthophotos,
RGB, CIR
Commercial/Authorities and partly scientific free Federal states, Federal Agency for Cartography and Geodesy
Table A3. Linking key challenges of monitoring agricultural land use intensity (A-LUI) with RS to the proposed indicator framework categories and potential RS/AI-based solutions.
Table A3. Linking key challenges of monitoring agricultural land use intensity (A-LUI) with RS to the proposed indicator framework categories and potential RS/AI-based solutions.
ChallengeRelevant Framework CategoryPossible RS/AI SolutionExample Application
Distinguishing intensive vs. extensive cultivation (e.g., organic vs. conventional)Trait indicatorsHyperspectral indices (red-edge, SIF) combined with AI crop classificationSeparation of organic vs. conventional wheat fields using Sentinel-2 red-edge indices
Seasonal dynamics and multiple harvestsGenesis indicatorsMulti-temporal analysis (Sentinel-1/2, SAR–optical fusion); AI-based phenology detectionIdentification of double-cropping systems in India
Irrigation and water managementFunctional indicatorsRadar-derived soil moisture (Sentinel-1), thermal RS for evapotranspiration, AI separation of natural vs. managed water stressMapping irrigation events in Mediterranean orchards
Fertiliser and pesticide application (not directly visible in RS)Trait and functional indicatorsIndirect proxies: leaf N content, chlorophyll indices, stress detection; ML calibration with in situ recordsEstimating nitrogen application in maize with UAV hyperspectral imaging
Small-scale heterogeneous fieldsStructural indicatorsHigh-resolution UAV/Planet imagery; OBIA; deep learning for parcel boundary delineationSmallholder mapping in Sub-Saharan Africa using PlanetScope + CNN
Agroforestry and mixed croppingTaxonomic indicatorsHyperspectral UAV imaging and AI spectral unmixingDifferentiating coffee under shade trees in agroforestry systems
Limited spectral resolution of standard satellitesTrait and functional indicatorsIntegration of hyperspectral missions (EnMAP, PRISMA, CHIME); AI-based spectral downscalingImproved stress detection in crops using EnMAP data
Climate and topographic confounding effectsGenesis & Functional indicatorsAI domain adaptation, topographic correction, normalisation with weather/soil dataAdjusting RS-based yield intensity estimates in mountainous regions
Table A4. Various indicators for measuring A-LUI that can be detected using RS. Here is a comprehensive table summarising the various indicators for measuring A-LUI and landscape structure that can be measured using RS.
Table A4. Various indicators for measuring A-LUI that can be detected using RS. Here is a comprehensive table summarising the various indicators for measuring A-LUI and landscape structure that can be measured using RS.
IndicatorsSatellitesReferences
Trait diversity of A-LUI
Chlorophyll-a/b Content
Leaf chlorophyll content (LCC)
Chlorophyllgehalt (Cab)
Canopy Chlorophyll Content (CCC)
Carotinoide, anthocyanin
Anthocyanin reflectance index (ARI)
Carotenoid reflectance index (CRI)
Sentinel-1 1, Sentinel-2 1, Landsat 8 1, CRIME 1,
EnMAP 1, Airborne hyperspectral CASI 2, Airborne Visible/Infrared Imaging Spectrometer AVIRIS 2, Airborne HyMap 2, UAV-(HSP,MSP) 3, Handheld portable hyperspectral camera (Specim IQ) ASD 4, Laboratory spectroscopy 5
[86,96,102,105,118,239,240,241,242,243,244,245,246,247,248,249,250,251]
Foliar Nitrogen, Phosphorus, Potassium—NPK UAV (LiDAR, MSP) 3, SVC HR-1024i spectrometer ASD 4[86,252,253]
Solar-induced chlorophyll fluorescence (SIF),
Photosynthesis activity
Sentinel-3 1, GOSIF data 1, AS-SpecFOM (ground-based) 6, FluoSpec2 system (ground-based) [72,107,254,255,256,257]
Leaf nitrogen content (LNC)
Nitrogen use efficiency,
Nitrogen nutrition index
Sentinel-2 1, CRIME 1, PRISMA 1,Airborne micro-hyperspec NIR-100 camera 2, UAV[86,96,98,119,120,258,259]
Plant water content
Leaf water content
Plant water stress
Cropland water-use efficiency
Crop Water Productivity
GLASS 1, Landsat 1, Sentinel-2 1,
UAV (MSP, HSP) 3, mmWave RADAR (Tower) 6, Cropland ecosystem flux sites 6, Local TIR Sensor 6,
[260,261,262,263,264,265,266,267,268]
Land Surface Temperature
Crop surface temperature
Landsat 1, High Spatio-Temporal Resolution Land Surface Temperature Monitoring (LSTM) Mission 1,
UAV (TIR, RGB, MSP) 3
[264,269,270,271,272,273]
Evapotranspriration (ET)
Crop evapotranspiration (ETc)
MODIS 1, DEIMOS-1 is a commercial tasking EO satellite 1, Landsat 1, Sentinel-2 1, SuperDove satellites (PlanetScope) 1, UAV-(RGB, MSP, TIR) 3[274,275,276,277,278,279,280,281,282]
Soil moistureMODIS-Terra 1, Landsat 1, AMSR-2 1, AMSR-E 1, NISAR 1, Sentinel-1 1, Sentinel-2 1, SMAP 1, Airborne hyperspectral (DAIS) 2, Airborne hyperspectral (AISA Eagle, Hawk) 2[283,284,285,286,287,288,289,290,291]
Irrigation
Irrigation Efficiency
Water Productivity and Efficiency
Irrigation patterns
Water-Ferilizer use efficency
Water Stress
Soil Water Deficit
Soil water stress
MODIS 1, Landsat 1, Sentinel-2 1, UAV (MSP) 3,
ASD 4,
[292,293,294,295,296,297,298,299,300,301,302,303]
LAI (Leaf Area Index) MODIS 1, Landsat 1, Sentinel-2 1, UAV-(HSP, TIR, LiDAR) 3, Ocean Optics USB2000 (Tower) 6[247,248,304,305,306]
Genesis Trait Diversity of A-LUI
Subsurface drainage systems,
Drainage density
RADAR (SAR) 1, Landsat 1, Senitnel-2 1, Airborne LiDAR 2, Airborne data 2, UAV–RGB, CIR, TIR 3[132,133,134,135,136,307,308]
Terrace mappingLandsat 1, Sentinel-1 1, Sentinel-2 1, GF-2 satellite image 1, WorldView-1 1, WorldView-3 1, Airborne LiDAR 2, UAV-LiDAR 3[87,137,138,143,144,145]
AllmendenAirborne LiDAR 3[146,147]
DeforestationMODIS 1, ALOS PALSAR data 1, RADARSAT-2 1, Landsat 1, Sentinel-1 1, Sentinel-2 1, UAV (RGB, NIR, IRT) 3[150,151,152,153,154,309,310,311,312]
Polder and single-polder systemsGoogle Earth RS data 1, Corona spy satellite imagery 1[313,314]
DEM (Digital Elevation Model)
DSM (Digital Surface Model)
SRTM 1, TerraSAR-X 1, TanDEM-X 1, Sentinel-1 1, Sentinel-3 1, ALOS-2 PALSAR-2 1, ALOS PRISM 1, Terra ASTER 1, ICESat GLAS 1, Airborne LiDAR 2, UAV (SAR, RGB) 3[61,315,316,317,318,319,320,321,322,323,324,325,326,327]
Soil Topography
Farmland microtopography feature
Landsat 1, Sentinel-1 1, Sentinel-2 1, CORONA KH-4B 1, Gaofen-7 satellite 1, Airborne LiDAR 2[171,328,329,330,331,332]
Soil metagenomics dataUAV (MSP, LiDAR) 3[333]
Structural traits of A-LUI
Soil, crop vegetation composition and configuration (e.g., patch size, distribution
Field size, Interspersion and Juxtaposition Index, Proximity Index,
Edge Density, Edge Contrast Index, Contagion Index, Core Area Index, Shape Index, Cropland Extent, Fragmentation, Homogeneity, Isolation, land use intensity patterns, Canopy structure
Farmland Boundary Extraction,
Cropland extent, Cropland area, Harvested Area Fraction, Structural Connectivity Index,
Vegetation Coherence Index, Crop Richness, Crop Evenness, Crop Simpson’s Diversity Index, Fractal Dimension Index, Entropy Index, Clumping Index,
Grassland plant species diversity
Plant density
MODIS 1, Landsat 1, Spot 1, Sentinel-2 1, WorldView-2/-3 1, QuickBird 1, Pleiades 1, GeoEye 1, GF-2 1, RapidEye 1, PlanetScope 1, Airborne Hyperspectral AVIRIS and HYDICE 2, Airborne data 2, UAV (RGB, MSP, HSP) 3[31,33,67,88,161,162,258,334,335,336,337,338,339,340,341,342,343,344,345,346,347,348,349]
Vertical Vegetation Structure,
Vegetation height, Plant height
3D-structures, 3D mapping
GEDI LiDAR 1, ICESat-2 1,
UAV (RGB, LiDAR)
Phenotyping robot “MARS-PhenoBot” 6, 6-DOT robot 6, RGB-Camera 6, Terrestrial LiDAR 6
[350,351,352,353,354,355]
Surface roughness
Canopy roughness
Sentinel-1 1, MODIS 1, UAV (RGB) 3[167,168,169]
Spektraler Heterogenität,
Rao’s Q diversity index,
Plant Species Richness
Spatiotemporal variability
MODIS 1, Landsat 1, Sentinel-2 1[165,166,356]
Homogeneity Index,
Grassland Homogeneity Index
Crop homogeneity
Sentinel-1 1, Sentinel-2 1, GF-2 1[357,358,359]
Soil Roughness,
Soil texture,
Farmland microtopography
Landsat 1, Sentinel-1 1, Sentinel-2 1, AHSI/ZY1-02D satellite 1, SRTM 1, Airborne LiDAR 2, ASD Handspectometer 4, Smartphone-captured digital images 6[171,172,332,360,361,362,363,364,365,366,367,368]
Taxonomic A-LUI
Cropping patterns
(single cropping, multiple cropping, sequential cropping, inter-cropping)
MODIS 1, Spot 1, Landsat 1, Sentinel-1 1, Sentinel-2 1, IRS 1, WiFS 1, Airborne AVIRIS 2, RADARSAT-2 1, Airborne LiDAR 2[155,173,174,175,176,177,178,179,180,181,182,183,186,369,370]
Crop classification,
Crop type classification
Crop type mapping
MODIS 1, Landsat 1, Sentinel-1 1, Sentinel-2 1, Sentinel-3 1, Airborne AVIRIS 2, UAV (HSP) 3[90,142,156,371,372,373,374,375,376,377]
Classification of grassland community typesLandsat 1, Sentinel-1 1, Sentinel-2 1[378,379,380]
Cropping frequency (single cropping/double cropping/triple cropping)
Crop rotation
Multi-cropping frequency (MCF)
Cropping intensity
Cropping intensity index
Change Detection crops
MODIS 1, Gaofen-1 1, GF-1 1, Landsat 1, Sentinel-1 1, Sentinel-2 1[174,187,341,370,381,382,383,384,385,386,387,388,389,390]
Crop residue cover mappingLandsat 1, Sentinel-2 1, Google Earth Engine 1, UAV 3,
FieldSpec Pro 4, Photo analysis surveys 6
[391,392,393,394,395,396]
Crop burning residueMODIS 1, AVHRR 1, LISS-III 1, LISS-IV 1, UAV 3 [397,398,399]
Classification between
cultivated and fallow fields
MODIS 1, Landsat 1, Sentinel-2 1[369,381,400,401,402]
Organic, conventional farming
Organic and non-organic farming
Landsat 1, Spot 1, Sentinel-2 1, KOMPSAT-2 1, WorldView-2 1, UAV (RGB) 3, Hyperspectral ASD 4 [403,404,405,406]
Phenotyping,
Phenology,
Phenology-Stadien (BBCH-Scale)
UAV (RGB, MSP, HSP, TIR, LiDAR) 3, UAV (RGB, VIS, NIR, TIR, LiDAR) 3, Labour-Hyperspectral–AISA-EAGLE 5[244,304,327,407,408,409,410,411,412]
Crop growth duration (GDa),MODIS 1, Landsat 1, Gaofen-1 1, Sentinel-2 1, RapidEye 1, UAV (SAR) 2[387,413,414,415,416]
Hedgerow map classifications,
Hedgerows and field margins
TerraSAR-X 1, Spot 1, IKONOS 1, Airborne MSP 2, Aerial photographs 2, UAV (RGB, MSP) 3[89,417,418,419,420,421]
Flower strip mapping
Flower Mapping
Airborne Hyperspectral (HySPEX, RGB, TIR) 2, Airborne Hyperspectral (AISA-Eagle) 2, Airborne MSP 2, UAV (MSP, HSP) 3[421,422,423,424,425,426]
Buffer Zone Efficiency
Agricultural Pesticides Drift zones
Landsat 1, Sentinel-2 1[427]
Classification of agroforestry systemsRapidEye 1, PlantetScope 1, LISS IV 1, Sentinel-2 1[428,429,430,431,432]
Plastic-covered greenhouses
Plasticulture detection
Plastic greenhouses (PGs) and Plastic-mulched farmland (PMF)
Landsat 1, Sentinel-1 1, Sentinel-2 1, GF-2[433,434,435,436,437]
Crop yield predictions
Grain Yield,
Protein estimation
MODIS 1, Landsat 1, Sentinel-2 1, UAV–(MSP, HSP) 3[258,438,439,440,441,442,443,444,445,446,447]
Hop cultivation classificationUAV (MSP) 3, Mobile phone camera 6[448,449]
Functional traits of A-LUI
Crop biomass,
Aboveground biomass (AGB),
Relative biomass potential (rel. BMP)
MODIS 1, Landsat 1, Sentinel-1 1, Sentinel-2 1, PlanetScope 1, UAV (MSP, RGB) 3, Smartphone 6[191,192,193,194,195,196,197,293,450]
Plant Nitrogen Concentration (PNC)
Leaf Nitrogen Content
Fertilisation Gradient
Sentinel-2 1, UAV (MSP, TIR) 3[102,205,206,207,451,452,453]
Soil organic carbon (SOC)
Soil organic matter (SOM)
ALOS-2 1, PALSAR-2 1, Landsat 1, Spot 4/5 1, GF-1 1, RADAR (PLAS) 1, Sentinel-1 1, Sentinel-2 1, Sentinel-3 1, Airborne hyperspectral (DAIS) 2, Airborne hyperspectral (AISA Eagle, Hawk) 2, Hyperspectral APEX 2, UAV (SAR) 3, VIS–NIR spectroscopy (Field) 1,[93,210,215,216,221,223,291,454,455,456,457,458,459,460,461,462,463,464,465,466]
Clay contentLandsat 1, Aster 1, Sentinel-2 1, Airborne hyperspectral (AISA Eagle, Hawk) 2[368,467,468,469,470,471,472,473]
Soil total nitrogen (TN)
N-Monitoring
Total soil nitrogen (TSN)
Nutritional Status
Soil Total Nitrogen
Soil Nutrients Contents
Sentinel-1 1, Sentinel-2 1, GF-1 1, UAV (HSP, MSP, TIR) 3, ASD (Field) 4[208,215,461,472,474,475,476,477,478,479,480]
C:N ratio soilLandsat 1, Sentinel-1 1, Sentinel-2 1, Sentinel-3 1[93,460,481,482,483,484]
Carbon use efficiency (CUE)MODIS 1, Landsat 1, Sentinel-2 1[485,486,487,488]
Silt contentGF-1 1, Airborne hyperspectral (AISA Eagle, Hawk) 2,[368,489]
Sand contentLandsat 1, Sentinel-2 1, Aster 1, GF-1 1, Planet/NICFI 1, Airborne hyperspectral (AISA Eagle, Hawk) 2[368,473,489,490,491,492,493]
Potassium contentPRISMA 1, UAV (MSP) 3 [476,477]
Phosphorus content (P)MODIS 1, Landsat 1, Sentinel-2 1, PRISMA 1, UAV (MSP, LiDAR) 3, ASD 4[333,476,477,478,479,494]
Pestizide, Herbizide, Fungizide
Pest management
Sentinel-2 1, UAV 3, Local hyperspectral camera 6, ASD—LeafSpec hyperspectral images 4[198,199,200,201,202,203,495,496]
Plant Disease Detection,
Crop vegetation health
Plant health
Sentinel-1 1, Sentinel-2 1, UAV (RGB, MSP, VIS, NIR, TIR, LiDAR) 3, ASD FieldSpec Pro FR 4[92,105,404,409,497,498,499,500,501,502,503,504,505,506,507,508]
CSR-Plant Strategy Types
Plant functional groups (PFGs)
Ellenberg Indicator Species
Landsat 1, Sentinel-2 1, Airborne hyperspectral data (AISA dual) 2, Airborne AISA Fenix 2, Airborne imaging spectrometer APEX 2, Airborne hyperspectral HySpex 2[509,510,511,512,513,514,515,516]
Gross Primary Production (GPP)
Dynamic of carbon emissions,
Carbon Fluxes
MODIS 1, Meris 1, Landsat 1, Sentinel-1 1, Sentinel-2 1, Sentinel-3 1, Hyperspectral Ocean Optics USB2000 (Tower) 6, LEDAPS-Aerosol Robotic Network (AERONET) 6[246,485,517,518,519,520,521,522,523,524]
Cropland NPP MODIS 1, Landsat 1, UAV (MSP) 3[306,347,485,525,526,527,528,529,530]
HANPP (Human Appropriation of Net Primary Production)MODIS 1, Landsat 1, Sentinel-2 1[531,532,533,534,535]
Water use efficiency (WUE)MODIS 1, Landsat 1, Sentinel-1 1, Sentinel-2 1[485,536,537,538,539,540]
Yield and QualityLandsat 1, Sentinel-1 1, Sentinel-2 1, UAV (MSP) 3[196,523,541,542,543,544,545,546,547,548]
Harvest IndexMODIS 1, HJ-1 satellite 1, Sentinel-2 1, UAV (HSP) 3, FieldSpec HandHeld Spectroradiometer (ASD) 4[549,550,551,552]
Soil quality index (SQI)Landsat 1, Sentinel-2 1, Airborne hyperspectral (AISA) 2[330,553,554]
Soil productivity potentialMODIS 1, Landsat 1, Sentinel-2 1, ASD FieldSpec 4[302,472,555,556,557]
Soil CrustKOMPSAT-2 satellite 1, Airborne hyperspectral (DAIS) 2, Airborne hyperspectral (AISA Eagle, Hawk) 2, UAV (RGB, MSP, HSP) 3, ASD Fieldspec 4[291,558,559,560,561,562,563,564]
Soil infiltrationAirborne hyperspectral (DAIS) 2, Airborne hyperspectral (AISA Eagle, Hawk) 2, Airborne CASI-1500 2, SASI-600 2, Airborne TASI-600 hyperspectral sensors 2, UAV (HSP, Cubert UHD-185) 3[291,565,566]
Soil pH valuePALSAR-1/2 1, SRTM 1, Landsat 1, PlantetScope 1, Sentinel-1 1, Sentinel-2 1, UAV (MSP) 3, ASD FieldSpec 4[290,361,547,567,568,569,570,571,572,573,574,575,576,577]
Soil salinity
Soil salinisation
Landsat 1, RADAR 1, Airborne LiDAR 2, HJ-1 Hyperspectral Imager Data 2[290,578,579,580,581,582,583,584,585]
Land degradation,
Soil degradation,
Soil erosion
Desertification
Landsat 1, SRTM 1, Sentinel-1 1, Sentinel-2 1, RapidEye 1, Airborne hyperspectral (DAIS) 2, Airborne hyperspectral (AISA Eagle, Hawk) 2, UAV (RGB) 3[291,586,587,588,589,590,591]
Soil compaction
Soil Compaction Index
Soil aggregation
Soil penetration resistance
Landsat 1, GoogleEarth aerial imagery 1, Sentinel-2 1, RapidEye 1, Airborne hyperspectral (CASI) 2, UAV (RGB, SAR, LiDAR, MSP, TIR) 3[587,592,593,594,595,596,597,598,599]
Cattle intensification,
Spatial distribution of cattle
Sentinel-1 1, Sentinel-2 1[600]
Grassland use intensity
Grassland management intensity
Landsat 1, Sentinel-1 1, Sentinel-2 1, RapidEye 1, [91,188,601,602,603,604,605]
Grassland fireMODIS 1, Sentinel-1 1, Sentinel-2 1, GF-6 WFV 1, UAV 3[606,607,608,609,610]
Grassland cut detectionSAR 1, Sentinel-1 1, Sentinel-2 1[611,612,613]
Different Water quality indicatorsAll RS Sensors with all RS characteristics (MSP, HSP, TIR, RADAR, LiDAR)[63]
The sensor is used on the RS platform: 1—spaceborne RS platforms, 2—airborne RS platform, 3—UAV, 4—Handheld portable hyperspectral camera (Specim IQ) ASD, 5—Laboratory spectroscopy, 6—Tower, Smartphone, ground-based.
Table A5. Illustrative examples linking plant and soil traits to RS observables, intensity signals, and key confounders.
Table A5. Illustrative examples linking plant and soil traits to RS observables, intensity signals, and key confounders.
Trait/ProcessRS Sensor/ModalityDirectionality
with Intensity
Key Confounders
(Non-Management)
Leaf N/chlorophyll contentRed-edge indices (Sentinel-2), hyperspectral (EnMAP, CHIME), solar-induced fluorescence (FLEX)↑ with higher fertilisation and improved managementCultivar-specific pigment traits; background soil reflectance; cloud/shadow effects
Canopy structure (LAI, height, biomass)Multispectral VIs (NDVI, EVI), LiDAR metrics (GEDI, UAV-LiDAR), SAR backscatter (Sentinel-1)↑ with higher input intensity, dense sowing, irrigationNatural soil fertility; precipitation regime; lodging events
Phenology (timing, cropping frequency)Time series (Sentinel-1 coherence for tillage/harvest; Sentinel-2 optical indices; PlanetScope)More frequent harvests or longer growing season → ↑ intensityClimate-driven shifts in growing season; interannual weather variability
Root traits (water/nutrient uptake)Thermal (ET proxies), SAR soil moisture (Sentinel-1), hyperspectral water stress proxiesIntensive irrigation/fertilisation → ↑ water use efficiency or altered root activitySoil texture; groundwater availability; drought stress independent of management
Canopy temperature/water statusThermal sensors (ECOSTRESS, UAV-TIR), ET modelling with optical+thermal fusion↓ canopy temperature and ↑ ET with irrigation intensityHeat waves, VPD variability, soil hydraulic properties
Structural diversity (field size, edges, hedgerows)High-res optical (PlanetScope, UAV), LiDAR for vertical structure, OBIA↑ intensity often linked to larger fields, reduced edge densityHistorical land consolidation, topography, land tenure
Crop type diversity (taxonomic composition)Multi-temporal Sentinel-2/Landsat, hyperspectral UAV, classification algorithms↑ intensity often → ↓ diversity, monocroppingRegional crop rotations, policy incentives, cultural practices
Soil organic matter/C:N ratioHyperspectral reflectance (VNIR-SWIR), SAR + optical fusion, regression models↓ SOM with long-term intensive use, ↑ mineral N inputs → altered C:NParent material, drainage, climate-driven decomposition
Harvest/tillage eventsSAR coherence (Sentinel-1), time-series change detection, UAV imagery↑ intensity = more frequent disturbance signals per seasonWeather-induced soil roughness, cloud cover gaps
Pest/disease stress signalsHyperspectral indices (red-edge, PRI), fluorescence (SIF), UAV multispectralIntensive management may ↓ visible stress due to pesticide controlPathogen pressure, local outbreak dynamics, cultivar resistance
Table A6. Integrative framework linking management practices, traits, RS proxies, A-LUI indicator categories, validation needs, and policy relevance. The table highlights how RS-derived observables can serve as bridges between field management, biophysical processes, and policy-relevant indicators of agricultural land use intensity.
Table A6. Integrative framework linking management practices, traits, RS proxies, A-LUI indicator categories, validation needs, and policy relevance. The table highlights how RS-derived observables can serve as bridges between field management, biophysical processes, and policy-relevant indicators of agricultural land use intensity.
Management PracticeTrait/Process AffectedRS Proxy/ObservableA-LUI Indicator CategoryValidation NeedsPolicy Relevance
Fertiliser applicationLeaf nitrogen, canopy chlorophyllRed-edge indices (Sentinel-2), hyperspectral retrievalsTrait/FunctionalGround sampling, cultivar comparisonsNutrient efficiency, sustainability reporting
IrrigationSoil moisture, evapotranspirationSAR backscatter (Sentinel-1), thermal RS, ET modelsFunctionalFlux tower data, lysimeter validationWater use efficiency, water policy compliance
Tillage/harvestSoil disturbance, residue coverSAR coherence, optical time seriesGenesis/StructuralIn situ soil disturbance surveysSoil conservation, monitoring sustainable practices
Crop rotationTemporal diversity, phenologyMulti-temporal NDVI/EVI, crop classificationGenesis/TaxonomicFarm records, phenological ground obs.Agri-environmental schemes, crop diversification targets
Field consolidationLandscape heterogeneity, field sizeHigh-res optical imagery, LiDAR boundariesStructuralField surveys, cadastral dataLand consolidation monitoring, biodiversity impacts
Intensified cropping cyclesAboveground biomass, multiple harvestsTime series (MODIS, Sentinel-2), SIF (FLEX, OCO-2)Genesis/FunctionalYield data, harvest recordsProductivity vs. sustainability trade-offs
Hedgerow removal/additionSemi-natural habitat, species richnessHigh-res imagery (UAV, Planet), landscape metricsStructural/TaxonomicBiodiversity field surveysCAP greening measures, landscape conservation
Table A7. Sensors and emerging technologies for A-LUI monitoring.
Table A7. Sensors and emerging technologies for A-LUI monitoring.
Technology/ApproachExample Missions or ToolsIndicator Categories AddressedSpatial/Temporal ResolutionDevelopment StageAdded Value
Multispectral opticalLandsat, Sentinel-2, PlanetScopeTrait (NDVI, chlorophyll, phenology)10–30 m/5–16 dOperationalLong time series, global coverage
HyperspectralEnMAP, CHIME, PRISMATrait (chlorophyll, N, stress proxies)20–30 m/<30 dOperational/newDetailed biochemical information
Thermal infraredECOSTRESS, Landsat TIRS, MODISFunctional (evapotranspiration, irrigation)70–1000 m/daily–16 dOperationalDirect link to water/energy fluxes
Radar (SAR)Sentinel-1, RADARSAT, ALOS PALSARStructure (tillage, harvest, soil moisture)10–30 m/6–12 dOperationalAll-weather, soil and canopy penetration
LiDARGEDI, ICESat-2, airborne LiDARStructure (canopy height, biomass, terraces)1–25 m/campaign-basedOperational/campaign3D structure, fine-scale terrain
UAV-based platformsMultispectral & thermal dronesTrait & Structure (field scale)cm–dm/flexibleOperationalUltra-high resolution, flexible timing
Solar-Induced Fluorescence (SIF)OCO-2, FLEX (upcoming)Functional (photosynthesis, GPP)300 m–2 km/dailyResearch/upcomingDirect proxy for photosynthesis
Multi-sensor fusionSentinel-1 + Sentinel-2, optical + thermalAll categoriesDepends on dataResearch & operationalImproves robustness & accuracy
AI/ML approachesDeep learning, data fusion methodsAll categoriesDepends on training dataResearch & early operationalEnhanced pattern recognition
Semantic web and linked dataRDF/OWL/SPARQL ontologiesData integrationN/AConceptualHarmonisation across datasets

References

  1. Diogo, V.; Helfenstein, J.; Mohr, F.; Varghese, V.; Debonne, N.; Levers, C.; Swart, R.; Sonderegger, G.; Nemecek, T.; Schader, C.; et al. Developing context-specific frameworks for integrated sustainability assessment of agricultural intensity change: An application for Europe. Environ. Sci. Policy 2022, 137, 128–142. [Google Scholar] [CrossRef]
  2. Tripathi, S.; Srivastava, P.; Devi, R.S.; Bhadouria, R. Influence of synthetic fertilizers and pesticides on soil health and soil microbiology. In Agrochemicals Detection, Treatment and Remediation; Elsevier: Amsterdam, The Netherlands, 2020; pp. 25–54. [Google Scholar]
  3. Yahaya, S.M.; Mahmud, A.A.; Abdullahi, M.; Haruna, A. Recent advances in the chemistry of nitrogen, phosphorus and potassium as fertilizers in soil: A review. Pedosphere 2023, 33, 385–406. [Google Scholar] [CrossRef]
  4. Shah, A.N.; Tanveer, M.; Shahzad, B.; Yang, G.; Fahad, S.; Ali, S.; Bukhari, M.A.; Tung, S.A.; Hafeez, A.; Souliyanonh, B. Soil compaction effects on soil health and cropproductivity: An overview. Environ. Sci. Pollut. Res. 2017, 24, 10056–10067. [Google Scholar] [CrossRef]
  5. Ingrao, C.; Strippoli, R.; Lagioia, G.; Huisingh, D. Water scarcity in agriculture: An overview of causes, impacts and approaches for reducing the risks. Heliyon 2023, 9, e18507. [Google Scholar] [CrossRef] [PubMed]
  6. Scanlon, B.R.; Jolly, I.; Sophocleous, M.; Zhang, L. Global impacts of conversions from natural to agricultural ecosystems on water resources: Quantity versus quality. Water Resour. Res. 2007, 43, W03437. [Google Scholar] [CrossRef]
  7. Sharma, K.; Rajan, S.; Nayak, S.K. Water pollution: Primary sources and associated human health hazards with special emphasis on rural areas. In Water Resources Management for Rural Development; Elsevier: Amsterdam, The Netherlands, 2024; pp. 3–14. [Google Scholar]
  8. Pannard, A.; Souchu, P.; Chauvin, C.; Delabuis, M.; Gascuel-Odoux, C.; Jeppesen, E.; Le Moal, M.; Ménesguen, A.; Pinay, G.; Rabalais, N.N.; et al. Why are there so many definitions of eutrophication? Ecol. Monogr. 2024, 94, e1616. [Google Scholar] [CrossRef]
  9. Zabel, F.; Delzeit, R.; Schneider, J.M.; Seppelt, R.; Mauser, W.; Václavík, T. Global impacts of future cropland expansion and intensification on agricultural markets and biodiversity. Nat. Commun. 2019, 10, 2844. [Google Scholar] [CrossRef] [PubMed]
  10. Mupepele, A.C.; Bruelheide, H.; Brühl, C.; Dauber, J.; Fenske, M.; Freibauer, A.; Gerowitt, B.; Krüß, A.; Lakner, S.; Plieninger, T.; et al. Biodiversity in European agricultural landscapes: Transformative societal changes needed. Trends Ecol. Evol. 2021, 36, 1067–1070. [Google Scholar] [CrossRef]
  11. Burian, A.; Kremen, C.; Wu, J.S.T.; Beckmann, M.; Bulling, M.; Garibaldi, L.A.; Krisztin, T.; Mehrabi, Z.; Ramankutty, N.; Seppelt, R. Biodiversity–production feedback effects lead to intensification traps in agricultural landscapes. Nat. Ecol. Evol. 2024, 8, 752–760. [Google Scholar] [CrossRef] [PubMed]
  12. Le Provost, G.; Thiele, J.; Westphal, C.; Penone, C.; Allan, E.; Neyret, M.; van der Plas, F.; Ayasse, M.; Bardgett, R.D.; Birkhofer, K.; et al. Contrasting responses of above-and belowground diversity to multiple components of land-use intensity. Nat. Commun. 2021, 12, 3918. [Google Scholar] [CrossRef]
  13. Beckmann, M.; Gerstner, K.; Akin-Fajiye, M.; Ceaușu, S.; Kambach, S.; Kinlock, N.L.; Phillips, H.R.P.; Verhagen, W.; Gurevitch, J.; Klotz, S.; et al. Conventional land-use intensification reduces species richness and increases production: A global meta-analysis. Glob. Change Biol. 2019, 25, 1941–1956. [Google Scholar] [CrossRef]
  14. Felipe-Lucia, M.R.; Soliveres, S.; Penone, C.; Fischer, M.; Ammer, C.; Boch, S.; Boeddinghaus, R.S.; Bonkowski, M.; Buscot, F.; Fiore-Donno, A.M.; et al. Land-use intensity alters networks between biodiversity, ecosystem functions, and services. Proc. Natl. Acad. Sci. 2020, 117, 28140–28149. [Google Scholar] [CrossRef] [PubMed]
  15. Gossner, M.M.; Lewinsohn, T.M.; Kahl, T.; Grassein, F.; Boch, S.; Prati, D.; Birkhofer, K.; Renner, S.C.; Sikorski, J.; Wubet, T.; et al. Land-use intensification causes multitrophic homogenization of grassland communities. Nature 2016, 540, 266–269. [Google Scholar] [CrossRef] [PubMed]
  16. Millard, J.; Outhwaite, C.L.; Kinnersley, R.; Freeman, R.; Gregory, R.D.; Adedoja, O.; Gavini, S.; Kioko, E.; Kuhlmann, M.; Ollerton, J.; et al. Global effects of land-use intensity on local pollinator biodiversity. Nat. Commun. 2021, 12, 2902. [Google Scholar] [CrossRef] [PubMed]
  17. Suarez, A.; Gwozdz, W. On the relation between monocultures and ecosystem services in the Global South: A review. Biol. Conserv. 2023, 278, 109870. [Google Scholar] [CrossRef]
  18. Cheng, C.; Liu, Z.; Song, W.; Chen, X.; Zhang, Z.; Li, B.; van Kleunen, M.; Wu, J. Biodiversity increases resistance of grasslands against plant invasions under multiple environmental changes. Nat. Commun. 2024, 15, 4506. [Google Scholar] [CrossRef]
  19. Scanes, C.G. Human Activity and Habitat Loss: Destruction, Fragmentation, and Degradation. In Animals and Human Society; Elsevier: Amsterdam, The Netherlands, 2018; pp. 451–482. ISBN 9780128052471. [Google Scholar]
  20. Oliver, T.H.; Marshall, H.H.; Morecroft, M.D.; Brereton, T.; Prudhomme, C.; Huntingford, C. Interacting effects of climate change and habitat fragmentation on drought-sensitive butterflies. Nat. Clim. Change 2015, 5, 941–946. [Google Scholar] [CrossRef]
  21. Peters, M.K.; Hemp, A.; Appelhans, T.; Becker, J.N.; Behler, C.; Classen, A.; Detsch, F.; Ensslin, A.; Ferger, S.W.; Frederiksen, S.B.; et al. Climate–land-use interactions shape tropical mountain biodiversity and ecosystem functions. Nature 2019, 568, 88–92. [Google Scholar] [CrossRef] [PubMed]
  22. Schneider, J.M.; Delzeit, R.; Neumann, C.; Heimann, T.; Seppelt, R.; Schuenemann, F.; Söder, M.; Mauser, W.; Zabel, F. Effects of profit-driven cropland expansion and conservation policies. Nat. Sustain. 2024, 7, 1335–1347. [Google Scholar] [CrossRef]
  23. Kaur, R.; Choudhary, D.; Bali, S.; Bandral, S.S.; Singh, V.; Ahmad, M.A.; Rani, N.; Singh, T.G.; Chandrasekaran, B. Pesticides: An alarming detrimental to health and environment. Sci. Total Environ. 2024, 915, 170113. [Google Scholar] [CrossRef]
  24. Bava, R.; Castagna, F.; Lupia, C.; Poerio, G.; Liguori, G.; Lombardi, R.; Naturale, M.D.; Mercuri, C.; Bulotta, R.M.; Britti, D.; et al. Antimicrobial Resistance in Livestock: A Serious Threat to Public Health. Antibiotics 2024, 13, 551. [Google Scholar] [CrossRef]
  25. Khmaissa, M.; Zouari-Mechichi, H.; Sciara, G.; Record, E.; Mechichi, T. Pollution from livestock farming antibiotics an emerging environmental and human health concern: A review. J. Hazard. Mater. Adv. 2024, 13, 100410. [Google Scholar] [CrossRef]
  26. Turner, B.L.; Doolittle, W.E. The Concept and Measure of Agricultural Intensity. Prof. Geogr. 1978, 30, 297–301. [Google Scholar] [CrossRef]
  27. Herzog, F.; Steiner, B.; Bailey, D.; Baudry, J.; Billeter, R.; Bukácek, R.; De Blust, G.; De Cock, R.; Dirksen, J.; Dormann, C.F.; et al. Assessing the intensity of temperate European agriculture at the landscape scale. Eur. J. Agron. 2006, 24, 165–181. [Google Scholar] [CrossRef]
  28. Lüscher, G.; Jeanneret, P.; Schneider, M.K.; Turnbull, L.A.; Arndorfer, M.; Balázs, K.; Báldi, A.; Bailey, D.; Bernhardt, K.G.; Choisis, J.P.; et al. Responses of plants, earthworms, spiders and bees to geographic location, agricultural management and surrounding landscape in European arable fields. Agric. Ecosyst. Environ. 2014, 186, 124–134. [Google Scholar] [CrossRef]
  29. Helfenstein, J.; Hepner, S.; Kreuzer, A.; Achermann, G.; Williams, T.; Bürgi, M.; Debonne, N.; Dimopoulos, T.; Diogo, V.; Fjellstad, W.; et al. Divergent agricultural development pathways across farm and landscape scales in Europe: Implications for sustainability and farmer satisfaction. Glob. Environ. Change 2024, 86, 102855. [Google Scholar] [CrossRef]
  30. Canisius, F.; Turral, H.; Molden, D. Fourier analysis of historical NOAA time series data to estimate bimodal agriculture. Int. J. Remote Sens. 2007, 28, 5503–5522. [Google Scholar] [CrossRef]
  31. Kuemmerle, T.; Hostert, P.; St-Louis, V.; Radeloff, V.C. Using image texture to map farmland field size: A case study in Eastern Europe. J. Land Use Sci. 2009, 4, 85–107. [Google Scholar] [CrossRef]
  32. Verburg, P.H.; Neumann, K.; Nol, L. Challenges in using land use and land cover data for global change studies. Glob. Change Biol. 2011, 17, 974–989. [Google Scholar] [CrossRef]
  33. Kuemmerle, T.; Erb, K.; Meyfroidt, P.; Müller, D.; Verburg, P.H.; Estel, S.; Haberl, H.; Hostert, P.; Jepsen, M.R.; Kastner, T.; et al. Challenges and opportunities in mapping land use intensity globally. Curr. Opin. Environ. Sustain. 2013, 5, 484–493. [Google Scholar] [CrossRef]
  34. Wellmann, T.; Haase, D.; Knapp, S.; Salbach, C.; Selsam, P.; Lausch, A. Urban land use intensity assessment: The potential of spatio-temporal spectral traits with remote sensing. Ecol. Indic. 2018, 85, 190–203. [Google Scholar] [CrossRef]
  35. Weber, D.; Schwieder, M.; Ritter, L.; Koch, T.; Psomas, A.; Huber, N.; Ginzler, C.; Boch, S. Grassland-use intensity maps for Switzerland based on satellite time series: Challenges and opportunities for ecological applications. Remote Sens. Ecol. Conserv. 2024, 10, 312–327. [Google Scholar] [CrossRef]
  36. Liu, H.; Zhou, Z.; Wen, Q.; Chen, J.; Kojima, S. Spatiotemporal Land Use/Land Cover Changes and Impact on Urban Thermal Environments: Analyzing Cool Island Intensity Variations. Sustainability 2024, 16, 3205. [Google Scholar] [CrossRef]
  37. Qiu, B.; Liu, B.; Tang, Z.; Dong, J.; Xu, W.; Liang, J.; Chen, N.; Chen, J.; Wang, L.; Zhang, C.; et al. National-scale 10-m maps of cropland use intensity in China during 2018–2023. Sci. Data 2024, 11, 691. [Google Scholar] [CrossRef]
  38. Hank, T.B.; Berger, K.; Bach, H.; Clevers, J.G.P.W.; Gitelson, A.; Zarco-Tejada, P.; Mauser, W. Spaceborne Imaging Spectroscopy for Sustainable Agriculture: Contributions and Challenges; Springer: Dordrecht, The Netherlands, 2019; Volume 40, ISBN 0123456789. [Google Scholar]
  39. Wulder, M.A.; Coops, N.C. Make Earth observations open access. Nature 2014, 513, 30–31. [Google Scholar] [CrossRef]
  40. 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, 113195. [Google Scholar] [CrossRef]
  41. Malenovský, Z.; Rott, H.; Cihlar, J.; Schaepman, M.E.; García-Santos, G.; Fernandes, R.; Berger, M. Sentinels for science: Potential of Sentinel-1, -2, and -3 missions for scientific observations of ocean, cryosphere, and land. Remote Sens. Environ. 2012, 120, 91–101. [Google Scholar] [CrossRef]
  42. Guanter, L.; Kaufmann, H.; Segl, K.; Foerster, S.; Rogass, C.; Chabrillat, S.; Kuester, T.; Hollstein, A.; Rossner, G.; Chlebek, C.; et al. The EnMAP Spaceborne Imaging Spectroscopy Mission for Earth Observation. Remote Sens. 2015, 7, 8830–8857. [Google Scholar] [CrossRef]
  43. Nieke, J.; Rast, M. Towards the Copernicus Hyperspectral Imaging Mission For The Environment (CHIME). In Proceedings of the IGARSS 2018—2018 IEEE International Geoscience and Remote Sensing Symposium, Valencia, Spain, 22–27 July 2018; pp. 157–159. [Google Scholar]
  44. Dubayah, R.; Blair, J.B.; Goetz, S.; Fatoyinbo, L.; Hansen, M.; Healey, S.; Hofton, M.; Hurtt, G.; Kellner, J.; Luthcke, S.; et al. The Global Ecosystem Dynamics Investigation: High-resolution laser ranging of the Earth’s forests and topography. Sci. Remote Sens. 2020, 1, 100002. [Google Scholar] [CrossRef]
  45. Lee, C.M.; Cable, M.L.; Hook, S.J.; Green, R.O.; Ustin, S.L.; Mandl, D.J.; Middleton, E.M. An introduction to the NASA Hyperspectral InfraRed Imager (HyspIRI) mission and preparatory activities. Remote Sens. Environ. 2015, 167, 6–19. [Google Scholar] [CrossRef]
  46. Kraft, S.; Del Bello, U.; Bouvet, M.; Drusch, M.; Moreno, J. FLEX: ESA’s Earth Explorer 8 candidate mission. In Proceedings of the 2012 IEEE International Geoscience and Remote Sensing Symposium, Munich, Germany, 22–27 July 2012; pp. 7125–7128. [Google Scholar]
  47. Omia, E.; Bae, H.; Park, E.; Kim, M.S.; Baek, I.; Kabenge, I.; Cho, B.K. Remote Sensing in Field Crop Monitoring: A Comprehensive Review of Sensor Systems, Data Analyses and Recent Advances. Remote Sens. 2023, 15, 354. [Google Scholar] [CrossRef]
  48. Sethy, P.K.; Pandey, C.; Sahu, Y.K.; Behera, S.K. Hyperspectral Imagery Applications for Precision Agriculture—A Systemic Survey; Springer: New York, NY, USA, 2022; Volume 81, ISBN 0123456789. [Google Scholar]
  49. Maes, W.H.; Steppe, K. Perspectives for Remote Sensing with Unmanned Aerial Vehicles in Precision Agriculture. Trends Plant Sci. 2019, 24, 152–164. [Google Scholar] [CrossRef]
  50. Khanal, S.; Fulton, J.; Shearer, S. An overview of current and potential applications of thermal remote sensing in precision agriculture. Comput. Electron. Agric. 2017, 139, 22–32. [Google Scholar] [CrossRef]
  51. Segarra, J.; Buchaillot, M.L.; Araus, J.L.; Kefauver, S.C. Remote Sensing for Precision Agriculture: Sentinel-2 Improved Features and Applications. Agronomy 2020, 10, 641. [Google Scholar] [CrossRef]
  52. Barbosa Júnior, M.R.; de Almeida Moreira, B.R.; Carreira, V.D.S.; Brito Filho, A.L.d.; Trentin, C.; Souza, F.L.P.d.; Tedesco, D.; Setiyono, T.; Flores, J.P.; Ampatzidis, Y.; et al. Precision agriculture in the United States: A comprehensive meta-review inspiring further research, innovation, and adoption. Comput. Electron. Agric. 2024, 221, 108993. [Google Scholar] [CrossRef]
  53. Marques, P.; Pádua, L.; Sousa, J.J.; Fernandes-Silva, A. Advancements in Remote Sensing Imagery Applications for Precision Management in Olive Growing: A Systematic Review. Remote Sens. 2024, 16, 1324. [Google Scholar] [CrossRef]
  54. Bonfanti, J.; Langridge, J.; Avadí Ef, A.; Casajus, N.; Chaudhary, A.; Damour Hi, G.; Estrada-Carmona, N.; Jones, S.K.; Makowski, D.; Mitchell, M.; et al. Global review of meta-analyses reveals key data gaps in agricultural impact studies on biodiversity in croplands. BioRxiv 2024. [Google Scholar] [CrossRef]
  55. Khanal, S.; Kushal, K.C.; Fulton, J.P.; Shearer, S.; Ozkan, E. Remote sensing in agriculture—Accomplishments, limitations, and opportunities. Remote Sens. 2020, 12, 3783. [Google Scholar] [CrossRef]
  56. Lu, B.; Dao, P.D.; Liu, J.; He, Y.; Shang, J. Recent advances of hyperspectral imaging technology and applications in agriculture. Remote Sens. 2020, 12, 2659. [Google Scholar] [CrossRef]
  57. Lausch, A.; Selsam, P.; Pause, M.; Bumberger, J. Monitoring vegetation- and geodiversity with remote sensing and traits. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2024, 382, 20230058. [Google Scholar] [CrossRef] [PubMed]
  58. Cavender-Bares, J.; Gamon, J.A.; Townsend, P.A. Remote Sensing of Plant Biodiversity; Cavender-Bares, J., Gamon, J.A., Townsend, P.A., Eds.; Springer International Publishing: Cham, Switzerland, 2020; ISBN 978-3-030-33156-6. [Google Scholar]
  59. Lausch, A.; Bastian, O.; Klotz, S.; Leitão, P.J.; Jung, A.; Rocchini, D.; Schaepman, M.E.; Skidmore, A.K.; Tischendorf, L.; Knapp, S. Understanding and assessing vegetation health by in situ species and remote-sensing approaches. Methods Ecol. Evol. 2018, 9, 1799–1809. [Google Scholar] [CrossRef]
  60. Lausch, A.; Baade, J.; Bannehr, L.; Borg, E.; Bumberger, J.; Chabrilliat, S.; Dietrich, P.; Gerighausen, H.; Glässer, C.; Hacker, J.; et al. Linking Remote Sensing and Geodiversity and Their Traits Relevant to Biodiversity—Part I: Soil Characteristics. Remote Sens. 2019, 11, 2356. [Google Scholar] [CrossRef]
  61. Lausch, A.; Schaepman, M.E.; Skidmore, A.K.; Truckenbrodt, S.C.; Hacker, J.M.; Baade, J.; Bannehr, L.; Borg, E.; Bumberger, J.; Dietrich, P.; et al. Linking the Remote Sensing of Geodiversity and Traits Relevant to Biodiversity—Part II: Geomorphology, Terrain and Surfaces. Remote Sens. 2020, 12, 3690. [Google Scholar] [CrossRef]
  62. Lausch, A.; Schaepman, M.E.; Skidmore, A.K.; Catana, E.; Bannehr, L.; Bastian, O.; Borg, E.; Bumberger, J.; Dietrich, P.; Glässer, C.; et al. Remote Sensing of Geomorphodiversity Linked to Biodiversity—Part III: Traits, Processes and Remote Sensing Characteristics. Remote Sens. 2022, 14, 2279. [Google Scholar] [CrossRef]
  63. Lausch, A.; Bannehr, L.; Berger, S.A.; Borg, E.; Bumberger, J. Monitoring Water Diversity and Water Quality with Remote Sensing and Traits. Remote Sens. 2024, 16, 2425. [Google Scholar] [CrossRef]
  64. Andersson, E.; Haase, D.; Anderson, P.; Cortinovis, C.; Goodness, J.; Kendal, D.; Lausch, A.; McPhearson, T.; Sikorska, D.; Wellmann, T. What are the traits of a social-ecological system: Towards a framework in support of urban sustainability. Npj Urban Sustain. 2021, 1, 14. [Google Scholar] [CrossRef]
  65. Wellmann, T.; Lausch, A.; Andersson, E.; Knapp, S.; Cortinovis, C.; Jache, J.; Scheuer, S.; Kremer, P.; Mascarenhas, A.; Kraemer, R.; et al. Remote sensing in urban planning: Contributions towards ecologically sound policies? Landsc. Urban Plan. 2020, 204, 103921. [Google Scholar] [CrossRef]
  66. Xie, C.; Wang, J.; Haase, D.; Wellmann, T.; Lausch, A. Measuring spatio-temporal heterogeneity and interior characteristics of green spaces in urban neighborhoods: A new approach using gray level co-occurrence matrix. Sci. Total Environ. 2023, 855, 158608. [Google Scholar] [CrossRef]
  67. Roilo, S.; Paulus, A.; Alarcón-Segura, V.; Kock, L.; Beckmann, M.; Klein, N.; Cord, A.F. Quantifying agricultural land-use intensity for spatial biodiversity modelling: Implications of different metrics and spatial aggregation methods. Landsc. Ecol. 2024, 39, 55. [Google Scholar] [CrossRef]
  68. Erb, K.H.; Haberl, H.; Jepsen, M.R.; Kuemmerle, T.; Lindner, M.; Müller, D.; Verburg, P.H.; Reenberg, A. A conceptual framework for analysing and measuring land-use intensity. Curr. Opin. Environ. Sustain. 2013, 5, 464–470. [Google Scholar] [CrossRef]
  69. Çakmakçı, R.; Salık, M.A.; Çakmakçı, S. Assessment and Principles of Environmentally Sustainable Food and Agri-culture Systems. Agriculture 2023, 13, 1073. [Google Scholar] [CrossRef]
  70. Dullinger, I.; Essl, F.; Moser, D.; Erb, K.; Haberl, H.; Dullinger, S. Biodiversity models need to represent land-use intensity more comprehensively. Glob. Ecol. Biogeogr. 2021, 30, 924–932. [Google Scholar] [CrossRef]
  71. Rascher, U.; Alonso, L.; Burkart, A.; Cilia, C.; Cogliati, S.; Colombo, R.; Damm, A.; Drusch, M.; Guanter, L.; Hanus, J.; et al. Sun-induced fluorescence—A new probe of photosynthesis: First maps from the imaging spectrometer HyPlant. Glob. Change Biol. 2015, 21, 4673–4684. [Google Scholar] [CrossRef] [PubMed]
  72. Mohammed, G.H.; Colombo, R.; Middleton, E.M.; Rascher, U.; van der Tol, C.; Nedbal, L.; Goulas, Y.; Pérez-Priego, O.; Damm, A.; Meroni, M.; et al. Remote sensing of solar-induced chlorophyll fluorescence (SIF) in vegetation: 50 years of progress. Remote Sens. Environ. 2019, 231, 111177. [Google Scholar] [CrossRef]
  73. Anderson, J.R.; Hardy, E.E.; Roach, J.T.; Witmer, R.E. A Land Use and Land Cover Classification System for Use with Remote Sensor Data; US Government Printing Office: Washington, DC, USA, 1976; Volume 964.
  74. Adewumi, J.R.; Akomolafe, J.K.; Ajibade, F.O.; Fabeku, B.B. Application of GIS and Remote Sensing Technique to Change Detection in Land Use/Land Cover Mapping of Igbokoda, Ondo State, Nigeria. J. Appl. Sci. Process Eng. 1970, 3, 34–54. [Google Scholar] [CrossRef]
  75. Matthews, E. Global Vegetation and Land Use: New High-Resolution Data Bases for Climate Studies. J. Clim. Appl. Meteorol. 1983, 22, 474–487. [Google Scholar] [CrossRef]
  76. Masek, J.G.; Wulder, M.A.; Markham, B.; McCorkel, J.; Crawford, C.J.; Storey, J.; Jenstrom, D.T. Landsat 9: Empowering open science and applications through continuity. Remote Sens. Environ. 2020, 248, 111968. [Google Scholar] [CrossRef]
  77. Copernicus Copernicus the European Union’s Earth Observation Programme. Available online: https://www.copernicus.eu (accessed on 2 August 2024).
  78. Rocchini, D.; Santos, M.J.; Ustin, S.L.; Féret, J.; Asner, G.P.; Beierkuhnlein, C.; Dalponte, M.; Feilhauer, H.; Foody, G.M.; Geller, G.N.; et al. The Spectral Species Concept in Living Color. J. Geophys. Res. Biogeosci. 2022, 127, e2022JG007026. [Google Scholar] [CrossRef]
  79. Maudet, S.; Brusse, T.; Poss, B.; Caro, G.; Marrec, R. Estimating landscape intensity through farming practices: An integrative and flexible approach to modelling farming intensity from field to landscape. Ecol. Modell. 2025, 501, 110975. [Google Scholar] [CrossRef]
  80. Lausch, A.; Menz, G. Bedeutung der Integration linearer Elemente in Fernerkundungsdaten zur Berechnung von Landschaftsstrukturmaßen. Photogramm. Fernerkund. Geoinf. 1999, 3, 185–194. [Google Scholar]
  81. Lausch, A.; Herzog, F. Applicability of landscape metrics for the monitoring of landscape change: Issues of scale, resolution and interpretability. Ecol. Indic. 2002, 2, 3–15. [Google Scholar] [CrossRef]
  82. Billeter, R.; Liira, J.; Bailey, D.; Bugter, R.; Arens, P.; Augenstein, I.; Aviron, S.; Baudry, J.; Bukacek, R.; Burel, F.; et al. Indicators for biodiversity in agricultural landscapes: A pan-European study. J. Appl. Ecol. 2008, 45, 141–150. [Google Scholar] [CrossRef]
  83. Mohr, F.; Diogo, V.; Helfenstein, J.; Debonne, N.; Dimopoulos, T.; Dramstad, W.; García-Martín, M.; Hernik, J.; Herzog, F.; Kizos, T.; et al. Why has farming in Europe changed? A farmers’ perspective on the development since the 1960s. Reg. Environ. Change 2023, 23, 156. [Google Scholar] [CrossRef] [PubMed]
  84. Martin, A.E.; Collins, S.J.; Crowe, S.; Girard, J.; Naujokaitis-Lewis, I.; Smith, A.C.; Lindsay, K.; Mitchell, S.; Fahrig, L. Effects of farmland heterogeneity on biodiversity are similar to—Or even larger than—The effects of farming practices. Agric. Ecosyst. Environ. 2020, 288, 106698. [Google Scholar] [CrossRef]
  85. Grenzdörffer, G. Grundlagen der Landwirtschaftlichen Fernerkundung; KTBL eV Kuratorium für Technik und Bauwesen in der Landwirtschaft: Darmstadt, Germany, 2022. [Google Scholar]
  86. Picado, E.F.; Romero, K.F. Mapping Spatial Variability of Sugarcane Foliar Nitrogen, Phosphorus, Potassium and Chlorophyll Concentrations Using Remote Sensing. Geomatics 2025, 5, 3. [Google Scholar] [CrossRef]
  87. Yu, M.; Rui, X.; Xie, W.; Xu, X.; Wei, W. Research on Automatic Identification Method of Terraces on the Loess Plateau Based on Deep Transfer Learning. Remote Sens. 2022, 14, 2446. [Google Scholar] [CrossRef]
  88. Wang, X.; Shu, L.; Han, R.; Yang, F.; Gordon, T.; Wang, X.; Xu, H. A Survey of Farmland Boundary Extraction Technology Based on Remote Sensing Images. Electronics 2023, 12, 1156. [Google Scholar] [CrossRef]
  89. Betbeder, J.; Hubert-Moy, L.; Burel, F.; Corgne, S.; Baudry, J. Assessing ecological habitat structure from local to landscape scales using synthetic aperture radar. Ecol. Indic. 2015, 52, 545–557. [Google Scholar] [CrossRef]
  90. Griffiths, P.; Nendel, C.; Hostert, P. Intra-annual reflectance composites from Sentinel-2 and Landsat for national-scale crop and land cover mapping. Remote Sens. Environ. 2019, 220, 135–151. [Google Scholar] [CrossRef]
  91. Bartold, M.; Kluczek, M.; Wróblewski, K.; Dąbrowska-Zielińska, K.; Goliński, P.; Golińska, B. Mapping management intensity types in grasslands with synergistic use of Sentinel-1 and Sentinel-2 satellite images. Sci. Rep. 2024, 14, 32066. [Google Scholar] [CrossRef]
  92. Günder, M.; Yamati, F.R.I.; Alcántara, A.A.B.; Mahlein, A.-K.; Sifa, R.; Bauckhage, C. SugarViT—Multi-objective Regression of UAV Images with Vision Transformers and Deep Label Distribution Learning Demonstrated on Disease Severity Prediction in Sugar Beet. PLoS ONE 2023, 20, e0318097. [Google Scholar] [CrossRef]
  93. Zhou, T.; Geng, Y.; Ji, C.; Xu, X.; Wang, H.; Pan, J.; Bumberger, J.; Haase, D.; Lausch, A. Prediction of soil organic carbon and the C:N ratio on a national scale using machine learning and satellite data: A comparison between Sentinel-2, Sentinel-3 and Landsat-8 images. Sci. Total Environ. 2021, 755, 142661. [Google Scholar] [CrossRef]
  94. Díaz, S.; Kattge, J.; Cornelissen, J.H.C.; Wright, I.J.; Lavorel, S.; Dray, S.; Reu, B.; Kleyer, M.; Wirth, C.; Colin Prentice, I.; et al. The global spectrum of plant form and function. Nature 2016, 529, 167–171. [Google Scholar] [CrossRef]
  95. Joswig, J.S.; Wirth, C.; Schuman, M.C.; Kattge, J.; Reu, B.; Wright, I.J.; Sippel, S.D.; Rüger, N.; Richter, R.; Schaepman, M.E.; et al. Climatic and soil factors explain the two-dimensional spectrum of global plant trait variation. Nat. Ecol. Evol. 2022, 6, 36–50. [Google Scholar] [CrossRef] [PubMed]
  96. Candiani, G.; Tagliabue, G.; Panigada, C.; Verrelst, J.; Picchi, V.; Rivera Caicedo, J.P.; Boschetti, M. Evaluation of Hybrid Models to Estimate Chlorophyll and Nitrogen Content of Maize Crops in the Framework of the Future CHIME Mission. Remote Sens. 2022, 14, 1792. [Google Scholar] [CrossRef] [PubMed]
  97. Verrelst, J.; Rivera, J.P.; Veroustraete, F.; Muñoz-Marí, J.; Clevers, J.G.P.W.; Camps-Valls, G.; Moreno, J. Experimental Sentinel-2 LAI estimation using parametric, non-parametric and physical retrieval methods—A comparison. ISPRS J. Photogramm. Remote Sens. 2015, 108, 260–272. [Google Scholar] [CrossRef]
  98. Berger, K.; Verrelst, J.; Féret, J.-B.; Hank, T.; Wocher, M.; Mauser, W.; Camps-Valls, G. Retrieval of aboveground crop nitrogen content with a hybrid machine learning method. Int. J. Appl. Earth Obs. Geoinf. 2020, 92, 102174. [Google Scholar] [CrossRef] [PubMed]
  99. Loizzo, R.; Daraio, M.; Guarini, R.; Longo, F.; Lorusso, R.; Dini, L.; Lopinto, E. Prisma Mission Status and Perspective. In Proceedings of the IGARSS 2019—2019 IEEE International Geoscience and Remote Sensing Symposium, Yokohama, Japan, 28 July–2 August 2019; pp. 4503–4506. [Google Scholar]
  100. Matsunaga, T.; Iwasaki, A.; Tachikawa, T.; Tanii, J.; Kashimura, O.; Mouri, K.; Inada, H.; Tsuchida, S.; Nakamura, R.; Yamamoto, H.; et al. Hyperspectral Imager Suite (HISUI): Its Launch and Current Status. In Proceedings of the IGARSS 2020–2020 IEEE International Geoscience and Remote Sensing Symposium, Waikoloa, HI, USA, 26 September–2 October 2020; pp. 3272–3273. [Google Scholar]
  101. Feingersh, T.; Dor, E. Ben SHALOM—A Commercial Hyperspectral Space Mission. In Optical Payloads for Space Missions; Wiley: New York, NY, USA, 2015; pp. 247–263. [Google Scholar]
  102. Delloye, C.; Weiss, M.; Defourny, P. Retrieval of the canopy chlorophyll content from Sentinel-2 spectral bands to estimate nitrogen uptake in intensive winter wheat cropping systems. Remote Sens. Environ. 2018, 216, 245–261. [Google Scholar] [CrossRef]
  103. Jacquemoud, S.; Verhoef, W.; Baret, F.; Bacour, C.; Zarco-Tejada, P.J.; Asner, G.P.; François, C.; Ustin, S.L. PROSPECT + SAIL models: A review of use for vegetation characterization. Remote Sens. Environ. 2009, 113, S56–S66. [Google Scholar] [CrossRef]
  104. Lhotáková, Z.; Neuwirthová, E.; Potůčková, M.; Červená, L.; Hunt, L.; Kupková, L.; Lukeš, P.; Campbell, P.; Albrechtová, J. Mind the leaf anatomy while taking ground truth with portable chlorophyll meters. Sci. Rep. 2025, 15, 1855. [Google Scholar] [CrossRef] [PubMed]
  105. Brewer, K.; Clulow, A.; Sibanda, M.; Gokool, S.; Naiken, V.; Mabhaudhi, T. Predicting the Chlorophyll Content of Maize over Phenotyping as a Proxy for Crop Health in Smallholder Farming Systems. Remote Sens. 2022, 14, 518. [Google Scholar] [CrossRef]
  106. Drusch, M.; Moreno, J.; Del Bello, U.; Franco, R.; Goulas, Y.; Huth, A.; Kraft, S.; Middleton, E.M.; Miglietta, F.; Mohammed, G.; et al. The FLuorescence EXplorer Mission Concept-ESA’s Earth Explorer 8. IEEE Trans. Geosci. Remote Sens. 2017, 55, 1273–1284. [Google Scholar] [CrossRef]
  107. De Grave, C.; Verrelst, J.; Morcillo-Pallarés, P.; Pipia, L.; Rivera-Caicedo, J.P.; Amin, E.; Belda, S.; Moreno, J. Quantifying vegetation biophysical variables from the Sentinel-3/FLEX tandem mission: Evaluation of the synergy of OLCI and FLORIS data sources. Remote Sens. Environ. 2020, 251, 112101. [Google Scholar] [CrossRef]
  108. Ač, A.; Malenovský, Z.; Olejníčková, J.; Gallé, A.; Rascher, U.; Mohammed, G. Meta-analysis assessing potential of steady-state chlorophyll fluorescence for remote sensing detection of plant water, temperature and nitrogen stress. Remote Sens. Environ. 2015, 168, 420–436. [Google Scholar] [CrossRef]
  109. Moreno, J.F.; Goulas, Y.; Huth, A.; Middleton, E.; Miglietta, F.; Mohammed, G.; Nedbal, L.; Rascher, U.; Verhoef, W.; Drusch, M. Very high spectral resolution imaging spectroscopy: The Fluorescence Explorer (FLEX) mission. Int. Geosci. Remote Sens. Symp. 2016, 2016, 264–267. [Google Scholar]
  110. Dong, N.; Prentice, I.C.; Wright, I.J.; Wang, H.; Atkin, O.K.; Bloomfield, K.J.; Domingues, T.F.; Gleason, S.M.; Maire, V.; Onoda, Y.; et al. Leaf nitrogen from the perspective of optimal plant function. J. Ecol. 2022, 110, 2585–2602. [Google Scholar] [CrossRef] [PubMed]
  111. Zheng, J.; Song, X.; Yang, G.; Du, X.; Mei, X.; Yang, X. Remote Sensing Monitoring of Rice and Wheat Canopy Nitrogen: A Review. Remote Sens. 2022, 14, 5712. [Google Scholar] [CrossRef]
  112. Silva, L.; Conceição, L.A.; Lidon, F.C.; Maçãs, B. Remote Monitoring of Crop Nitrogen Nutrition to Adjust Crop Models: A Review. Agriculture 2023, 13, 835. [Google Scholar] [CrossRef]
  113. Arogoundade, A.M.; Mutanga, O.; Odindi, J.; Naicker, R. The role of remote sensing in tropical grassland nutrient estimation: A review. Environ. Monit. Assess. 2023, 195, 954. [Google Scholar] [CrossRef] [PubMed]
  114. Xi, R.; Gu, Y.; Zhang, X.; Ren, Z. Nitrogen monitoring and inversion algorithms of fruit trees based on spectral remote sensing: A deep review. Front. Plant Sci. 2024, 15, 1489151. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, J.; Hu, Y.; Li, F.; Fue, K.G.; Yu, K. Meta-Analysis Assessing Potential of Drone Remote Sensing in Estimating Plant Traits Related to Nitrogen Use Efficiency. Remote Sens. 2024, 16, 838. [Google Scholar] [CrossRef]
  116. Ravikumar, S.; Vellingiri, G.; Sellaperumal, P.; Pandian, K.; Sivasankar, A.; Sangchul, H. Real-time nitrogen monitoring and management to augment N use efficiency and ecosystem sustainability—A review. J. Hazard. Mater. Adv. 2024, 16, 100466. [Google Scholar] [CrossRef]
  117. Chowdhury, M.; Kumar Khura, T.; Ahmad Parray, R.; Kushwaha, H.L.; Upadhyay, P.K.; Jha, A.; Patra, K.; Kushwah, A.; Kumar Prajapati, V. The use of destructive and nondestructive techniques in concrete nitrogen assessment in plants. J. Plant Nutr. 2024, 47, 2271–2294. [Google Scholar] [CrossRef]
  118. Tasnim, M.; Simic, A.; Verrelst, J.; Tian, Q.; Soleil, A.; Poku, H.; Rahman, A.; Processing, I.; Científic, P.; Val, U. De Optimizing Empirical and Hybrid Modeling for Advanced Canopy Chlorophyll and Nitrogen Retrieval Technique Using EnMAP Data. Environ. Chall. 2025, 18, 101114. [Google Scholar]
  119. Verrelst, J.; Rivera-Caicedo, J.P.; Reyes-Muñoz, P.; Morata, M.; Amin, E.; Tagliabue, G.; Panigada, C.; Hank, T.; Berger, K. Mapping landscape canopy nitrogen content from space using PRISMA data. ISPRS J. Photogramm. Remote Sens. 2021, 178, 382–395. [Google Scholar] [CrossRef]
  120. Wang, Y.; Suarez, L.; Hornero, A.; Poblete, T.; Ryu, D.; Gonzalez-Dugo, V.; Zarco-Tejada, P.J. Assessing plant traits derived from Sentinel-2 to characterize leaf nitrogen variability in almond orchards: Modeling and validation with airborne hyperspectral imagery. Precis. Agric. 2025, 26, 13. [Google Scholar] [CrossRef]
  121. Zhang, X.; Han, L.; Sobeih, T.; Lappin, L.; Lee, M.A.; Howard, A.; Kisdi, A. The Self-Supervised Spectral–Spatial Vision Transformer Network for Accurate Prediction of Wheat Nitrogen Status from UAV Imagery. Remote Sens. 2022, 14, 1400. [Google Scholar] [CrossRef]
  122. Xu, S.; Xu, X.; Blacker, C.; Gaulton, R.; Zhu, Q.; Yang, M.; Yang, G.; Zhang, J.; Yang, Y.; Yang, M.; et al. Estimation of Leaf Nitrogen Content in Rice Using Vegetation Indices and Feature Variable Optimization with Information Fusion of Multiple-Sensor Images from UAV. Remote Sens. 2023, 15, 854. [Google Scholar] [CrossRef]
  123. Smith, D.R.; King, K.W.; Johnson, L.; Francesconi, W.; Richards, P.; Baker, D.; Sharpley, A.N. Surface Runoff and Tile Drainage Transport of Phosphorus in the Midwestern United States. J. Environ. Qual. 2015, 44, 495–502. [Google Scholar] [CrossRef] [PubMed]
  124. Valipour, M.; Krasilnikof, J.; Yannopoulos, S.; Kumar, R.; Deng, J.; Roccaro, P.; Mays, L.; Grismer, M.E.; Angelakis, A.N. The evolution of agricultural drainage from the earliest times to the present. Sustainability 2020, 12, 416. [Google Scholar] [CrossRef]
  125. Jepsen, M.R.; Kuemmerle, T.; Müller, D.; Erb, K.; Verburg, P.H.; Haberl, H.; Vesterager, J.P.; Andrič, M.; Antrop, M.; Austrheim, G.; et al. Transitions in European land-management regimes between 1800 and 2010. Land Use Policy 2015, 49, 53–64. [Google Scholar] [CrossRef]
  126. Davidson, N.C. How much wetland has the world lost? Long-term and recent trends in global wetland area. Mar. Freshw. Res. 2014, 65, 934–941. [Google Scholar] [CrossRef]
  127. Koch, J.; Elsgaard, L.; Greve, M.H.; Gyldenkærne, S.; Hermansen, C.; Levin, G.; Wu, S.; Stisen, S. Water-table-driven greenhouse gas emission estimates guide peatland restoration at national scale. Biogeosciences 2023, 20, 2387–2403. [Google Scholar] [CrossRef]
  128. Williamson, T.N.; Dobrowolski, E.G.; Meyer, S.M.; Frey, J.W.; Allred, B.J. Delineation of tile-drain networks using thermal and multispectral imagery—Implications for water quantity and quality differences from paired edge-of-field sites. J. Soil Water Conserv. 2018, 74, 1–11. [Google Scholar] [CrossRef]
  129. Carlsen, A.H.; Fensholt, R.; Looms, M.C.; Gominski, D.; Stisen, S.; Jepsen, M.R. Systematic review of the detection of subsurface drainage systems in agricultural fields using remote sensing systems. Agric. Water Manag. 2024, 299, 108892. [Google Scholar] [CrossRef]
  130. Karbs, H.H. Subsurface Drainage Mapping By Airborne Infrared. In Proceedings of the Oklahoma Academy of Science, Norman, OK, USA, 31 August–4 September 1970; Volume 18, pp. 10–18. [Google Scholar]
  131. Gökkaya, K.; Budhathoki, M.; Christopher, S.F.; Hanrahan, B.R.; Tank, J.L. Subsurface tile drained area detection using GIS and remote sensing in an agricultural watershed. Ecol. Eng. 2017, 108, 370–379. [Google Scholar] [CrossRef]
  132. Cho, E.; Jacobs, J.M.; Jia, X.; Kraatz, S. Identifying Subsurface Drainage using Satellite Big Data and Machine Learning via Google Earth Engine. Water Resour. Res. 2019, 55, 8028–8045. [Google Scholar] [CrossRef]
  133. Lai, Q.; Xin, Q.; Tian, Y.; Chen, X.; Li, Y.; Wu, R. Structural Analysis and 3D Reconstruction of Underground Pipeline Systems Based on LiDAR Point Clouds. Remote Sens. 2025, 17, 341. [Google Scholar] [CrossRef]
  134. Woo, D.K.; Ji, J.; Song, H. Subsurface drainage pipe detection using an ensemble learning approach and aerial images. Agric. Water Manag. 2023, 287, 108455. [Google Scholar] [CrossRef]
  135. Allred, B.; Eash, N.; Freeland, R.; Martinez, L.; Wishart, D.B. Effective and efficient agricultural drainage pipe mapping with UAS thermal infrared imagery: A case study. Agric. Water Manag. 2018, 197, 132–137. [Google Scholar] [CrossRef]
  136. Becker, A.M.; Becker, R.H.; Doro, K.O. Locating Drainage Tiles at a Wetland Restoration Site within the Oak Openings Region of Ohio, United States Using UAV and Land Based Geophysical Techniques. Wetlands 2021, 41, 116. [Google Scholar] [CrossRef]
  137. Liu, Z.; Chen, G.K.; Tang, B.; Wen, Q.; Tan, R.; Huang, Y. Regional scale terrace mapping in fragmented mountainous areas using multi-source remote sensing data and sample purification strategy. Sci. Total Environ. 2024, 925, 171366. [Google Scholar] [CrossRef]
  138. Lu, Y.; Li, X.; Xin, L.; Song, H.; Wang, X. Mapping the terraces on the Loess Plateau based on a deep learning-based model at 1.89 m resolution. Sci. Data 2023, 10, 115. [Google Scholar] [CrossRef]
  139. Ding, H.; Na, J.; Jiang, S.; Zhu, J.; Liu, K.; Fu, Y.; Li, F. Evaluation of three different machine learning methods for object-based artificial terrace mapping—A case study of the loess plateau, China. Remote Sens. 2021, 13, 1021. [Google Scholar] [CrossRef]
  140. Zhao, F.; Xiong, L.; Wang, C.; Wang, H.; Wei, H.; Tang, G. Terraces mapping by using deep learning approach from remote sensing images and digital elevation models. Trans. GIS 2021, 25, 2438–2454. [Google Scholar] [CrossRef]
  141. Zhai, D.; Dong, J.; Cadisch, G.; Wang, M.; Kou, W.; Xu, J.; Xiao, X.; Abbas, S. Comparison of pixel- and object-based approaches in phenology-based rubber plantation mapping in fragmented landscapes. Remote Sens. 2018, 10, 44. [Google Scholar] [CrossRef]
  142. Blickensdörfer, L.; Schwieder, M.; Pflugmacher, D.; Nendel, C.; Erasmi, S.; Hostert, P. Mapping of crop types and crop sequences with combined time series of Sentinel-1, Sentinel-2 and Landsat 8 data for Germany. Remote Sens. Environ. 2022, 269, 112831. [Google Scholar] [CrossRef]
  143. Godone, D.; Giordan, D.; Baldo, M. Rapid mapping application of vegetated terraces based on high resolution airborne lidar. Geomat. Nat. Hazards Risk 2018, 9, 970–985. [Google Scholar] [CrossRef]
  144. Le Vot, T.; Cohen, M.; Nowak, M.; Passy, P.; Sumera, F. Resilience of Terraced Landscapes to Human and Natural Impacts: A GIS-Based Reconstruction of Land Use Evolution in a Mediterranean Mountain Valley. Land 2024, 13, 592. [Google Scholar] [CrossRef]
  145. Garzón-Oechsle, A.; Johanson, E.; Nagarajan, S.; Martínez, V. In between the Sites: Understanding Late Holocene Manteño Agricultural Contexts in the Chongón-Colonche Mountains of Coastal Ecuador through UAV-Lidar and Excavation. J. F. Archaeol. 2025, 50, 42–59. [Google Scholar] [CrossRef]
  146. Lozić, E. Application of airborne lidar data to the archaeology of agrarian land use. The case study of the early medieval microregion of bled (Slovenia). Remote Sens. 2021, 13, 3228. [Google Scholar] [CrossRef]
  147. Masini, N.; Gizzi, F.T.; Biscione, M.; Fundone, V.; Sedile, M.; Sileo, M.; Pecci, A.; Lacovara, B.; Lasaponara, R. Medieval archaeology under the canopy with LiDAR. the (Re)discovery of a medieval fortified settlement in southern Italy. Remote Sens. 2018, 10, 1598. [Google Scholar] [CrossRef]
  148. Smith, C.; Baker, J.C.A.; Spracklen, D.V. Tropical deforestation causes large reductions in observed precipitation. Nature 2023, 615, 270–275. [Google Scholar] [CrossRef] [PubMed]
  149. Hansen, M.C.; Potapov, P.V.; Moore, R.; Hancher, M.; Turubanova, S.A.; Tyukavina, A.; Thau, D.; Stehman, S.V.; Goetz, S.J.; Loveland, T.R.; et al. High-Resolution Global Maps of 21st-Century Forest Cover Change. Science 2013, 342, 850–853. [Google Scholar] [CrossRef]
  150. Ponvert-Delisles Batista, D.R.; Estrada-Medina, H.; Gijón-Yescas, G.N.; Álvarez-Rivera, O.O. Land covers analyses during slash and burn agriculture by using multispectral imagery obtained with Unattended Aerial Vehicles (UAVs). Trop. Subtrop. Agroecosystems 2021, 24, 21. [Google Scholar] [CrossRef]
  151. Lechner, A.M.; Foody, G.M.; Boyd, D.S. Applications in Remote Sensing to Forest Ecology and Management. One Earth 2020, 2, 405–412. [Google Scholar] [CrossRef]
  152. Lehmann, E.A.; Caccetta, P.; Lowell, K.; Mitchell, A.; Zhou, Z.S.; Held, A.; Milne, T.; Tapley, I. SAR and optical remote sensing: Assessment of complementarity and interoperability in the context of a large-scale operational forest monitoring system. Remote Sens. Environ. 2015, 156, 335–348. [Google Scholar] [CrossRef]
  153. Ballère, M.; Bouvet, A.; Mermoz, S.; Le Toan, T.; Koleck, T.; Bedeau, C.; André, M.; Forestier, E.; Frison, P.L.; Lardeux, C. SAR data for tropical forest disturbance alerts in French Guiana: Benefit over optical imagery. Remote Sens. Environ. 2021, 252, 112159. [Google Scholar] [CrossRef]
  154. Tarazona, Y.; Mantas, V.M.; Pereira, A.J.S.C. Improving tropical deforestation detection through using photosynthetic vegetation time series–(PVts-β). Ecol. Indic. 2018, 94, 367–379. [Google Scholar] [CrossRef]
  155. Bégué, A.; Arvor, D.; Bellon, B.; Betbeder, J.; de Abelleyra, D.; Ferraz, R.P.D.; Lebourgeois, V.; Lelong, C.; Simões, M.; Verón, S.R. Remote sensing and cropping practices: A review. Remote Sens. 2018, 10, 99. [Google Scholar] [CrossRef]
  156. Preidl, S.; Lange, M.; Doktor, D. Introducing APiC for regionalised land cover mapping on the national scale using Sentinel-2A imagery. Remote Sens. Environ. 2020, 240, 111673. [Google Scholar] [CrossRef]
  157. Baessler, C.; Klotz, S. Effects of changes in agricultural land-use on landscape structure and arable weed vegetation over the last 50 years. Agric. Ecosyst. Environ. 2006, 115, 43–50. [Google Scholar] [CrossRef]
  158. Chapman, T. Calculating the Interspersion and Juxtaposition Rates for Saint Clair Flats State Wildlife Area. Ph.D. Thesis, University of Redlands, Redlands, CA, USA, 2022. [Google Scholar]
  159. Uuemaa, E.; Antrop, M.; Roosaare, J.; Marja, R.; Mander, Ü. Landscape metrics and indices: An overview of their use in landscape research. Living Rev. Landsc. Res. 2009, 3, 1–28. [Google Scholar] [CrossRef]
  160. Oksanen, T. Shape-describing indices for agricultural field plots and their relationship to operational efficiency. Comput. Electron. Agric. 2013, 98, 252–259. [Google Scholar] [CrossRef]
  161. Griffel, L.M.; Vazhnik, V.; Hartley, D.S.; Hansen, J.K.; Roni, M. Agricultural field shape descriptors as predictors of field efficiency for perennial grass harvesting: An empirical proof. Comput. Electron. Agric. 2020, 168, 105088. [Google Scholar] [CrossRef]
  162. Salas, E.A.L.; Subburayalu, S.K. Correction: Modified shape index for object-based random forest image classification of agricultural systems using airborne hyperspectral datasets. PLoS ONE 2019, 14, e0222474. [Google Scholar] [CrossRef]
  163. Blüthgen, N.; Dormann, C.F.; Prati, D.; Klaus, V.H.; Kleinebecker, T.; Hölzel, N.; Alt, F.; Boch, S.; Gockel, S.; Hemp, A.; et al. A quantitative index of land-use intensity in grasslands: Integrating mowing, grazing and fertilization. Basic Appl. Ecol. 2012, 13, 207–220. [Google Scholar] [CrossRef]
  164. Rocchini, D.; Balkenhol, N.; Carter, G.A.; Foody, G.M.; Gillespie, T.W.; He, K.S.; Kark, S.; Levin, N.; Lucas, K.; Luoto, M.; et al. Remotely sensed spectral heterogeneity as a proxy of species diversity: Recent advances and open challenges. Ecol. Inform. 2010, 5, 318–329. [Google Scholar] [CrossRef]
  165. Rocchini, D.; Marcantonio, M.; Ricotta, C. Measuring Rao’s Q diversity index from remote sensing: An open source solution. Ecol. Indic. 2017, 72, 234–238. [Google Scholar] [CrossRef]
  166. Rocchini, D.; Bacaro, G.; Chirici, G.; Da Re, D.; Feilhauer, H.; Foody, G.M.; Galluzzi, M.; Garzon-Lopez, C.X.; Gillespie, T.W.; He, K.S.; et al. Remotely sensed spatial heterogeneity as an exploratory tool for taxonomic and functional diversity study. Ecol. Indic. 2018, 85, 983–990. [Google Scholar] [CrossRef]
  167. Steele-Dunne, S.C.; McNairn, H.; Monsivais-Huertero, A.; Judge, J.; Liu, P.W.; Papathanassiou, K. Radar Remote Sensing of Agricultural Canopies: A Review. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2017, 10, 2249–2273. [Google Scholar] [CrossRef]
  168. Howison, R.A.; Piersma, T.; Kentie, R.; Hooijmeijer, J.C.E.W.; Olff, H. Quantifying landscape–level land–use intensity patterns through radar–based remote sensing. J. Appl. Ecol. 2018, 55, 1276–1287. [Google Scholar] [CrossRef]
  169. Herrero-Huerta, M.; Bucksch, A.; Puttonen, E.; Rainey, K.M. Canopy roughness: A new phenotypic trait to estimate aboveground biomass from unmanned aerial system. Plant Phenomics 2020, 2020, 6735967. [Google Scholar] [CrossRef]
  170. Alfieri, J.G.; Kustas, W.P.; Nieto, H.; Prueger, J.H.; Hipps, L.E.; McKee, L.G.; Gao, F.; Los, S. Influence of wind direction on the surface roughness of vineyards. Irrig. Sci. 2019, 37, 359–373. [Google Scholar] [CrossRef]
  171. Singh, A.; Gaurav, K.; Rai, A.K.; Beg, Z. Machine learning to estimate surface roughness from satellite images. Remote Sens. 2021, 13, 3794. [Google Scholar] [CrossRef]
  172. Turner, R.; Panciera, R.; Tanase, M.A.; Lowell, K.; Hacker, J.M.; Walker, J.P. Estimation of soil surface roughness of agricultural soils using airborne LiDAR. Remote Sens. Environ. 2014, 140, 107–117. [Google Scholar] [CrossRef]
  173. Mahlayeye, M.; Darvishzadeh, R.; Nelson, A. Cropping Patterns of Annual Crops: A Remote Sensing Review. Remote Sens. 2022, 14, 2404. [Google Scholar] [CrossRef]
  174. Qiu, B.; Hu, X.; Chen, C.; Tang, Z.; Yang, P.; Zhu, X.; Yan, C.; Jian, Z. Maps of cropping patterns in China during 2015–2021. Sci. Data 2022, 9, 479. [Google Scholar] [CrossRef]
  175. El Hajj, M.; Bégué, A.; Guillaume, S.; Martiné, J.F. Integrating SPOT-5 time series, crop growth modeling and expert knowledge for monitoring agricultural practices—The case of sugarcane harvest on Reunion Island. Remote Sens. Environ. 2009, 113, 2052–2061. [Google Scholar] [CrossRef]
  176. Maponya, M.G.; van Niekerk, A.; Mashimbye, Z.E. Pre-harvest classification of crop types using a Sentinel-2 time-series and machine learning. Comput. Electron. Agric. 2020, 169, 105164. [Google Scholar] [CrossRef]
  177. Ahlawat, I.; Sheoran, H.S.; Roohi; Dahiya, G.; Sihag, P. Analysis of sentinel-1 data for regional crop classification: A multi-data approach for rabi crops of district Hisar (Haryana). J. Appl. Nat. Sci. 2020, 12, 165–170. [Google Scholar] [CrossRef]
  178. Orynbaikyzy, A.; Gessner, U.; Conrad, C. Crop type classification using a combination of optical and radar remote sensing data: A review. Int. J. Remote Sens. 2019, 40, 6553–6595. [Google Scholar] [CrossRef]
  179. Zhao, J.; Zhong, Y.; Hu, X.; Wei, L.; Zhang, L. A robust spectral-spatial approach to identifying heterogeneous crops using remote sensing imagery with high spectral and spatial resolutions. Remote Sens. Environ. 2020, 239, 111605. [Google Scholar] [CrossRef]
  180. Nigam, R.; Tripathy, R.; Dutta, S.; Bhagia, N.; Nagori, R.; Chandrasekar, K.; Kot, R.; Bhattacharya, B.K.; Ustin, S. Crop type discrimination and health assessment using hyperspectral imaging. Curr. Sci. 2019, 116, 1108–1123. [Google Scholar] [CrossRef]
  181. Prins, A.J.; Van Niekerk, A. Crop type mapping using LiDAR, Sentinel-2 and aerial imagery with machine learning algorithms. Geo-Spat. Inf. Sci. 2020, 24, 215–227. [Google Scholar] [CrossRef]
  182. He, T.; Zhang, M.; Xiao, W.; Zhai, G.; Fang, K.; Chen, Y.; Wu, C. Terend and potential enhancement of cropping intensity. Comput. Electron. Agric. 2025, 229, 109777. [Google Scholar] [CrossRef]
  183. Du, G.; Han, L.; Yao, L.; Faye, B. Spatiotemporal Dynamics and Evolution of Grain Cropping Patterns in Northeast China: Insights from Remote Sensing and Spatial Overlay Analysis. Agriculture 2024, 14, 1443. [Google Scholar] [CrossRef]
  184. Manjunath, K.R.; Kundu, N.; Ray, S.S.; Panigrahy, S.; Parihar, J.S. Cropping Systems Dynamics in the Lower Gangetic Plains of India using Geospatial Technologies. Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 2012, 38, 40–45. [Google Scholar] [CrossRef]
  185. Manjunath, K.R.; More, R.S.; Jain, N.K.; Panigrahy, S.; Parihar, J.S. Mapping of rice-cropping pattern and cultural type using remote-sensing and ancillary data: A case study for South and Southeast Asian countries. Int. J. Remote Sens. 2015, 36, 6008–6030. [Google Scholar] [CrossRef]
  186. Aduvukha, G.R.; Abdel-Rahman, E.M.; Sichangi, A.W.; Makokha, G.O.; Landmann, T.; Mudereri, B.T.; Tonnang, H.E.Z.; Dubois, T. Cropping pattern mapping in an agro-natural heterogeneous landscape using sentinel-2 and sentinel-1 satellite datasets. Agric. 2021, 11, 530. [Google Scholar] [CrossRef]
  187. Zhang, M.; Wu, B.; Zeng, H.; He, G.; Liu, C.; Tao, S.; Zhang, Q.; Nabil, M.; Tian, F.; Bofana, J.; et al. GCI30: A global dataset of 30m cropping intensity using multisource remote sensing imagery. Earth Syst. Sci. Data 2021, 13, 4799–4817. [Google Scholar] [CrossRef]
  188. Lange, M.; Feilhauer, H.; Kühn, I.; Doktor, D. Mapping land-use intensity of grasslands in Germany with machine learning and Sentinel-2 time series. Remote Sens. Environ. 2022, 277, 112888. [Google Scholar] [CrossRef]
  189. Griffiths, P.; Nendel, C.; Pickert, J.; Hostert, P. Towards national-scale characterization of grassland use intensity from integrated Sentinel-2 and Landsat time series. Remote Sens. Environ. 2020, 238, 111124. [Google Scholar] [CrossRef]
  190. Fischer, M.; Bossdorf, O.; Gockel, S.; Hänsel, F.; Hemp, A.; Hessenmöller, D.; Korte, G.; Nieschulze, J.; Pfeiffer, S.; Prati, D.; et al. Implementing large-scale and long-term functional biodiversity research: The Biodiversity Exploratories. Basic Appl. Ecol. 2010, 11, 473–485. [Google Scholar] [CrossRef]
  191. Sousa Júnior, V.d.P.; Sparacino, J.; Espindola, G.M.d.; Assis, R.J.S.d. Carbon Biomass Estimation Using Vegetation Indices in Agriculture–Pasture Mosaics in the Brazilian Caatinga Dry Tropical Forest. ISPRS Int. J. Geo-Inf. 2023, 12, 354. [Google Scholar] [CrossRef]
  192. Kumar, L.; Mutanga, O. Remote sensing of above-ground biomass. Remote Sens. 2017, 9, 935. [Google Scholar] [CrossRef]
  193. Zhang, P.; Lu, B.; Ge, J.; Wang, X.; Yang, Y.; Shang, J.; La, Z.; Zang, H.; Zeng, Z. Using UAV-based multispectral and RGB imagery to monitor above-ground biomass of oat-based diversified cropping. Eur. J. Agron. 2025, 162, 127422. [Google Scholar] [CrossRef]
  194. Da, H.; Li, Y.; Xu, L.; Wang, S.; Hu, L.; Hu, Z.; Wei, Q.; Zhu, R.; Chen, Q.; Xin, D.; et al. Advancing soybean biomass estimation through multi-source UAV data fusion and machine learning algorithms. Smart Agric. Technol. 2025, 10, 100778. [Google Scholar] [CrossRef]
  195. Breunig, F.M.; Dalagnol, R.; Galvão, L.S.; Bispo, P.d.C.; Liu, Q.; Berra, E.F.; Gaida, W.; Liesenberg, V.; Sampaio, T.V.M. Monitoring Cover Crop Biomass in Southern Brazil Using Combined PlanetScope and Sentinel-1 SAR Data. Remote Sens. 2024, 16, 2686. [Google Scholar] [CrossRef]
  196. Hagn, L.; Schuster, J.; Mittermayer, M.; Hülsbergen, K.J. A new method for satellite-based remote sensing analysis of plant-specific biomass yield patterns for precision farming applications. Precis. Agric. 2024, 25, 2801–2830. [Google Scholar] [CrossRef]
  197. Burger, R.; Aouizerats, B.; den Besten, N.; Guillevic, P.; Catarino, F.; van der Horst, T.; Jackson, D.; Koopmans, R.; Ridderikhoff, M.; Robson, G.; et al. The Biomass Proxy: Unlocking Global Agricultural Monitoring through Fusion of Sentinel-1 and Sentinel-2. Remote Sens. 2024, 16, 835. [Google Scholar] [CrossRef]
  198. Bloem, E.; Gerighausen, H.; Chen, X.; Schnug, E. The potential of spectral measurements for identifying glyphosate application to agricultural fields. Agronomy 2020, 10, 1409. [Google Scholar] [CrossRef]
  199. Niu, Z.; Rehman, T.; Young, J.; Johnson, W.G.; Yokoo, T.; Young, B.; Jin, J. Hyperspectral Analysis for Discriminating Herbicide Site of Action: A Novel Approach for Accelerating Herbicide Research. Sensors 2023, 23, 9300. [Google Scholar] [CrossRef] [PubMed]
  200. Zhang, T.; Huang, Y.; Reddy, K.N.; Yang, P.; Zhao, X.; Zhang, J. Using Machine Learning and Hyperspectral Images to Assess Damages to Corn Plant Caused by Glyphosate and to Evaluate Recoverability. Agronomy 2021, 11, 583. [Google Scholar] [CrossRef]
  201. Chu, H.; Zhang, C.; Wang, M.; Gouda, M.; Wei, X.; He, Y.; Liu, Y. Hyperspectral imaging with shallow convolutional neural networks (SCNN) predicts the early herbicide stress in wheat cultivars. J. Hazard. Mater. 2022, 421, 126706. [Google Scholar] [CrossRef]
  202. Xiao, T.; Yang, L.; He, X.; Wang, L.; Zhang, D.; Cui, T.; Zhang, K.; Bao, L.; An, S.; Zhang, X. A green and efficient method for detecting nicosulfuron residues in field maize using hyperspectral imaging and deep learning. J. Hazard. Mater. 2025, 484, 136724. [Google Scholar] [CrossRef] [PubMed]
  203. Pause, M.; Raasch, F.; Marrs, C.; Csaplovics, E. Csaplovics Monitoring Glyphosate-Based Herbicide Treatment Using Sentinel-2 Time Series—A Proof-of-Principle. Remote Sens. 2019, 11, 2541. [Google Scholar] [CrossRef]
  204. Pon Arasan, A.; Radhamani, S.; Pazhanivelan, S.; Kavitha, R.; Raja, R.; Kumaraperumal, R. Mapping and monitoring of weeds using unmanned aircraft systems and remote sensing. Plant Prot. Sci. 2024, 61, 44–55. [Google Scholar] [CrossRef]
  205. Li, J.; Ge, Y.; Puntel, L.A.; Heeren, D.M.; Bai, G.; Balboa, G.R.; Gamon, J.A.; Arkebauer, T.J.; Shi, Y. Devising optimized maize nitrogen stress indices in complex field conditions from UAV hyperspectral imagery. Precis. Agric. 2024, 26, 3. [Google Scholar] [CrossRef]
  206. Yin, H.; Li, F.; Yang, H.; Di, Y.; Hu, Y.; Yu, K. Mapping Plant Nitrogen Concentration and Aboveground Biomass of Potato Crops from Sentinel-2 Data Using Ensemble Learning Models. Remote Sens. 2024, 16, 349. [Google Scholar] [CrossRef]
  207. Almawazreh, A.; Buerkert, A.; Vazhacharickal, P.J.; Peth, S. Assessing canopy temperature responses to nitrogen fertilisation in South Indian crops using UAV -based thermal sensing. Int. J. Remote Sens. 2025, 4, 2389–2417. [Google Scholar] [CrossRef]
  208. Hossen, M.A.; Diwakar, P.K.; Ragi, S. Total nitrogen estimation in agricultural soils via aerial multispectral imaging and LIBS. Sci. Rep. 2021, 11, 12693. [Google Scholar] [CrossRef] [PubMed]
  209. Tiessen, H.; Cuevas, E.; Chacon, P. The role of soil organic matter in sustaining soil fertility. Nature 1994, 371, 783–785. [Google Scholar] [CrossRef]
  210. Castaldi, F.; Chabrillat, S.; Jones, A.; Vreys, K.; Bomans, B.; van Wesemael, B. Soil organic carbon estimation in croplands by hyperspectral remote APEX data using the LUCAS topsoil database. Remote Sens. 2018, 10, 153. [Google Scholar] [CrossRef]
  211. Wilcox, C.H.; Frazier, B.E.; Ball, S.T. Relationship between soil organic carbon and landsat tm data in eastern Washington. Photogramm. Eng. Rem. Sens 1994, 60, 777–781. [Google Scholar]
  212. Batjes, N.H. Total carbon and nitrogen in the soils of the world. Eur. J. Soil Sci. 1996, 47, 151–163. [Google Scholar] [CrossRef]
  213. Basnyat, P.; McConkey, B.; Meinert, B.; Gatkze, C.; Noble, G. Agriculture Field Characterization Using Aerial Photograph and Satellite Imagery. IEEE Geosci. Remote Sens. Lett. 2004, 1, 7–10. [Google Scholar] [CrossRef]
  214. Nguyen, T.T.; Pham, T.D.; Nguyen, C.T.; Delfos, J.; Archibald, R.; Dang, K.B.; Hoang, N.B.; Guo, W.; Ngo, H.H. A novel intelligence approach based active and ensemble learning for agricultural soil organic carbon prediction using multispectral and SAR data fusion. Sci. Total Environ. 2022, 804, 150187. [Google Scholar] [CrossRef]
  215. Zhou, T.; Lv, W.; Geng, Y.; Xiao, S.; Chen, J.; Xu, X.; Pan, J.; Si, B.; Lausch, A. National-scale spatial prediction of soil organic carbon and total nitrogen using long-term optical and microwave satellite observations in Google Earth Engine. Comput. Electron. Agric. 2023, 210, 107928. [Google Scholar] [CrossRef]
  216. Castaldi, F.; Halil Koparan, M.; Wetterlind, J.; Žydelis, R.; Vinci, I.; Özge Savaş, A.; Kıvrak, C.; Tunçay, T.; Volungevičius, J.; Obber, S.; et al. Assessing the capability of Sentinel-2 time-series to estimate soil organic carbon and clay content at local scale in croplands. ISPRS J. Photogramm. Remote Sens. 2023, 199, 40–60. [Google Scholar] [CrossRef]
  217. Wang, X.; Zhang, Y.; Atkinson, P.M.; Yao, H. Predicting soil organic carbon content in Spain by combining Landsat TM and ALOS PALSAR images. Int. J. Appl. Earth Obs. Geoinf. 2020, 92, 102182. [Google Scholar] [CrossRef]
  218. Zhou, T.; Geng, Y.; Chen, J.; Pan, J.; Haase, D.; Lausch, A. High-resolution digital mapping of soil organic carbon and soil total nitrogen using DEM derivatives, Sentinel-1 and Sentinel-2 data based on machine learning algorithms. Sci. Total Environ. 2020, 729, 138244. [Google Scholar] [CrossRef] [PubMed]
  219. Li, H.; Zhang, C.; Zhang, S.; Atkinson, P.M. Full year crop monitoring and separability assessment with fully-polarimetric L-band UAVSAR: A case study in the Sacramento Valley, California. Int. J. Appl. Earth Obs. Geoinf. 2019, 74, 45–56. [Google Scholar] [CrossRef]
  220. Ottinger, M.; Kuenzer, C. Spaceborne L-Band Synthetic Aperture Radar Data for Geoscientific Analyses in Coastal Land Applications: A Review. Remote Sens. 2020, 12, 2228. [Google Scholar] [CrossRef]
  221. Shafizadeh-Moghadam, H.; Minaei, F.; Talebi-khiavi, H.; Xu, T.; Homaee, M. Synergetic use of multi-temporal Sentinel-1, Sentinel-2, NDVI, and topographic factors for estimating soil organic carbon. Catena 2022, 212, 106077. [Google Scholar] [CrossRef]
  222. Wang, L.; Wang, X.; Wang, D.; Qi, B.; Zheng, S.; Liu, H.; Luo, C.; Li, H.; Meng, L.; Meng, X.; et al. Spatiotemporal Changes and Driving Factors of Cultivated Soil Organic Carbon in Northern China’s Typical Agro-Pastoral Ecotone in the Last 30 Years. Remote Sens. 2021, 13, 3607. [Google Scholar] [CrossRef]
  223. Odebiri, O.; Mutanga, O.; Odindi, J.; Naicker, R. Modelling soil organic carbon stock distribution across different land-uses in South Africa: A remote sensing and deep learning approach. ISPRS J. Photogramm. Remote Sens. 2022, 188, 351–362. [Google Scholar] [CrossRef]
  224. Wang, X.; Zeng, H.; Yang, X.; Shu, J.; Wu, Q.; Que, Y.; Yang, X.; Yi, X.; Khalil, I.; Zomaya, A.Y. Remote sensing revolutionizing agriculture: Toward a new frontier. Futur. Gener. Comput. Syst. 2025, 166, 107691. [Google Scholar] [CrossRef]
  225. Zhu, X.; Cai, F.; Tian, J.; Williams, T.K.A. Spatiotemporal fusion of multisource remote sensing data: Literature survey, taxonomy, principles, applications, and future directions. Remote Sens. 2018, 10, 527. [Google Scholar] [CrossRef]
  226. Wang, D.; Cao, W.; Zhang, F.; Li, Z.; Xu, S.; Wu, X. A Review of Deep Learning in Multiscale Agricultural Sensing. Remote Sens. 2022, 14, 559. [Google Scholar] [CrossRef]
  227. Osco, L.P.; Junior, J.M.; Ramos, A.P.M.; Furuya, D.E.G.; Santana, D.C.; Teodoro, L.P.R.; Gonçalves, W.N.; Baio, F.H.R.; Pistori, H.; Junior, C.A.d.S.; et al. Leaf nitrogen concentration and plant height prediction for maize using UAV-based multispectral imagery and machine learning techniques. Remote Sens. 2020, 12, 3237. [Google Scholar] [CrossRef]
  228. Reda, R.; Saffaj, T.; Ilham, B.; Saidi, O.; Issam, K.; Brahim, L.; El Hadrami, E.M. A comparative study between a new method and other machine learning algorithms for soil organic carbon and total nitrogen prediction using near infrared spectroscopy. Chemom. Intell. Lab. Syst. 2019, 195, 103873. [Google Scholar] [CrossRef]
  229. Castaldi, F.; Hueni, A.; Chabrillat, S.; Ward, K.; Buttafuoco, G.; Bomans, B.; Vreys, K.; Brell, M.; Wesemael, B. Van ISPRS Journal of Photogrammetry and Remote Sensing Evaluating the capability of the Sentinel 2 data for soil organic carbon prediction in croplands. ISPRS J. Photogramm. Remote Sens. 2019, 147, 267–282. [Google Scholar] [CrossRef]
  230. Shi, P.; Wang, Y.; Xu, J.; Zhao, Y.; Yang, B.; Yuan, Z.; Sun, Q. Rice nitrogen nutrition estimation with RGB images and machine learning methods. Comput. Electron. Agric. 2021, 180, 105860. [Google Scholar] [CrossRef]
  231. Sahabiev, I.; Smirnova, E.; Giniyatullin, K. Spatial prediction of agrochemical properties on the scale of a single field using machine learning methods based on remote sensing data. Agronomy 2021, 11, 2266. [Google Scholar] [CrossRef]
  232. Wolanin, A.; Mateo-Garciá, G.; Camps-Valls, G.; Gómez-Chova, L.; Meroni, M.; Duveiller, G.; Liangzhi, Y.; Guanter, L. Estimating and understanding crop yields with explainable deep learning in the Indian Wheat Belt. Environ. Res. Lett. 2020, 15, 024019. [Google Scholar] [CrossRef]
  233. Ferentinos, K.P. Deep learning models for plant disease detection and diagnosis. Comput. Electron. Agric. 2018, 145, 311–318. [Google Scholar] [CrossRef]
  234. Jaihuni, M.; Khan, F.; Lee, D.; Basak, J.K.; Bhujel, A.; Moon, B.E.; Park, J.; Kim, H.T. Determining Spatiotemporal Distribution of Macronutrients in a Cornfield Using Remote Sensing and a Deep Learning Model. IEEE Access 2021, 9, 30256–30266. [Google Scholar] [CrossRef]
  235. Gebbers, R.; Adamchuk, V.I. Precision Agriculture and Food Security. Science 2010, 327, 828–831. [Google Scholar] [CrossRef]
  236. Jiang, H.; Zhang, C.; Qiao, Y.; Zhang, Z.; Zhang, W.; Song, C. CNN feature based graph convolutional network for weed and crop recognition in smart farming. Comput. Electron. Agric. 2020, 174, 105450. [Google Scholar] [CrossRef]
  237. Khusro, S.; Jabeen, F.; Mashwani, S.R.; Alam, I. Linked open data: Towards the realization of semantic web-A review. Indian J. Sci. Technol. 2014, 7, 745–764. [Google Scholar] [CrossRef]
  238. Lausch, A.; Schmidt, A.; Tischendorf, L. Data mining and linked open data—New perspectives for data analysis in environmental research. Ecol. Modell. 2015, 295, 5–17. [Google Scholar] [CrossRef]
  239. Zhou, X.; Zhang, J.; Chen, D.; Huang, Y.; Kong, W.; Yuan, L.; Ye, H.; Huang, W. Assessment of Leaf Chlorophyll Content Models for Winter Wheat Using Landsat-8 Multispectral Remote Sensing Data. Remote Sens. 2020, 12, 2574. [Google Scholar] [CrossRef]
  240. Zolotukhina, A.; Machikhin, A.; Guryleva, A.; Gresis, V.; Kharchenko, A.; Dekhkanova, K.; Polyakova, S.; Fomin, D.; Nesterov, G.; Pozhar, V. Evaluation of Leaf Chlorophyll Content from Acousto-Optic Hyperspectral Data: A Multi-Crop Study. Remote Sens. 2024, 16, 1073. [Google Scholar] [CrossRef]
  241. Cheng, J.; Yang, H.; Qi, J.; Sun, Z.; Han, S.; Feng, H.; Jiang, J.; Xu, W.; Li, Z.; Yang, G.; et al. Estimating canopy-scale chlorophyll content in apple orchards using a 3D radiative transfer model and UAV multispectral imagery. Comput. Electron. Agric. 2022, 202, 107401. [Google Scholar] [CrossRef]
  242. Haboudane, D.; Miller, J.R.; Tremblay, N.; Zarco-Tejada, P.J.; Dextraze, L. Integrated narrow-band vegetation indices for prediction of crop chlorophyll content for application to precision agriculture. Remote Sens. Environ. 2002, 81, 416–426. [Google Scholar] [CrossRef]
  243. Qian, B.; Ye, H.; Huang, W.; Xie, Q.; Pan, Y.; Xing, N.; Ren, Y.; Guo, A.; Jiao, Q.; Lan, Y. A sentinel-2-based triangular vegetation index for chlorophyll content estimation. Agric. For. Meteorol. 2022, 322, 109000. [Google Scholar] [CrossRef]
  244. Holtgrave, A.K.; Röder, N.; Ackermann, A.; Erasmi, S.; Kleinschmit, B. Comparing Sentinel-1 and -2 data and indices for agricultural land use monitoring. Remote Sens. 2020, 12, 2919. [Google Scholar] [CrossRef]
  245. Wu, Q.; Zhang, Y.; Zhao, Z.; Xie, M.; Hou, D. Estimation of Relative Chlorophyll Content in Spring Wheat Based on Multi-Temporal UAV Remote Sensing. Agronomy 2023, 13, 211. [Google Scholar] [CrossRef]
  246. Gitelson, A.A.; Viña, A.; Solovchenko, A. Spectral response of gross primary production to in situ canopy light absorption coefficient of chlorophyll. Photosynth. Res. 2025, 163, 20. [Google Scholar] [CrossRef]
  247. Gitelson, A.; Arkebauer, T.; Solovchenko, A.; Nguy-Robertson, A.; Inoue, Y. An insight into spectral composition of light available for photosynthesis via remotely assessed absorption coefficient at leaf and canopy levels. Photosynth. Res. 2022, 151, 47–60. [Google Scholar] [CrossRef]
  248. Abdelbaki, A.; Schlerf, M.; Retzlaff, R.; Machwitz, M.; Verrelst, J.; Udelhoven, T. Comparison of crop trait retrieval strategies using UAV-based VNIR hyperspectral imaging. Remote Sens. 2021, 13, 1748. [Google Scholar] [CrossRef]
  249. Blackburn, G.A. Hyperspectral remote sensing of plant pigments. J. Exp. Bot. 2007, 58, 855–867. [Google Scholar] [CrossRef]
  250. Lee, S.; Ghimire, A.; Kim, Y.; Lee, J.D. Automatic optimization of regions of interest in hyperspectral images for detecting vegetative indices in soybeans. Front. Plant Sci. 2025, 16, 1511646. [Google Scholar] [CrossRef]
  251. Gitelson, A.; Solovchenko, A.; Viña, A. Foliar absorption coefficient derived from reflectance spectra: A gauge of the efficiency of in situ light-capture by different pigment groups. J. Plant Physiol. 2020, 254, 153277. [Google Scholar] [CrossRef] [PubMed]
  252. Shen, J.; Huang, Y.; Chen, W.; Li, M.; Tan, W.; Wang, R.; Deng, Y.; Gong, Y.; Ai, S.; Liu, N. Assessing the Transferability of Models for Predicting Foliar Nutrient Concentrations Across Maize Cultivars. Remote Sens. 2025, 17, 652. [Google Scholar] [CrossRef]
  253. Kharel, T.P.; Tyler, H.L.; Mubvumba, P.; Huang, Y.; Bhandari, A.B.; Fletcher, R.S.; Anapalli, S.; Joshi, D.R.; Mengistu, A.; Birru, G.; et al. Machine learning on multi-spectral imagery to estimate nutrient yield of mixed-species cover crops. Agric. Environ. Lett. 2025, 10, e70009. [Google Scholar] [CrossRef]
  254. Geng, G.; Gu, Q.; Zhou, H.; Zhang, B.; He, Z.; Zheng, R. Random forest model that incorporates solar-induced chlorophyll fluorescence data can accurately track crop yield variations under drought conditions. Ecol. Inform. 2025, 85, 102972. [Google Scholar] [CrossRef]
  255. Wu, Y.; Zhang, Z.; Wu, L.; Zhang, Y. Solar-induced chlorophyll fluorescence and its relationship with photosynthesis during waterlogging in a maize field. Agric. For. Meteorol. 2025, 363, 110404. [Google Scholar] [CrossRef]
  256. Li, Z.; Zhang, Q.; Li, J.; Yang, X.; Wu, Y.; Zhang, Z.; Wang, S.; Wang, H.; Zhang, Y. Solar-induced chlorophyll fluorescence and its link to canopy photosynthesis in maize from continuous ground measurements. Remote Sens. Environ. 2020, 236, 111420. [Google Scholar] [CrossRef]
  257. Song, L.; Cai, J.; Wu, K.; Li, Y.; Hou, G.; Du, S.; Duan, J.; He, L.; Guo, T.; Feng, W. Early diagnosis of wheat powdery mildew using solar-induced chlorophyll fluorescence and hyperspectral reflectance. Eur. J. Agron. 2025, 162, 127427. [Google Scholar] [CrossRef]
  258. Jin, X.; Zarco-Tejada, P.J.; Schmidhalter, U.; Reynolds, M.P.; Hawkesford, M.J.; Varshney, R.K.; Yang, T.; Nie, C.; Li, Z.; Ming, B.; et al. High-Throughput Estimation of Crop Traits: A Review of Ground and Aerial Phenotyping Platforms. IEEE Geosci. Remote Sens. Mag. 2021, 9, 200–231. [Google Scholar] [CrossRef]
  259. Camino, C.; González-Dugo, V.; Hernández, P.; Sillero, J.C.; Zarco-Tejada, P.J. Improved nitrogen retrievals with airborne-derived fluorescence and plant traits quantified from VNIR-SWIR hyperspectral imagery in the context of precision agriculture. Int. J. Appl. Earth Obs. Geoinf. 2018, 70, 105–117. [Google Scholar] [CrossRef]
  260. Gerhards, M.; Schlerf, M.; Mallick, K.; Udelhoven, T. Challenges and future perspectives of multi-/Hyperspectral thermal infrared remote sensing for crop water-stress detection: A review. Remote Sens. 2019, 11, 1240. [Google Scholar] [CrossRef]
  261. Abbas, T.; Farooq, N.; Nadeem, M.A. Application of leaf water content measurement to improve herbicide efficacy for effective weed management in a changing climate—A review. Crop Prot. 2025, 191, 107138. [Google Scholar] [CrossRef]
  262. Cardamis, M.; Jia, H.; Qian, H.; Chen, W.; Yan, Y.; Ghannoum, O.; Quigley, A.; Chou, C.T.; Hu, W. Leafeon: Towards Accurate Sensing of Leaf Water Content for Protected Cropping, with mmWave Radar. IEEE Internet Things J. 2025, 12, 19646–19659. [Google Scholar] [CrossRef]
  263. Dong, H.; Dong, J.; Sun, S.; Bai, T.; Zhao, D.; Yin, Y.; Shen, X.; Wang, Y.; Zhang, Z.; Wang, Y. Crop water stress detection based on UAV remote sensing systems. Agric. Water Manag. 2024, 303, 109059. [Google Scholar] [CrossRef]
  264. Ahmad, U.; Alvino, A.; Marino, S. A Review of Crop Water Stress Assessment Using Remote Sensing. Remote Sens. 2021, 13, 4155. [Google Scholar] [CrossRef]
  265. Romero-Trigueros, C.; Bayona Gambín, J.M.; Nortes Tortosa, P.A.; Alarcón Cabañero, J.J.; Nicolás, E.N. Determination of Crop Water Stress index by infrared thermometry in grapefruit trees irrigated with saline reclaimed water combined with deficit irrigation. Remote Sens. 2019, 11, 757. [Google Scholar] [CrossRef]
  266. Jiang, M.; Zheng, C.; Jia, L.; Chen, J. A 20-year dataset (2001—2020) of global cropland water-use efficiency at 1-km grid resolution. Sci. Data 2025, 12, 574. [Google Scholar] [CrossRef] [PubMed]
  267. Cheng, M.; Yin, D.; Wu, W.; Cui, N.; Nie, C.; Shi, L.; Liu, S.; Yu, X.; Bai, Y.; Liu, Y.; et al. A review of remote sensing estimation of crop water productivity: Definition, methodology, scale, and evaluation. Int. J. Remote Sens. 2023, 44, 5033–5068. [Google Scholar] [CrossRef]
  268. Safi, A.R.; Karimi, P.; Mul, M.; Chukalla, A.; de Fraiture, C. Translating open-source remote sensing data to crop water productivity improvement actions. Agric. Water Manag. 2022, 261, 107373. [Google Scholar] [CrossRef]
  269. Koetz, B.; Bastiaanssen, W.; Berger, M.; Defourney, P.; Del Bello, U.; Drusch, M.; Drinkwater, M.; Duca, R.; Fernandez, V.; Ghent, D.; et al. High spatio-temporal resolution land surface temperature mission—A copernicus candidate mission in support of agricultural monitoring. In Proceedings of the IEEE International Geoscience and Remote Sensing Symposium, Valencia, Spain, 22–27 July 2018; pp. 8160–8162. [Google Scholar]
  270. Heinemann, S.; Siegmann, B.; Thonfeld, F.; Muro, J.; Jedmowski, C.; Kemna, A.; Kraska, T.; Muller, O.; Schultz, J.; Udelhoven, T.; et al. Land surface temperature retrieval for agricultural areas using a novel UAV platform equipped with a thermal infrared and multispectral sensor. Remote Sens. 2020, 12, 1075. [Google Scholar] [CrossRef]
  271. Lee, J. Estimating Near-Surface Air Temperature from Satellite-Derived Land Surface Temperature Using Temporal Deep Learning: A Comparative Analysis. IEEE Access 2025, 13, 28935–28945. [Google Scholar] [CrossRef]
  272. Majumder, A.; Kingra, P.K.; Setia, R.; Singh, S.P.; Pateriya, B. Influence of land use/land cover changes on surface temperature and its effect on crop yield in different agro-climatic regions of Indian Punjab. Geocarto Int. 2020, 35, 663–686. [Google Scholar] [CrossRef]
  273. Barai, K.; Wallhead, M.; Hall, B.; Rahimzadeh-Bajgiran, P.; Meireles, J.; Herrmann, I.; Zhang, Y.J. Detecting spatial variation in wild blueberry water stress using UAV-borne thermal imagery: Distinct temporal and reference temperature effects. Precis. Agric. 2025, 26, 25. [Google Scholar] [CrossRef]
  274. García-Santos, V.; Sánchez, J.M.; Cuxart, J. Evapotranspiration Acquired with Remote Sensing Thermal-Based Algorithms: A State-of-the-Art Review. Remote Sens. 2022, 14, 3440. [Google Scholar] [CrossRef]
  275. El Hazdour, I.; Le Page, M.; Hanich, L.; Chakir, A.; Lopez, O.; Jarlan, L. A GEE TSEB workflow for daily high-resolution fully remote sensing evapotranspiration: Validation over four crops in semi-arid conditions and comparison with the SSEBop experimental product. Environ. Model. Softw. 2025, 187, 106365. [Google Scholar] [CrossRef]
  276. Bhattarai, N.; Wagle, P. Recent advances in remote sensing of evapotranspiration. Remote Sens. 2021, 13, 4260. [Google Scholar] [CrossRef]
  277. Pelosi, A.; Villani, P.; Bolognesi, S.F.; Chirico, G.B.; D’urso, G. Predicting crop evapotranspiration by integrating ground and remote sensors with air temperature forecasts. Sensors 2020, 20, 1740. [Google Scholar] [CrossRef] [PubMed]
  278. Ippolito, M.; De Caro, D.; Ciraolo, G.; Minacapilli, M.; Provenzano, G. Estimating crop coefficients and actual evapotranspiration in citrus orchards with sporadic cover weeds based on ground and remote sensing data. Irrig. Sci. 2023, 41, 5–22. [Google Scholar] [CrossRef]
  279. Wolff, W.; Francisco, J.P.; Flumignan, D.L.; Marin, F.R.; Folegatti, M.V. Optimized algorithm for evapotranspiration retrieval via remote sensing. Agric. Water Manag. 2022, 262, 107390. [Google Scholar] [CrossRef]
  280. Garofalo, S.P.; Ardito, F.; Sanitate, N.; De Carolis, G.; Ruggieri, S.; Giannico, V.; Rana, G.; Ferrara, R.M. Robustness of Actual Evapotranspiration Predicted by Random Forest Model Integrating Remote Sensing and Meteorological Information: Case of Watermelon (Citrullus lanatus, (Thunb.) Matsum. & Nakai, 1916). Water 2025, 17, 323. [Google Scholar] [CrossRef]
  281. Ratshiedana, P.E.; Elbasit, M.A.M.; Adam, E.; Chirima, G.J. Evaluation of remote sensing algorithms for estimating actual evapotranspiration in arid agricultural environments. Hydrol. Earth Syst. Sci. 2025. [Google Scholar] [CrossRef]
  282. Liu, M.; Lei, H.; Wang, X.; Paredes, P. High-resolution mapping of evapotranspiration over heterogeneous cropland affected by soil salinity. Agric. Water Manag. 2025, 308, 109301. [Google Scholar] [CrossRef]
  283. Kaihotsu, I.; Asanuma, J.; Aida, K.; Oyunbaatar, D. Evaluation of the AMSR2 L2 soil moisture product of JAXA on the Mongolian Plateau over seven years (2012–2018). SN Appl. Sci. 2019, 1, 1477. [Google Scholar] [CrossRef]
  284. Kolassa, J.; Gentine, P.; Prigent, C.; Aires, F. Soil moisture retrieval from AMSR-E and ASCAT microwave observation synergy. Part 1: Satellite data analysis. Remote Sens. Environ. 2016, 173, 202–217. [Google Scholar] [CrossRef]
  285. NISAR. NISAR: The NASA-ISRO SAR Mission. Water: Vital for Life and Civilization. © 2019 California Institute of Technology. Government Sponsorship Acknowledged. Available online: https://nisar.jpl.nasa.gov/system/documents/files/15_NISARApplications_SoilMoisture1.pdf (accessed on 19 October 2025).
  286. Benninga, H.J.F.; van der Velde, R.; Su, Z. Sentinel-1 soil moisture content and its uncertainty over sparsely vegetated fields. J. Hydrol. X 2020, 9, 100066. [Google Scholar] [CrossRef]
  287. Abbaszadeh, P.; Moradkhani, H.; Gavahi, K.; Kumar, S.; Hain, C.; Zhan, X.; Duan, Q.; Peters-Lidard, C.; Karimiziarani, S. High-resolution smap satellite soil moisture product: Exploring the opportunities. Bull. Am. Meteorol. Soc. 2021, 102, 309–315. [Google Scholar] [CrossRef]
  288. Noory, H.; Khoshsima, M.; Tsunekawa, A.; Tsubo, M.; Haregeweyn, N.; Pashapour, S. Developing a method for root-zone soil moisture monitoring at the field scale using remote sensing and simulation modeling. Agric. Water Manag. 2025, 308, 109263. [Google Scholar] [CrossRef]
  289. Jalilvand, E.; Tajrishy, M.; Ghazi Zadeh Hashemi, S.A.; Brocca, L. Quantification of irrigation water using remote sensing of soil moisture in a semi-arid region. Remote Sens. Environ. 2019, 231, 111226. [Google Scholar] [CrossRef]
  290. Ghazali, M.F.; Wikantika, K.; Harto, A.B.; Kondoh, A. Generating soil salinity, soil moisture, soil pH from satellite imagery and its analysis. Inf. Process. Agric. 2020, 7, 294–306. [Google Scholar] [CrossRef]
  291. Chabrillat, S.; Ben-Dor, E.; Cierniewski, J.; Gomez, C.; Schmid, T.; van Wesemael, B. Imaging Spectroscopy for Soil Mapping and Monitoring; Springer: Dordrecht, The Netherlands, 2019; Volume 40, ISBN 0123456789. [Google Scholar]
  292. Alvino, A.; Marino, S. Remote sensing for irrigation of horticultural crops. Horticulturae 2017, 3, 40. [Google Scholar] [CrossRef]
  293. Mpakairi, K.S.; Dube, T.; Sibanda, M.; Mutanga, O. Leveraging remote sensing for optimised national scale agricultural water management in South Africa. Sci. Total Environ. 2025, 974, 179199. [Google Scholar] [CrossRef]
  294. Zhou, D.; Zheng, C.; Jia, L.; Menenti, M.; Lu, J.; Chen, Q. Evaluating the Performance of Irrigation Using Remote Sensing Data and the Budyko Hypothesis: A Case Study in Northwest China. Remote Sens. 2025, 17, 1085. [Google Scholar] [CrossRef]
  295. Pitoro, V.S.J.; Franco, J.R.; de Souza Correa, L.R.; Manjate, M.J.; Román, R.M.S. Evaluating Water Productivity and Efficiency in Irrigated Fields Using Remote Sensing: A Case Study in Mozambique’s Sugarcane Cultivation. Water Conserv. Sci. Eng. 2025, 10, 14. [Google Scholar] [CrossRef]
  296. Ahmad, U.; Sohel, F. Evaluating decision support systems for precision irrigation and water use efficiency. Digit. Eng. 2025, 4, 100038. [Google Scholar] [CrossRef]
  297. Akbar, M.U.; Mirchi, A.; Arshad, A.; Alian, S.; Mehata, M.; Taghvaeian, S.; Khodkar, K.; Kettner, J.; Datta, S.; Wagner, K. Multi-model ensemble mapping of irrigated areas using remote sensing, machine learning, and ground truth data. Agric. Water Manag. 2025, 312, 109416. [Google Scholar] [CrossRef]
  298. Zhai, W.; Cheng, Q.; Duan, F.; Huang, X.; Chen, Z. Remote sensing-based analysis of yield and water-fertilizer use efficiency in winter wheat management. Agric. Water Manag. 2025, 311, 109390. [Google Scholar] [CrossRef]
  299. Liu, Y.; Wang, Y.; Liao, Y.; Liao, R.; Šimůnek, J. Generating high-precision farmland irrigation pattern maps using remotely sensed ecological indices and machine learning algorithms. Agric. Water Manag. 2025, 308, 109302. [Google Scholar] [CrossRef]
  300. Ozdogan, M.; Yang, Y.; Allez, G.; Cervantes, C. Remote sensing of irrigated agriculture: Opportunities and challenges. Remote Sens. 2010, 2, 2274–2304. [Google Scholar] [CrossRef]
  301. Alordzinu, K.E.; Li, J.; Lan, Y.; Appiah, S.A.; Al Aasmi, A.; Wang, H.; Liao, J.; Sam-amoah, L.K.; Qiao, S. Ground-based hyperspectral remote sensing for estimating water stress in tomato growth in sandy loam and silty loam soils. Sensors 2021, 21, 5705. [Google Scholar] [CrossRef]
  302. Wang, X.; Guo, Z.; Zhang, K.; Fu, Z.; Lee, C.K.F.; Yang, D.; Detto, M.; Zhang, Y.; Wu, J. Can Large—Scale Satellite Products Track the Effects of Atmospheric Dryness and Soil Water Deficit on Ecosystem Productivity Under Droughts? Geophys. Res. Lett. 2025, 52, e2024GL110785. [Google Scholar] [CrossRef]
  303. Yu, J.; Wang, W.; Chen, Z.; Cao, M.; Qian, H. Disentangling the dominance of atmospheric and soil water stress on vegetation productivity in global drylands. J. Hydrol. 2025, 657, 133043. [Google Scholar] [CrossRef]
  304. Sankey, T.T. UAV hyperspectral-thermal-lidar fusion in phenotyping: Genetic trait differences among Fremont cottonwood populations. Landsc. Ecol. 2025, 40, 45. [Google Scholar] [CrossRef]
  305. Gitelson, A.; Arkebauer, T.; Viña, A.; Skakun, S.; Inoue, Y. Evaluating plant photosynthetic traits via absorption coefficient in the photosynthetically active radiation region. Remote Sens. Environ. 2021, 258, 112401. [Google Scholar] [CrossRef]
  306. Wang, X.; Lei, H.; Li, J.; Huo, Z.; Zhang, Y.; Qu, Y. Estimating evapotranspiration and yield of wheat and maize croplands through a remote sensing-based model. Agric. Water Manag. 2023, 282, 108294. [Google Scholar] [CrossRef]
  307. Koganti, T.; Ghane, E.; Martinez, L.R.; Iversen, B.V.; Allred, B.J. Mapping of agricultural subsurface drainage systems using unmanned aerial vehicle imagery and ground penetrating radar. Sensors 2021, 21, 2800. [Google Scholar] [CrossRef]
  308. Rahmani, S.R.; Schulze, D.G. Mapping subsurface tile lines on a research farm using aerial photography, paper maps, and expert knowledge. Agrosystems Geosci. Environ. 2023, 6, e20362. [Google Scholar] [CrossRef]
  309. Zhao, C.; Pan, Y.; Zhu, X.; Li, L.; Xia, X.; Ren, S.; Gao, Y. Monitoring of deforestation events in the tropics using multidimensional features of Sentinel 1 radar data. Front. For. Glob. Change 2023, 6, 1257806. [Google Scholar] [CrossRef]
  310. Tang, X.; Bratley, K.H.; Cho, K.; Bullock, E.L.; Olofsson, P.; Woodcock, C.E. Near real-time monitoring of tropical forest disturbance by fusion of Landsat, Sentinel-2, and Sentinel-1 data. Remote Sens. Environ. 2023, 294, 113626. [Google Scholar] [CrossRef]
  311. Zhao, F.; Sun, R.; Zhong, L.; Meng, R.; Huang, C.; Zeng, X.; Wang, M.; Li, Y.; Wang, Z. Monthly mapping of forest harvesting using dense time series Sentinel-1 SAR imagery and deep learning. Remote Sens. Environ. 2022, 269, 112822. [Google Scholar] [CrossRef]
  312. Bullock, E.L.; Woodcock, C.E.; Olofsson, P. Monitoring tropical forest degradation using spectral unmixing and Landsat time series analysis. Remote Sens. Environ. 2020, 238, 110968. [Google Scholar] [CrossRef]
  313. Li, Y.; Ye, Y.; Fang, X.; Zhang, C.; Zhao, Z. Loss of wetlands due to the expansion of polder in the Dongting Plain, China, AD 1368–1980. Holocene 2020, 30, 646–656. [Google Scholar] [CrossRef]
  314. Li, Y.; Dai, H.; Dai, Z.; Zhang, L. The polder systems legacies in the early twentieth century affect the contemporary landscape in the Jianghan Plain of Hubei, China. Herit. Sci. 2024, 12, 311. [Google Scholar] [CrossRef]
  315. Rosen, P.A.; Hensley, S.; Gurrola, E.; Rogez, F.; Chan, S.; Martin, J.; Rodriguez, E. SRTM C-band topographic data: Quality assessments and calibration activities. In Proceedings of the IGARSS 2001. Scanning the Present and Resolving the Future. Proceedings. IEEE 2001 International Geoscience and Remote Sensing Symposium (Cat. No.01CH37217), Sydney, Australia, 9–13 July 2001; Volume 2, pp. 739–741. [Google Scholar]
  316. Wulder, M.A.; Loveland, T.R.; Roy, D.P.; Crawford, C.J.; Masek, J.G.; Woodcock, C.E.; Allen, R.G.; Anderson, M.C.; Belward, A.S.; Cohen, W.B.; et al. Current status of Landsat program, science, and applications. Remote Sens. Environ. 2019, 225, 127–147. [Google Scholar] [CrossRef]
  317. Zink, M.; Fiedler, H.; Hajnsek, I.; Krieger, G.; Moreira, A.; Werner, M. The TanDEM-X Mission Concept. In Proceedings of the 2006 IEEE International Symposium on Geoscience and Remote Sensing, Denver, CO, USA, 31 July–4 August 2006; pp. 1938–1941. [Google Scholar]
  318. Wessel, B.; Huber, M.; Wohlfart, C.; Marschalk, U.; Kosmann, D.; Roth, A. Accuracy assessment of the global TanDEM-X Digital Elevation Model with GPS data. ISPRS J. Photogramm. Remote Sens. 2018, 139, 171–182. [Google Scholar] [CrossRef]
  319. Potin, P.; Rosich, B.; Roeder, J.; Bargellini, P. Sentinel-1 Mission operations concept. In Proceedings of the 2014 IEEE Geoscience and Remote Sensing Symposium, Munich, Germany, 22–27 July 2014; pp. 1465–1468. [Google Scholar]
  320. Kankaku, Y.; Suzuki, S.; OSAWA, Y. ALOS-2 mission and development status. In Proceedings of the 2013 IEEE International Geoscience and Remote Sensing Symposium—IGARSS, Melbourne, Australia, 21–26 July 2013; pp. 2396–2399. [Google Scholar]
  321. Hirano, A.; Welch, R.; Lang, H. Mapping from ASTER stereo image data: DEM validation and accuracy assessment. ISPRS J. Photogramm. Remote Sens. 2003, 57, 356–370. [Google Scholar] [CrossRef]
  322. Takaku, J.; Tadono, T. High resolution DSM generation from ALOS PRISM—Mosaic dataset-. In Proceedings of the 2012 IEEE International Geoscience and Remote Sensing Symposium, Munich, Germany, 22–27 July 2012; pp. 2687–2690. [Google Scholar]
  323. Rosenqvist, A.; Shimada, M.; Ito, N.; Watanabe, M. ALOS PALSAR: A pathfinder mission for global-scale monitoring of the environment. IEEE Trans. Geosci. Remote Sens. 2007, 45, 3307–3316. [Google Scholar] [CrossRef]
  324. Schutz, B.E.; Zwally, H.J.; Shuman, C.A.; Hancock, D.; DiMarzio, J.P. Overview of the ICESat mission. Geophys. Res. Lett. 2005, 32, L21S01. [Google Scholar] [CrossRef]
  325. Roy, Y.L.; Deschaux-Beaume, M. Sral, A Radar Altime Ter Designed to Meas Ure A Wide Range of Surface Types. Power 2009, 1, 445–448. [Google Scholar]
  326. Donnellan, A.; Parker, J.; Hensley, S.; Pierce, M.; Wang, J.; Rundle, J. UAVSAR observations of triggered slip on the Imperial, Superstition Hills, and East Elmore Ranch Faults associated with the 2010 M 7.2 El Mayor-Cucapah earthquake. Geochem. Geophys. Geosyst. 2014, 15, 815–829. [Google Scholar] [CrossRef]
  327. Dillner, R.P.; Wimmer, M.A.; Porten, M.; Udelhoven, T.; Retzlaff, R. Combining a Standardized Growth Class Assessment, UAV Sensor Data, GIS Processing, and Machine Learning Classification to Derive a Correlation with the Vigour and Canopy Volume of Grapevines. Sensors 2025, 25, 431. [Google Scholar] [CrossRef] [PubMed]
  328. Mulder, V.L.; de Bruin, S.; Schaepman, M.E.; Mayr, T.R. The use of remote sensing in soil and terrain mapping—A review. Geoderma 2011, 162, 1–19. [Google Scholar] [CrossRef]
  329. Kehoe, M.; Harding, A.; Pagdilao, S.J.; Appels, W.M. Effect of topographical and soil complexity on potato yields in irrigated fields. Agric. Water Manag. 2025, 307, 109216. [Google Scholar] [CrossRef]
  330. Perović, V.; Čakmak, D.; Jakšić, D.; Milanović, M.; Matić, M.; Pavlović, D.; Mitrović, M.; Pavlović, P. Development and evaluation approach of soil quality in agricultural soils: Integrated system for a more reliable delineation of homogeneous management zones. Appl. Soil Ecol. 2025, 206, 105860. [Google Scholar] [CrossRef]
  331. Tang, J.; Xie, Y.; Cheng, H.; Liu, G. Catena Impact of farmland landscape characteristics on gully erosion in the black soil region of Northeast China. Catena 2025, 249, 108623. [Google Scholar] [CrossRef]
  332. Liu, G.; Xia, J.; Zheng, K.; Cheng, J.; Wang, K.; Liu, Z.; Wei, Y.; Xie, D. Measurement and evaluation method of farmland microtopography feature information based on 3D LiDAR and inertial measurement unit. Soil Tillage Res. 2024, 236, 105921. [Google Scholar] [CrossRef]
  333. Ambaru, B.; Manvitha, R.; Madas, R. Synergistic integration of remote sensing and soil metagenomics data: Advancing precision agriculture through interdisciplinary approaches. Front. Sustain. Food Syst. 2024, 8, 1499973. [Google Scholar] [CrossRef]
  334. Wrbka, T.; Erb, K.H.; Schulz, N.B.; Peterseil, J.; Hahn, C.; Haberl, H. Linking pattern and process in cultural landscapes. An empirical study based on spatially explicit indicators. Land Use Policy 2004, 21, 289–306. [Google Scholar] [CrossRef]
  335. Sertel, E.; Topaloğlu, R.H.; Şallı, B.; Algan, I.Y.; Aksu, G.A. Comparison of landscape metrics for three different level land cover/land use maps. ISPRS Int. J. Geo-Inf. 2018, 7, 408. [Google Scholar] [CrossRef]
  336. Cai, Z.; Hu, Q.; Zhang, X.; Yang, J.; Wei, H.; He, Z.; Song, Q.; Wang, C.; Yin, G.; Xu, B. An Adaptive Image Segmentation Method with Automatic Selection of Optimal Scale for Extracting Cropland Parcels in Smallholder Farming Systems. Remote Sens. 2022, 14, 3067. [Google Scholar] [CrossRef]
  337. Wagner, M.P.; Oppelt, N. Extracting agricultural fields from remote sensing imagery using graph-based growing contours. Remote Sens. 2020, 12, 1205. [Google Scholar] [CrossRef]
  338. Watkins, B.; Van Niekerk, A. Automating field boundary delineation with multi-temporal Sentinel-2 imagery. Comput. Electron. Agric. 2019, 167, 105078. [Google Scholar] [CrossRef]
  339. Tieskens, K.F.; Schulp, C.J.E.; Levers, C.; Lieskovský, J.; Kuemmerle, T.; Plieninger, T.; Verburg, P.H. Characterizing European cultural landscapes: Accounting for structure, management intensity and value of agricultural and forest landscapes. Land Use Policy 2017, 62, 29–39. [Google Scholar] [CrossRef]
  340. Ye, S.; Ren, S.; Song, C.; Du, Z.; Wang, K.; Du, B.; Cheng, F.; Zhu, D. Spatial pattern of cultivated land fragmentation in mainland China: Characteristics, dominant factors, and countermeasures. Land Use Policy 2024, 139, 107070. [Google Scholar] [CrossRef]
  341. Erikstad, L.; Simensen, T.; Bakkestuen, V.; Halvorsen, R. Index Measuring Land Use Intensity—A Gradient-Based Approach. Geomatics 2023, 3, 188–204. [Google Scholar] [CrossRef]
  342. Meier, E.S.; Indermaur, A.; Ginzler, C.; Psomas, A. An effective way to map land-use intensity with a high spatial resolution based on habitat type and environmental data. Remote Sens. 2020, 12, 969. [Google Scholar] [CrossRef]
  343. Forkuor, G.; Conrad, C.; Thiel, M.; Zoungrana, B.J.B.; Tondoh, J.E. Multiscale remote sensing to map the spatial distribution and extent of cropland in the sudanian savanna of West Africa. Remote Sens. 2017, 9, 839. [Google Scholar] [CrossRef]
  344. Eckert, S.; Kiteme, B.; Njuguna, E.; Zaehringer, J.G. Agricultural expansion and intensification in the foothills of Mount Kenya: A landscape perspective. Remote Sens. 2017, 9, 784. [Google Scholar] [CrossRef]
  345. Knauer, K.; Gessner, U.; Fensholt, R.; Forkuor, G.; Kuenzer, C. Monitoring agricultural expansion in Burkina Faso over 14 years with 30 m resolution time series: The role of population growth and implications for the environment. Remote Sens. 2017, 9, 132. [Google Scholar] [CrossRef]
  346. Gilcher, M.; Udelhoven, T. Field geometry and the spatial and temporal generalization of crop classification algorithms—A randomized approach to compare pixel based and convolution based methods. Remote Sens. 2021, 13, 775. [Google Scholar] [CrossRef]
  347. Potapov, P.; Turubanova, S.; Hansen, M.C.; Tyukavina, A.; Zalles, V.; Khan, A.; Song, X.P.; Pickens, A.; Shen, Q.; Cortez, J. Global maps of cropland extent and change show accelerated cropland expansion in the twenty-first century. Nat. Food 2022, 3, 19–28. [Google Scholar] [CrossRef] [PubMed]
  348. Janus, J. Measuring land fragmentation considering the shape of transportation network: A method to increase the accuracy of modeling the spatial structure of agriculture with case study in Poland. Comput. Electron. Agric. 2018, 148, 259–271. [Google Scholar] [CrossRef]
  349. Wang, H.; Feng, C.; Li, X.; Yang, Y.; Zhang, Y.; Su, J.; Luo, D.; Wei, D.; He, Y. Plant Species Diversity Assessment in the Temperate Grassland Region of China Using UAV Hyperspectral Remote Sensing. Diversity 2024, 16, 775. [Google Scholar] [CrossRef]
  350. Xu, D.; Chen, H.; Ji, F.; Zhu, J.; Wang, Z.; Zhang, R.; Hou, M.; Huang, X.; Wang, D.; Lu, T.; et al. New insights on canopy heterogeneous analysis and light micro-climate simulation in Chinese solar greenhouse. Comput. Electron. Agric. 2025, 233, 110179. [Google Scholar] [CrossRef]
  351. Contreras, F.; Cayuela, M.L.; Sánchez-Monedero, M.Á.; Pérez-Cutillas, P. Multi-Source Remote Sensing for large-scale biomass estimation in mediterranean olive orchards using GEDI LiDAR and Machine Learning. EGUsphere 2025. [Google Scholar]
  352. Feng, H.; Fan, Y.; Yue, J.; Bian, M.; Liu, Y.; Chen, R.; Ma, Y.; Fan, J.; Yang, G.; Zhao, C. Estimation of potato above-ground biomass based on the VGC-AGB model and deep learning. Comput. Electron. Agric. 2025, 232, 110122. [Google Scholar] [CrossRef]
  353. Chen, M.; Tang, Y.; Zou, X.; Huang, Z.; Zhou, H.; Chen, S. 3D global mapping of large-scale unstructured orchard integrating eye-in-hand stereo vision and SLAM. Comput. Electron. Agric. 2021, 187, 106237. [Google Scholar] [CrossRef]
  354. Hadadi, M.; Saraeian, M.; Godbersen, J.; Jubery, T.; Li, Y.; Attigala, L.; Aditya Balu, S.S.; Schnable, P.S.; Krishnamurthy, A.; Ganapathysubramanian, B.; et al. Procedural Generation of 3D Maize Plant Architecture from LIDAR. Comput. Sci. 2025, 236, 110382. [Google Scholar] [CrossRef]
  355. Bao, Y.; Tang, L.; Srinivasan, S.; Schnable, P.S. ScienceDirect Field-based architectural traits characterisation of maize plant using time-of-flight 3D imaging. Biosyst. Eng. 2018, 178, 86–101. [Google Scholar] [CrossRef]
  356. Rocchini, D.; Dadalt, L.; Delucchi, L.; Neteler, M.; Palmer, M.W. Disentangling the role of remotely sensed spectral heterogeneity as a proxy for North American plant species richness. Community Ecol. 2014, 15, 37–43. [Google Scholar] [CrossRef]
  357. Yang, H.; Yang, Z.; Wu, Y.; Wang, C.; Wu, Y.; Zhang, P.; Wang, B. High-Resolution Remote Sensing Farmland Extraction Network Based on Dense-Feature Overlay Fusion and Information. IEEE Geosci. Remote Sens. Lett. 2025, 22, 1–5. [Google Scholar] [CrossRef]
  358. Pooya, M.R.; Hasankhani, A.; Fathololomi, S.; Karimi Firozjaei, M. A Spatial Multi-Criteria Decision-Making Approach to Evaluating Homogeneous Areas for Rainfed Wheat Yield Assessment. Water 2025, 17, 1045. [Google Scholar] [CrossRef]
  359. Fang, G.; Wang, C.; Dong, T.; Wang, Z.; Cai, C.; Chen, J.; Liu, M.; Zhang, H. A Landscape-Clustering Zoning Strategy to Map Multi-Crops in Fragmented Cropland Regions Using Sentinel-2 and Sentinel-1 Imagery with Feature Selection. Agriculture 2025, 15, 186. [Google Scholar] [CrossRef]
  360. Abdulraheem, M.I.; Zhang, W.; Li, S.; Moshayedi, A.J.; Farooque, A.A.; Hu, J. Advancement of Remote Sensing for Soil Measurements and Applications: A Comprehensive Review. Sustain. 2023, 15, 15444. [Google Scholar] [CrossRef]
  361. Yuzugullu, O.; Fajraoui, N.; Liebisch, F. Soil Texture and pH Mapping Using Remote Sensing and Support Sampling. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2024, 17, 12685–12705. [Google Scholar] [CrossRef]
  362. Gao, L.; Zhang, Y.; Zang, D.; Yang, Q.; Liu, H.; Luo, C. Using Satellites to Monitor Soil Texture in Typical Black Soil Areas and Assess Its Impact on Crop Growth. Agriculture 2025, 15, 912. [Google Scholar] [CrossRef]
  363. Liu, F.; Zhang, G.L.; Song, X.; Li, D.; Zhao, Y.; Yang, J.; Wu, H.; Yang, F. High-resolution and three-dimensional mapping of soil texture of China. Geoderma 2020, 361, 114061. [Google Scholar] [CrossRef]
  364. Bahrami, A.; Danesh, M.; Bahrami, M. Studying sand component of soil texture using the spectroscopic method. Infrared Phys. Technol. 2022, 122, 104056. [Google Scholar] [CrossRef]
  365. Swetha, R.K.; Bende, P.; Singh, K.; Gorthi, S.; Biswas, A.; Li, B.; Weindorf, D.C.; Chakraborty, S. Predicting soil texture from smartphone-captured digital images and an application. Geoderma 2020, 376, 114562. [Google Scholar] [CrossRef]
  366. Taghizadeh-Mehrjardi, R.; Emadi, M.; Cherati, A.; Heung, B.; Mosavi, A.; Scholten, T. Bio-inspired hybridization of artificial neural networks: An application for mapping the spatial distribution of soil texture fractions. Remote Sens. 2021, 13, 1025. [Google Scholar] [CrossRef]
  367. Swain, S.R.; Chakraborty, P.; Panigrahi, N.; Vasava, H.B.; Reddy, N.N.; Roy, S.; Majeed, I.; Das, B.S. Estimation of soil texture using Sentinel-2 multispectral imaging data: An ensemble modeling approach. Soil Tillage Res. 2021, 213, 105134. [Google Scholar] [CrossRef]
  368. Shen, Q.; Shang, K.; Xiao, C.; Tang, H.; Wu, T.; Wang, C. A novel hyperspectral remote sensing estimation model for surface soil texture using AHSI/ZY1-02D satellite image. Int. J. Appl. Earth Obs. Geoinf. 2025, 138, 104453. [Google Scholar] [CrossRef]
  369. Tong, X.; Brandt, M.; Hiernaux, P.; Herrmann, S.M.; Tian, F.; Prishchepov, A.V.; Fensholt, R. Revisiting the coupling between NDVI trends and cropland changes in the Sahel drylands: A case study in western Niger. Remote Sens. Environ. 2017, 191, 286–296. [Google Scholar] [CrossRef]
  370. Tao, J.; Jiang, Q.; Zhang, X.; Huang, J.; Wang, Y.; Wu, W. From frequency to intensity—A new index for annual large-scale cropping intensity mapping. Comput. Electron. Agric. 2023, 215, 108428. [Google Scholar] [CrossRef]
  371. Estel, S.; Kuemmerle, T.; Levers, C.; Baumann, M.; Hostert, P. Mapping cropland-use intensity across Europe using MODIS NDVI time series. Environ. Res. Lett. 2016, 11, 024015. [Google Scholar] [CrossRef]
  372. Defourny, P.; Bontemps, S.; Bellemans, N.; Cara, C.; Dedieu, G.; Guzzonato, E.; Hagolle, O.; Inglada, J.; Nicola, L.; Rabaute, T.; et al. Near real-time agriculture monitoring at national scale at parcel resolution: Performance assessment of the Sen2-Agri automated system in various cropping systems around the world. Remote Sens. Environ. 2019, 221, 551–568. [Google Scholar] [CrossRef]
  373. Xu, F.; Yao, X.; Zhang, K.; Yang, H.; Feng, Q.; Li, Y.; Yan, S.; Gao, B.; Li, S.; Yang, J.; et al. Deep learning in cropland field identification: A review. Comput. Electron. Agric. 2024, 222, 109042. [Google Scholar] [CrossRef]
  374. Bouguettaya, A.; Zarzour, H.; Kechida, A.; Taberkit, A.M. Deep learning techniques to classify agricultural crops through UAV imagery: A review. Neural Comput. Appl. 2022, 34, 9511–9536. [Google Scholar] [CrossRef] [PubMed]
  375. Choukri, M.; Laamrani, A.; Chehbouni, A. Use of Optical and Radar Imagery for Crop Type Classification in Africa: A Review. Sensors 2024, 24, 3618. [Google Scholar] [CrossRef]
  376. Tan, Y.; Gu, J.; Lu, L.; Zhang, L.; Huang, J.; Pan, L.; Lv, Y.; Wang, Y.; Chen, Y. Hyperspectral Band Selection for Crop Identification and Mapping of Agriculture. Remote Sens. 2025, 17, 663. [Google Scholar] [CrossRef]
  377. Bourriz, M.; Hajji, H.; Laamrani, A.; Elbouanani, N.; Abdelali, H.A.; Bourzeix, F.; El-Battay, A.; Amazirh, A.; Chehbouni, A. Integration of Hyperspectral Imaging and AI Techniques for Crop Type Mapping: Present Status, Trends, and Challenges. Remote Sens. 2025, 17, 1574. [Google Scholar] [CrossRef]
  378. Muthoka, J.M.; Rowhani, P.; Salakpi, E.E.; Balzter, H.; Antonarakis, A.S. Classification of grassland community types and palatable pastures in semi-arid savannah grasslands of Kenya using multispectral Sentinel-2 imagery. Front. Sustain. Food Syst. 2025, 9, 1543491. [Google Scholar] [CrossRef]
  379. Li, Y.; Liu, T.; Wang, Y.; Duan, L.; Li, M.; Zhang, J.; Zhang, G. A more effective approach for species-level classifications using multi-source remote sensing data: Validation and application to an arid and semi-arid grassland. Ecol. Indic. 2024, 160, 111853. [Google Scholar] [CrossRef]
  380. Zhang, M.; Yu, W.; Chen, A.; Xu, C.; Guo, J.; Xing, X.; Yang, D.; Wang, Z.; Yang, X. Two-tier classification framework for mapping grassland types using multisource earth observation data. GIScience Remote Sens. 2024, 61, 2385170. [Google Scholar] [CrossRef]
  381. Liu, C.; Zhang, Q.; Tao, S.; Qi, J.; Ding, M.; Guan, Q.; Wu, B.; Zhang, M.; Nabil, M.; Tian, F.; et al. A new framework to map fine resolution cropping intensity across the globe: Algorithm, validation, and implication. Remote Sens. Environ. 2020, 251, 112095. [Google Scholar] [CrossRef]
  382. Krishna, G.; Biradar, C.M. Spatiotemporal mapping of rice fallows and soil moisture dynamics for sustainable intensification using time series geospatial big data. J. Appl. Remote Sens. 2025, 19, 014515. [Google Scholar] [CrossRef]
  383. Tao, J.; Wu, W.; Liu, W. Spatial-temporal dynamics of cropping frequency in hubei province over 2001–2015. Sensors 2017, 17, 2622. [Google Scholar] [CrossRef] [PubMed]
  384. Xiang, M.; Yu, Q.; Wu, W. From multiple cropping index to multiple cropping frequency: Observing cropland use intensity at a finer scale. Ecol. Indic. 2019, 101, 892–903. [Google Scholar] [CrossRef]
  385. Arjasakusuma, S.; Kusuma, S.S.; Mahendra, W.K.; Astriviany, N. Mapping paddy field extent and temporal pattern variation in a complex terrain area using sentinel 1-time series data: Case study of magelang district, indonesia. Int. J. Geoinform. 2021, 17, 79–88. [Google Scholar] [CrossRef]
  386. Pazhanivelan, S.; Kumaraperumal, R.; Vishnu Priya, M.; Rengabashyam, K.; Shankar, K.; Nivas Raj, M.; Yadav, M.K. Multi-Temporal Analysis of Cropping Patterns and Intensity Using Optical and SAR Satellite Data for Sustaining Agricultural Production in Tamil Nadu, India. Sustainability 2025, 17, 1613. [Google Scholar] [CrossRef]
  387. Yu, Q.; Xiang, M.; Sun, Z.; Wu, W. The complexity of measuring cropland use intensity: An empirical study. Agric. Syst. 2021, 192, 103180. [Google Scholar] [CrossRef]
  388. Conrad, C.; Schönbrodt-Stitt, S.; Löw, F.; Sorokin, D.; Paeth, H. Cropping intensity in the Aral Sea Basin and its dependency from the runoffformation 2000–2012. Remote Sens. 2016, 8, 630. [Google Scholar] [CrossRef]
  389. Rufin, P.; Levers, C.; Baumann, M.; Jägermeyr, J.; Krueger, T.; Kuemmerle, T.; Hostert, P. Global-scale patterns and determinants of cropping frequency in irrigation dam command areas. Glob. Environ. Change 2018, 50, 110–122. [Google Scholar] [CrossRef]
  390. Barbieri, P.; Pellerin, S.; Nesme, T. Comparing crop rotations between organic and conventional farming. Sci. Rep. 2017, 7, 13761. [Google Scholar] [CrossRef] [PubMed]
  391. Zheng, B.; Campbell, J.B.; de Beurs, K.M. Remote sensing of crop residue cover using multi-temporal Landsat imagery. Remote Sens. Environ. 2012, 117, 177–183. [Google Scholar] [CrossRef]
  392. Zhang, W.; Li, W.; Wang, C.; Yu, Q.; Tang, H.; Wu, W. A novel index for mapping crop residue covered cropland using remote sensing data. Comput. Electron. Agric. 2025, 231, 109995. [Google Scholar] [CrossRef]
  393. Du, J.; Jacinthe, P.-A.; Song, K.; Zhang, L.; Zhao, B.; Liu, H.; Wang, Y.; Zhang, W.; Zheng, Z.; Yu, W.; et al. Maize crop residue cover mapping using Sentinel-2 MSI data and random forest algorithms. Int. Soil Water Conserv. Res. 2025, 13, 189–202. [Google Scholar] [CrossRef]
  394. Yao, Y.; Ren, H.; Liu, Y. Remote sensing estimation of winter wheat residue cover with dry and wet soil background. Agric. Water Manag. 2025, 307, 109227. [Google Scholar] [CrossRef]
  395. Williams, F.; Gelder, B.; Presley, D.A.; Pape, B.; Einck, A. Estimation of Crop Residue Cover Utilizing Multiple Ground Truth Survey Techniques and Multi-Satellite Regression Models. Remote Sens. 2024, 16, 4185. [Google Scholar] [CrossRef]
  396. Kavoosi, Z.; Raoufat, M.H.; Dehghani, M.; Abdolabbas, J.; Kazemeini, S.A.; Nazemossadat, M.J. Feasibility of satellite and drone images for monitoring soil residue cover. J. Saudi Soc. Agric. Sci. 2020, 19, 56–64. [Google Scholar] [CrossRef]
  397. Singh, G.; Kant, Y.; Dadhwal, V.K. Remote sensing of crop residue burning in Punjab (India): A study on burned area estimation using multi-sensor approach. Geocarto Int. 2009, 24, 273–292. [Google Scholar] [CrossRef]
  398. Sharma, A.; Kumar Singh, P. Applicability of UAVs in detecting and monitoring burning residue of paddy crops with IoT Integration: A step towards greener environment. Comput. Ind. Eng. 2023, 184, 109524. [Google Scholar] [CrossRef]
  399. Walker, K. Overcoming Common Pitfalls to Improve the Accuracy of Crop Residue Burning Measurement Based on Remote Sensing Data. Remote Sens. 2024, 16, 342. [Google Scholar] [CrossRef]
  400. Xie, Z.; Zhao, Y.; Jiang, R.; Zhang, M.; Hammer, G.; Chapman, S.; Brider, J.; Potgieter, A.B. Seasonal dynamics of fallow and cropping lands in the broadacre cropping region of Australia. Remote Sens. Environ. 2024, 305, 114070. [Google Scholar] [CrossRef]
  401. Oliphant, A.J.; Thenkabail, P.S.; Teluguntla, P.G.; Aneece, I.P.; Foley, D.J.; McCormick, R.L. Automated Cropland Fallow Algorithm (ACFA) for the Northern Great Plains of USA. Int. J. Digit. Earth 2024, 17, 2337221. [Google Scholar] [CrossRef]
  402. Tong, X.; Brandt, M.; Hiernaux, P.; Herrmann, S.; Rasmussen, L.V.; Rasmussen, K.; Tian, F.; Tagesson, T.; Zhang, W.; Fensholt, R. The forgotten land use class: Mapping of fallow fields across the Sahel using Sentinel-2. Remote Sens. Environ. 2020, 239, 111598. [Google Scholar] [CrossRef]
  403. Denis, A.; Desclee, B.; Migdall, S.; Hansen, H.; Bach, H.; Ott, P.; Kouadio, A.L.; Tychon, B. Multispectral remote sensing as a tool to support organic crop certification: Assessment of the discrimination level between organic and conventional maize. Remote Sens. 2021, 13, 117. [Google Scholar] [CrossRef]
  404. Atanasova, D.; Bozhanova, V.; Biserkov, V.; Maneva, V. Distinguishing areas of organic, biodynamic and conventional farming by means of multispectral images. A pilot study. Biotechnol. Biotechnol. Equip. 2021, 35, 977–993. [Google Scholar] [CrossRef]
  405. Denis, A.; Tychon, B. Remote sensing enables high discrimination between organic and non-organic cotton for organic cotton certification in West Africa. Agron. Sustain. Dev. 2015, 35, 1499–1510. [Google Scholar] [CrossRef]
  406. Serrano-Grijalva, L.; Ochoa-Hueso, R.; Veen, G.F.; Repeto-Deudero, I.; Van Rijssel, S.Q.; Koorneef, G.J.; Van der Putten, W.H. Normalized difference vegetation index analysis reveals increase of biomass production and stability during the conversion from conventional to organic farming. Glob. Change Biol. 2024, 30, e17461. [Google Scholar] [CrossRef] [PubMed]
  407. Cheng, T.; Zhang, D.; Zhang, G.; Wang, T.; Ren, W.; Yuan, F.; Liu, Y.; Wang, Z.; Zhao, C. High-throughput phenotyping techniques for forage: Status, bottleneck, and challenges. Artif. Intell. Agric. 2025, 15, 98–115. [Google Scholar] [CrossRef]
  408. Lausch, A.; Salbach, C.; Schmidt, A.; Doktor, D.; Merbach, I.; Pause, M. Deriving phenology of barley with imaging hyperspectral remote sensing. Ecol. Modell. 2015, 295, 123–135. [Google Scholar] [CrossRef]
  409. Ispizua Yamati, F.R.; Bömer, J.; Noack, N.; Linkugel, T.; Paulus, S.; Mahlein, A.K. Configuration of a multisensor platform for advanced plant phenotyping and disease detection: Case study on Cercospora leaf spot in sugar beet. Smart Agric. Technol. 2025, 10, 100740. [Google Scholar] [CrossRef]
  410. Grubinger, S.; Coops, N.C.; O’Neill, G.A.; Degner, J.C.; Wang, T.; Waite, O.J.M.; Riofrío, J.; Koch, T.L. Seasonal vegetation dynamics for phenotyping using multispectral drone imagery: Genetic differentiation, climate adaptation, and hybridization in a common-garden trial of interior spruce (Picea engelmannii × glauca). Remote Sens. Environ. 2025, 317, 114512. [Google Scholar] [CrossRef]
  411. Peng, Y.; Solovchenko, A.; Zhang, C.; Shurygin, B.; Liu, X.; Wu, X.; Gong, Y.; Fang, S.; Gitelson, A. Remote sensing of rice phenology and physiology via absorption coefficient derived from unmanned aerial vehicle imaging. Precis. Agric. 2024, 25, 285–302. [Google Scholar] [CrossRef]
  412. Zheng, C.; Abd-elrahman, A.; Whitaker, V. Remote sensing and machine learning in crop phenotyping and management, with an emphasis on applications in strawberry farming. Remote Sens. 2021, 13, 531. [Google Scholar] [CrossRef]
  413. French, A.; Sanchez, C.A.; Wirth, T.; Scott, A.T.; Shields, J.; Bautista, E.; Saber, M.N.; Wisniewski, E. Estimating Fao-56 Crop Growth Stage Lengths with Sentinel 2. SSRN Electron. J. 2022, 4068442. [Google Scholar] [CrossRef]
  414. Xiang, M.; Yu, Q.; Li, Y.; Shi, Z.; Wu, W. Increasing multiple cropping for land use intensification: The role of crop choice. Land Use Policy 2022, 112, 105846. [Google Scholar] [CrossRef]
  415. Bahrami, H.; Homayouni, S.; Safari, A.; Mirzaei, S.; Mahdianpari, M.; Reisi-Gahrouei, O. Deep learning-based estimation of crop biophysical parameters using multi-source and multi-temporal remote sensing observations. Agronomy 2021, 11, 1363. [Google Scholar] [CrossRef]
  416. Mokhtari, A.; Noory, H.; Vazifedoust, M.; Palouj, M.; Bakhtiari, A.; Barikani, E.; Zabihi Afrooz, R.A.; Fereydooni, F.; Sadeghi Naeni, A.; Pourshakouri, F.; et al. Evaluation of single crop coefficient curves derived from Landsat satellite images for major crops in Iran. Agric. Water Manag. 2019, 218, 234–249. [Google Scholar] [CrossRef]
  417. Pirbasti, M.A.; Mcardle, G.; Akbari, V. Hedgerows Monitoring in Remote Sensing: A Comprehensive Review. IEEE Access 2024, 12, 156184–156207. [Google Scholar] [CrossRef]
  418. Ahlswede, S.; Asam, S.; Röder, A. Hedgerow object detection in very high-resolution satellite images using convolutional neural networks. J. Appl. Remote Sens. 2021, 15, 018501. [Google Scholar] [CrossRef]
  419. Wolstenholme, J.M.; Cooper, F.; Thomas, R.E.; Ahmed, J.; Parsons, K.J.; Parsons, D.R. Automated identification of hedgerows and hedgerow gaps using deep learning. Remote Sens. Ecol. Conserv. 2025, 11, 411–424. [Google Scholar] [CrossRef]
  420. Smigaj, M.; Gaulton, R. Capturing hedgerow structure and flowering abundance with UAV remote sensing. Remote Sens. Ecol. Conserv. 2021, 7, 521–533. [Google Scholar] [CrossRef]
  421. Barnsley, S.L.; Lovett, A.A.; Dicks, L.V. Mapping nectar-rich pollinator floral resources using airborne multispectral imagery. J. Environ. Manag. 2022, 313, 114942. [Google Scholar] [CrossRef] [PubMed]
  422. Prokoph, S.; Cheema, J.; Kirmer, A.; Lausch, A.; Bannehr, L. Monitoring von blütenreichen Flächen mittels Fernerkundung. DGPF 2022, 30, 220–236. [Google Scholar]
  423. Abdel-Rahman, E.M.; Makori, D.M.; Landmann, T.; Piiroinen, R.; Gasim, S.; Pellikka, P.; Raina, S.K. The utility of AISA eagle hyperspectral data and random forest classifier for flower mapping. Remote Sens. 2015, 7, 13298–13318. [Google Scholar] [CrossRef]
  424. Gallmann, J.; Schüpbach, B.; Jacot, K.; Albrecht, M.; Winizki, J.; Kirchgessner, N.; Aasen, H. Flower Mapping in Grasslands With Drones and Deep Learning. Front. Plant Sci. 2022, 12, 774965. [Google Scholar] [CrossRef]
  425. Gonzales, D.; Hempel de Ibarra, N.; Anderson, K. Remote Sensing of Floral Resources for Pollinators—New Horizons From Satellites to Drones. Front. Ecol. Evol. 2022, 10, 869751. [Google Scholar] [CrossRef]
  426. Schnalke, M.; Funk, J.; Wagner, A. Bridging technology and ecology: Enhancing applicability of deep learning and UAV-based flower recognition. Front. Plant Sci. 2025, 16, 1498913. [Google Scholar] [CrossRef]
  427. El Afandi, G.; Ismael, H.; Fall, S.; Ankumah, R. Effectiveness of Utilizing Remote Sensing and GIS Techniques to Estimate the Exposure to Organophosphate Pesticides Drift over Macon, Alabama. Agronomy 2023, 13, 1759. [Google Scholar] [CrossRef]
  428. Bolívar-Santamaría, S.; Reu, B. Detection and characterization of agroforestry systems in the Colombian Andes using sentinel-2 imagery. Agrofor. Syst. 2021, 95, 499–514. [Google Scholar] [CrossRef]
  429. Koralewicz, A.; Vlcek, J.; Menor, I.O.; Hirons, M.; Akinyugha, A.; Olowoyo, O.S.; Ajayi-Ebenezer, M.; Owen, O. Mapping the extent and exploring the drivers of cocoa agroforestry in Nigeria, insights into trends for climate change adaptation. Agrofor. Syst. 2025, 99, 38. [Google Scholar] [CrossRef]
  430. Yang, W.; Ortiz-Gonzalo, D.; Tong, X.; Gominski, D.; Fensholt, R. Mapping forest-agroforest frontiers in the Peruvian Amazon with deep learning and PlanetScope satellite data. Ecol. Inform. 2025, 86, 103034. [Google Scholar] [CrossRef]
  431. Ahmad, T.; Sahoo, P.M.; Jally, S.K. Estimation of area under agroforestry using high resolution satellite data. Agrofor. Syst. 2016, 90, 289–303. [Google Scholar] [CrossRef]
  432. Brandt, M.; Gominski, D.; Reiner, F.; Kariryaa, A.; Guthula, V.B.; Ciais, P.; Tong, X.; Zhang, W.; Govindarajulu, D.; Ortiz-Gonzalo, D.; et al. Severe decline in large farmland trees in India over the past decade. Nat. Sustain. 2024, 7, 860–868. [Google Scholar] [CrossRef]
  433. Veettil, B.K.; Van, D.D.; Quang, N.X.; Hoai, P.N. Remote sensing of plastic-covered greenhouses and plastic-mulched farmlands: Current trends and future perspectives. Land Degrad. Dev. 2023, 34, 591–609. [Google Scholar] [CrossRef]
  434. Niu, B.; Feng, Q.; Qiu, B.; Su, S.; Zhang, X.; Cui, R.; Zhang, X.; Sun, F.; Yan, W.; Zhao, S.; et al. Global-PCG-10: A 10-m global map of plastic-covered greenhouses derived from Sentinel-2 in 2020. Earth Syst. Sci. Data 2025, 17, 5065–5088. [Google Scholar] [CrossRef]
  435. Zhang, M.; Dong, J.; Ge, Q.; Hasituya; Hao, P. A Review of Agricultural Film Mapping: Current Status, Challenges, and Future Directions. J. Remote Sens. 2025, 5, 0395. [Google Scholar] [CrossRef]
  436. Wu, Y.; Dai, J.; Zhu, Y.; Zhang, T. Intelligent Agricultural Greenhouse Extraction Method Based on Multi-Feature Modeling: Fusion of Geometric, Spatial, and Spectral Characteristics. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2025, 18, 5568–5581. [Google Scholar] [CrossRef]
  437. Zhang, X.; Li, J.; Li, H.; Guo, Z.; Chang, C.; Xu, X.; Zhen, T.; Yu, K.; Li, P. Identification Of Plastic Film Mulched Farmland in the Core Area of the Beijing-Tianjin Sand Source Region Using Multi-Temporal Remote Sensing Features. Remote Sens. Appl. Soc. Environ. 2025, 38, 101600. [Google Scholar] [CrossRef]
  438. Shafi, U.; Mumtaz, R.; Anwar, Z.; Ajmal, M.M.; Khan, M.A.; Mahmood, Z.; Qamar, M.; Jhanzab, H.M. Tackling Food Insecurity Using Remote Sensing and Machine Learning-Based Crop Yield Prediction. IEEE Access 2023, 11, 108640–108657. [Google Scholar] [CrossRef]
  439. Haseeb, M.; Tahir, Z.; Mahmood, S.A.; Tariq, A. Winter wheat yield prediction using linear and nonlinear machine learning algorithms based on climatological and remote sensing data. Inf. Process. Agric. 2025, in press. [Google Scholar] [CrossRef]
  440. Fita, D.; Rubio, C.; Franch, B.; Castiñeira-Ibáñez, S.; Tarrazó-Serrano, D.; San Bautista, A. Improving harvester yield maps postprocessing leveraging remote sensing data in rice crop. Precis. Agric. 2025, 26, 33. [Google Scholar] [CrossRef]
  441. Hou, X.; Zhang, J.; Luo, X.; Zeng, S.; Lu, Y.; Wei, Q.; Liu, J.; Feng, W.; Li, Q. Peanut yield prediction using remote sensing and machine learning approaches based on phenological characteristics. Comput. Electron. Agric. 2025, 232, 110084. [Google Scholar] [CrossRef]
  442. Liu, Y.; Feng, H.; Fan, Y.; Yue, J.; Yang, F.; Fan, J.; Ma, Y.; Chen, R.; Bian, M.; Yang, G. Utilizing UAV-based hyperspectral remote sensing combined with various agronomic traits to monitor potato growth and estimate yield. Comput. Electron. Agric. 2025, 231, 109984. [Google Scholar] [CrossRef]
  443. Abdul-Jabbar, T.S.; Ziboon, A.T.; Albayati, M.M. Crop yield estimation using different remote sensing data: Literature review. IOP Conf. Ser. Earth Environ. Sci. 2023, 1129, 012004. [Google Scholar] [CrossRef]
  444. Badagliacca, G.; Messina, G.; Presti, E.L.; Preiti, G.; Di Fazio, S.; Monti, M.; Modica, G.; Praticò, S. Durum Wheat (Triticum durum Desf.) Grain Yield and Protein Estimation by Multispectral UAV Monitoring and Machine Learning Under Mediterranean Conditions. AgriEngineering 2025, 7, 99. [Google Scholar] [CrossRef]
  445. Raza, A.; Shahid, M.A.; Zaman, M.; Miao, Y.; Huang, Y.; Safdar, M.; Maqbool, S.; Muhammad, N.E. Improving Wheat Yield Prediction with Multi-Source Remote Sensing Data and Machine Learning in Arid Regions. Remote Sens. 2025, 17, 774. [Google Scholar] [CrossRef]
  446. Mena, F.; Pathak, D.; Najjar, H.; Sanchez, C.; Helber, P.; Bischke, B.; Habelitz, P.; Miranda, M.; Siddamsetty, J.; Nuske, M.; et al. Adaptive fusion of multi-modal remote sensing data for optimal sub-field crop yield prediction. Remote Sens. Environ. 2025, 318, 114547. [Google Scholar] [CrossRef]
  447. Dhillon, M.S.; Kübert-Flock, C.; Dahms, T.; Rummler, T.; Arnault, J.; Steffan-Dewenter, I.; Ullmann, T. Evaluation of MODIS, Landsat 8 and Sentinel-2 Data for Accurate Crop Yield Predictions: A Case Study Using STARFM NDVI in Bavaria, Germany. Remote Sens. 2023, 15, 1830. [Google Scholar] [CrossRef]
  448. Řeřicha, J.; Kohútek, M.; Vandírková, V.; Krofta, K.; Kumhála, F.; Kumhálová, J. Assessment of UAV Imageries for Estimating Growth Vitality, Yield and Quality of Hop (Humulus lupulus L.) Crops. Remote Sens. 2025, 17, 970. [Google Scholar] [CrossRef]
  449. Brito, L.G.d.; Jorge, R.C.; Oliveira, V.C.d.; Cassemiro, P.F.; Dal Pai, A.; Sarnighausen, V.C.R.; Rodrigues, S.A. Classification Models for Nitrogen Concentration in Hop Leaves Using Digital Image Processing. Appl. Sci. 2025, 15, 4799. [Google Scholar] [CrossRef]
  450. Dietrich, P.; Elias, M.; Dietrich, P.; Harpole, S.; Roscher, C.; Bumberger, J. Advancing plant biomass measurements: Integrating smartphone-based 3D scanning techniques for enhanced ecosystem monitoring. Methods Ecol. Evol. 2025, 2025, 1723–1732. [Google Scholar] [CrossRef]
  451. Zhang, H.; Zhang, L.; Wu, H.; Wang, D.; Ma, X.; Shao, Y.; Jiang, M.; Chen, X. Unmanned-Aerial-Vehicle-Based Multispectral Monitoring of Nitrogen Content in Canopy Leaves of Processed Tomatoes. Agriculture 2025, 15, 309. [Google Scholar] [CrossRef]
  452. Acharya, B.; Dodla, S.; Tubana, B.; Gentimis, T.; Rontani, F.; Adhikari, R.; Duron, D.; Bortolon, G.; Setiyono, T. Characterizing Optimum N Rate in Waterlogged Maize (Zea mays L.) with Unmanned Aerial Vehicle (UAV) Remote Sensing. Agronomy 2025, 15, 434. [Google Scholar] [CrossRef]
  453. Segarra, J.; Buchaillot, M.L.; Stefani, U.; Araus, J.L.; Kefauver, S.C. Sentinel-2 Responsiveness to Fertilization Gradients in Wheat at Field Level in Córdoba Province, Argentina. In Proceedings of the 2020 Mediterranean and Middle-East Geoscience and Remote Sensing Symposium (M2GARSS), Tunis, Tunisia, 9–11 March 2020; pp. 322–325. [Google Scholar]
  454. Zhou, T.; Geng, Y.; Lv, W.; Xiao, S.; Zhang, P.; Xu, X.; Chen, J.; Wu, Z.; Pan, J.; Si, B.; et al. Effects of optical and radar satellite observations within Google Earth Engine on soil organic carbon prediction models in Spain. J. Environ. Manag. 2023, 338, 117810. [Google Scholar] [CrossRef]
  455. Zhou, T.; Geng, Y.; Chen, J.; Liu, M.; Haase, D.; Lausch, A. Mapping soil organic carbon content using multi-source remote sensing variables in the Heihe River Basin in China. Ecol. Indic. 2020, 114, 106288. [Google Scholar] [CrossRef]
  456. Suleymanov, A.; Gabbasova, I.; Suleymanov, R.; Abakumov, E.; Polyakov, V.; Liebelt, P. Mapping soil organic carbon under erosion processes using remote sensing. Hung. Geogr. Bull. 2021, 70, 49–64. [Google Scholar] [CrossRef]
  457. Yuzugullu, O.; Fajraoui, N.; Don, A.; Liebisch, F. Satellite-based soil organic carbon mapping on European soils using available datasets and support sampling. Sci. Remote Sens. 2024, 9, 100118. [Google Scholar] [CrossRef]
  458. Radočaj, D.; Gašparović, M.; Jurišić, M. Open Remote Sensing Data in Digital Soil Organic Carbon Mapping: A Review. Agric. 2024, 14, 1005. [Google Scholar] [CrossRef]
  459. Qian, J.; Yang, J.; Sun, W.; Zhao, L.; Shi, L.; Shi, H.; Liao, L.; Dang, C. Soil Organic Carbon Estimation and Transfer Framework in Agricultural Areas Based on Spatiotemporal Constraint Strategy Combined with Active and Passive Remote Sensing. Remote. Sens. 2025, 17, 333. [Google Scholar] [CrossRef]
  460. Geng, J.; Tan, Q.; Lv, J.; Fang, H. Assessing spatial variations in soil organic carbon and C:N ratio in Northeast China’s black soil region: Insights from Landsat-9 satellite and crop growth information. Soil Tillage Res. 2024, 235, 105897. [Google Scholar] [CrossRef]
  461. Xu, D.; Chen, S.; Zhou, Y.; Ji, W.; Shi, Z. Spatial Estimation of Soil Organic Matter and Total Nitrogen by Fusing Field Vis–NIR Spectroscopy and Multispectral Remote Sensing Data. Remote Sens. 2025, 17, 729. [Google Scholar] [CrossRef]
  462. Zhang, Y.; Luo, C.; Zhang, W.; Wu, Z.; Zang, D. Mapping Soil Organic Matter in Black Soil Cropland Areas Using Remote Sensing and Environmental Covariates. Agriculture 2025, 15, 339. [Google Scholar] [CrossRef]
  463. Zhang, J.; Gan, S.; Yang, P.; Zhou, J.; Huang, X.; Chen, H.; He, H.; Saintilan, N.; Sanders, C.J.; Wang, F. A global assessment of mangrove soil organic carbon sources and implications for blue carbon credit. Nat. Commun. 2024, 15, 8994. [Google Scholar] [CrossRef]
  464. Chen, X.; Yuan, F.; Ata-Ul-Karim, S.T.; Liu, X.; Tian, Y.; Zhu, Y.; Cao, W.; Cao, Q. A bibliometric analysis of research on remote sensing-based monitoring of soil organic matter conducted between 2003 and 2023. Artif. Intell. Agric. 2025, 15, 26–38. [Google Scholar] [CrossRef]
  465. Van Wasmael, B. A European soil organic carbon monitoring system leveraging Sentinel 2 imagery and the LUCAS soil data base. Geoderma 2024, 452, 117113. [Google Scholar] [CrossRef]
  466. Petropoulos, T.; Benos, L.; Busato, P.; Kyriakarakos, G.; Kateris, D.; Aidonis, D.; Bochtis, D. Soil Organic Carbon Assessment for Carbon Farming: A Review. Agriculture 2025, 15, 567. [Google Scholar] [CrossRef]
  467. Gomez, C.; Oltra-carrió, R.; Bacha, S.; Lagacherie, P.; Briottet, X. Evaluating the sensitivity of clay content prediction to atmospheric effects and degradation of image spatial resolution using Hyperspectral VNIR/SWIR imagery. Remote Sens. Environ. 2015, 164, 1–15. [Google Scholar] [CrossRef]
  468. Shabou, M.; Mougenot, B.; Chabaane, Z.L.; Walter, C.; Boulet, G.; Aissa, N.B.; Zribi, M. Soil clay content mapping using a time series of Landsat TM data in semi-arid lands. Remote Sens. 2015, 7, 6059–6078. [Google Scholar] [CrossRef]
  469. Gomez, C.; Adeline, K.; Bacha, S.; Driessen, B.; Gorretta, N.; Lagacherie, P.; Roger, J.M.; Briottet, X. Sensitivity of clay content prediction to spectral configuration of VNIR/SWIR imaging data, from multispectral to hyperspectral scenarios. Remote Sens. Environ. 2018, 204, 18–30. [Google Scholar] [CrossRef]
  470. Paul, S.S.; Coops, N.C.; Johnson, M.S.; Krzic, M.; Chandna, A.; Smukler, S.M. Mapping soil organic carbon and clay using remote sensing to predict soil workability for enhanced climate change adaptation. Geoderma 2020, 363, 114177. [Google Scholar] [CrossRef]
  471. Gasmi, A.; Gomez, C.; Lagacherie, P.; Zouari, H. Surface soil clay content mapping at large scales using multispectral (VNIR–SWIR) ASTER data. Int. J. Remote Sens. 2019, 40, 1506–1533. [Google Scholar] [CrossRef]
  472. Franz, A.; Sowiński, J.; Głogowski, A.; Fiałkiewicz, W. A Preliminary Study on the Use of Remote Sensing Techniques to Determine the Nutritional Status and Productivity of Oats on Spatially Variable Sandy Soils. Agronomy 2025, 15, 616. [Google Scholar] [CrossRef]
  473. Fongaro, C.T.; Demattê, J.A.M.; Rizzo, R.; Safanelli, J.L.; Mendes, W.d.S.; Dotto, A.C.; Vicente, L.E.; Franceschini, M.H.D.; Ustin, S.L. Improvement of clay and sand quantification based on a novel approach with a focus on multispectral satellite images. Remote Sens. 2018, 10, 1555. [Google Scholar] [CrossRef]
  474. Cerasola, V.A.; Orsini, F.; Pennisi, G.; Moretti, G.; Bona, S.; Mirone, F.; Verrelst, J.; Berger, K.; Gianquinto, G. Hyperspectral imaging for precision nitrogen management: A comparative exploration of two methodological approaches to estimate optimal nitrogen rate in processing tomato. Smart Agric. Technol. 2025, 10, 100802. [Google Scholar] [CrossRef]
  475. Wang, J.; Wang, W.; Liu, S.; Hui, X.; Zhang, H.; Yan, H.; Maes, W.H. UAV-Based Multiple Sensors for Enhanced Data Fusion and Nitrogen Monitoring in Winter Wheat Across Growth Seasons. Remote Sens. 2025, 17, 498. [Google Scholar] [CrossRef]
  476. Misbah, K.; Laamrani, A.; Khechba, K.; Dhiba, D.; Chehbouni, A. Multi-sensors remote sensing applications for assessing, monitoring, and mapping npk content in soil and crops in african agricultural land. Remote Sens. 2022, 14, 81. [Google Scholar] [CrossRef]
  477. Peng, X.; Chen, D.; Zhou, Z.; Zhang, Z.; Xu, C.; Zha, Q.; Wang, F.; Hu, X. Prediction of the Nitrogen, Phosphorus and Potassium Contents in Grape Leaves at Different Growth Stages Based on UAV Multispectral Remote Sensing. Remote Sens. 2022, 14, 2659. [Google Scholar] [CrossRef]
  478. Zhang, W.; Zhu, L.; Zhuang, Q.; Chen, D.; Sun, T. Mapping Cropland Soil Nutrients Contents Based on Multi-Spectral Remote Sensing and Machine Learning. Agriculture 2023, 13, 1592. [Google Scholar] [CrossRef]
  479. Lin, C.; Ma, R.; Zhu, Q.; Li, J. Using hyper-spectral indices to detect soil phosphorus concentration for various land use patterns. Environ. Monit. Assess. 2015, 187, 4130. [Google Scholar] [CrossRef]
  480. Guo, C.; Zhang, L.; Zhou, X.; Zhu, Y.; Cao, W.; Qiu, X.; Cheng, T.; Tian, Y. Integrating remote sensing information with crop model to monitor wheat growth and yield based on simulation zone partitioning. Precis. Agric. 2018, 19, 55–78. [Google Scholar] [CrossRef]
  481. Wang, S.; Adhikari, K.; Wang, Q.; Jin, X.; Li, H. Role of environmental variables in the spatial distribution of soil carbon (C), nitrogen (N), and C:N ratio from the northeastern coastal agroecosystems in China. Ecol. Indic. 2018, 84, 263–272. [Google Scholar] [CrossRef]
  482. Wang, X.; Geng, Y.; Zhou, T.; Zhao, Y.; Li, H.; Liu, Y.; Li, H.; Ren, R.; Zhang, Y.; Xu, X.; et al. Mapping the soil C:N ratio at the European scale by combining multi-year Sentinel radar and optical data via cloud computing. Soil Tillage Res. 2025, 245, 106311. [Google Scholar] [CrossRef]
  483. Rapiya, M.; Ramoelo, A.; Truter, W. Seasonal monitoring of forage C:N:ADF ratio in natural rangeland using remote sensing data. Environ. Monit. Assess. 2025, 197, 137. [Google Scholar] [CrossRef] [PubMed]
  484. Wang, Y.; Chen, S.; Hong, Y.; Hu, B.; Peng, J.; Shi, Z. A comparison of multiple deep learning methods for predicting soil organic carbon in Southern Xinjiang, China. Comput. Electron. Agric. 2023, 212, 108067. [Google Scholar] [CrossRef]
  485. Lei, S.; Zhou, P.; Lin, J.; Tan, Z.; Huang, J.; Yan, P.; Chen, H. Spatiotemporal Variation in Carbon and Water Use Efficiency and Their Influencing Variables Based on Remote Sensing Data in the Nanling Mountains Region. Remote Sens. 2025, 17, 648. [Google Scholar] [CrossRef]
  486. Yue, F.; Liu, D.; Xiong, L.; Chen, J.; Chen, H.; Yin, J. Understanding the roles of climate change, land use and land cover change and water diversion project in modulating water- and carbon-use efficiency in Han River Basin. J. Environ. Manag. 2025, 373, 123445. [Google Scholar] [CrossRef] [PubMed]
  487. Xu, C.; Chen, X.; Yu, Q.; Avirmed, B.; Zhao, J.; Liu, W.; Sun, W. Relationship between ecological spatial network and vegetation carbon use efficiency in the Yellow River Basin, China. GISci. Remote Sens. 2024, 61, 2318070. [Google Scholar] [CrossRef]
  488. Nair, R.; Luo, Y.; El-Madany, T.; Rolo, V.; Pacheco-Labrador, J.; Caldararu, S.; Morris, K.A.; Schrumpf, M.; Carrara, A.; Moreno, G.; et al. Nitrogen availability and summer drought, but not N:P imbalance, drive carbon use efficiency of a Mediterranean tree-grass ecosystem. Glob. Change Biol. 2024, 30, e17486. [Google Scholar] [CrossRef] [PubMed]
  489. Wu, J.; Gao, Z.; Liu, Q.; Li, Z.; Zhong, B. Methods for sandy land detection based on multispectral remote sensing data. Geoderma 2018, 316, 89–99. [Google Scholar] [CrossRef]
  490. Di Raimo, L.A.D.L.; Couto, E.G.; Poppiel, R.R.; Mello, D.C.d.; Amorim, R.S.S.; Torres, G.N.; Bocuti, E.D.; Veloso, G.V.; Fernandes-Filho, E.I.; Francelino, M.R.; et al. Sand subfractions by proximal and satellite sensing: Optimizing agricultural expansion in tropical sandy soils. Catena 2024, 234, 107604. [Google Scholar] [CrossRef]
  491. Secu, C.V.; Stoleriu, C.C.; Lesenciuc, C.D.; Ursu, A. Normalized Sand Index for Identification of Bare Sand Areas in Temperate Climates Using Landsat Images, Application to the South of Romania. Remote Sens. 2022, 14, 3802. [Google Scholar] [CrossRef]
  492. Meng, J.; Chu, N.; Luo, C.; Liu, H.; Li, X. High-Resolution Mapping of Topsoil Sand Content in Planosol Regions Using Temporal and Spectral Feature Optimization. Remote Sens. 2025, 17, 553. [Google Scholar] [CrossRef]
  493. Espinoza, N.S.; Piedade, M.T.F.; Demarchi, L.O.; Lima, G.R.; Resende, A.F.; Ferreira, R.R.; Silva, F.A.G.; Machado, L.A.T.; Schörgart, J. Detection of white sand patches in central Amazonia using remote sensing and meteorological data. Int. J. Remote Sens. 2025, 46, 3446–3465. [Google Scholar] [CrossRef]
  494. Lou, H.; Yang, S.; Zhao, C.; Shi, L.; Wu, L.; Wang, Y.; Wang, Z. Detecting and analyzing soil phosphorus loss associated with critical source areas using a remote sensing approach. Sci. Total Environ. 2016, 573, 397–408. [Google Scholar] [CrossRef]
  495. Aziz, D.; Rafiq, S.; Saini, P.; Ahad, I.; Gonal, B.; Rehman, S.A.; Rashid, S.; Saini, P.; Rohela, G.K.; Aalum, K.; et al. Remote sensing and artificial intelligence: Revolutionizing pest management in agriculture. Front. Sustain. Food Syst. 2025, 9, 1551460. [Google Scholar] [CrossRef]
  496. Bautista, A.S.; Tarrazó-Serrano, D.; Uris, A.; Blesa, M.; Estruch-Guitart, V.; Castiñeira-Ibáñez, S.; Rubio, C. Remote Sensing Evaluation Drone Herbicide Application Effectiveness for Controlling Echinochloa spp. in Rice Crop in Valencia (Spain). Sensors 2024, 24, 804. [Google Scholar] [CrossRef]
  497. Mahlein, A.-K. Present and Future Trends in Plant Disease Detection. Plant Dis 2016, 100, 1–11. [Google Scholar]
  498. Okole, N.; Yamati, F.R.I.; Hossain, R.; Varrelmann, M.; Mahlein, A.K.; Heim, R.H.J. Hyperspectral signatures and betalain indicator for beet mosaic virus infection in sugar beet. In Proceedings of the 2023 IEEE International Workshop on Metrology for Agriculture and Forestry (MetroAgriFor), Pisa, Italy, 6–8 November 2023; pp. 506–511. [Google Scholar]
  499. Okole, N.; Ispizua Yamati, F.R.; Hossain, R.; Varrelmann, M.; Mahlein, A.K.; Heim, R.H.J. Aerial low-altitude remote sensing and deep learning for in-field disease incidence scoring of virus yellows in sugar beet. Plant Pathol. 2024, 73, 2310–2324. [Google Scholar] [CrossRef]
  500. Hossain, R.; Ispizua Yamati, F.R.; Barreto, A.; Savian, F.; Varrelmann, M.; Mahlein, A.K.; Paulus, S. Elucidation of turnip yellows virus (TuYV) spectral reflectance pattern in Nicotiana benthamiana by non-imaging sensor technology. J. Plant Dis. Prot. 2023, 130, 35–43. [Google Scholar] [CrossRef]
  501. Barreto, A.; Ispizua Yamati, F.R.; Varrelmann, M.; Paulus, S.; Mahlein, A.K. Disease Incidence and Severity of Cercospora Leaf Spot in Sugar Beet Assessed by Multispectral Unmanned Aerial Images and Machine Learning. Plant Dis. 2023, 107, 188–200. [Google Scholar] [CrossRef]
  502. Isip, M.F.; Alberto, R.T.; Biagtan, A.R. Exploring vegetation indices adequate in detecting twister disease of onion using Sentinel-2 imagery. Spat. Inf. Res. 2020, 28, 369–375. [Google Scholar] [CrossRef]
  503. Navrozidis, L.; Alexandridis, T.K.; Moshou, D.; Pantazi, X.E.; Alexandra Tamouridou, A.; Kozhukh, D.; Castef, F.; Lagopodi, A.; Zartaloudis, Z.; Mourelatos, S.; et al. Olive Trees Stress Detection Using Sentinel-2 Images. In Proceedings of the 2019 IEEE International Geoscience and Remote Sensing Symposium, Yokohama, Japan, 28 July–2 August 2019; pp. 7220–7223. [Google Scholar]
  504. Safari, M.M.; Malian, A. Plant disease mapping in paddy growing stages using remotely sensed data. Environ. Earth Sci. 2025, 84, 1. [Google Scholar] [CrossRef]
  505. Taneja, A.; Rani, S. Target localization and communication for remote sensing with use case of self-supervised learning in plant disease detection. Int. J. Remote Sens. 2025, 1–22. [Google Scholar] [CrossRef]
  506. Zhu, H.; Lin, C.; Liu, G.; Wang, D.; Qin, S.; Li, A.; Xu, J.L.; He, Y. Intelligent agriculture: Deep learning in UAV-based remote sensing imagery for crop diseases and pests detection. Front. Plant Sci. 2024, 15, 1435016. [Google Scholar] [CrossRef] [PubMed]
  507. Zheng, Q.; Huang, W.; Xia, Q.; Dong, Y.; Ye, H.; Jiang, H.; Chen, S.; Huang, S. Remote Sensing Monitoring of Rice Diseases and Pests from Different Data Sources: A Review. Agronomy 2023, 13, 1851. [Google Scholar] [CrossRef]
  508. Zhang, T.; Cai, Y.; Zhuang, P.; Li, J. Remotely Sensed Crop Disease Monitoring by Machine Learning Algorithms: A Review. Unmanned Syst. 2024, 12, 161–171. [Google Scholar] [CrossRef]
  509. Schmidtlein, S. Imaging spectroscopy as a tool for mapping Ellenberg indicator values. J. Appl. Ecol. 2005, 42, 966–974. [Google Scholar] [CrossRef]
  510. Schmidt, J.; Fassnacht, F.E.; Lausch, A.; Schmidtlein, S. Assessing the functional signature of heathland landscapes via hyperspectral remote sensing. Ecol. Indic. 2017, 73, 505–512. [Google Scholar] [CrossRef]
  511. Schmidtlein, S.; Feilhauer, H.; Bruelheide, H. Mapping plant strategy types using remote sensing. J. Veg. Sci. 2012, 23, 395–405. [Google Scholar] [CrossRef]
  512. Cushnahan, T.A.; Grafton, M.C.E.; Pearson, D.; Ramilan, T. Hyperspectral Data Can Classify Plant Functional Groups Within New Zealand Hill Farm Pasture. Remote Sens. 2025, 17, 1120. [Google Scholar] [CrossRef]
  513. Kanta, C.; Kumar, A.; Chauhan, A.; Singh, H.; Sharma, I.P. The Interplay Between Plant Functional Traits and Climate Change; Springer: Berlin/Heidelberg, Germany, 2024; ISBN 9789819715107. [Google Scholar]
  514. Schweiger, A.K.; Schütz, M.; Risch, A.C.; Kneubühler, M.; Haller, R.; Schaepman, M.E. How to predict plant functional types using imaging spectroscopy: Linking vegetation community traits, plant functional types and spectral response. Methods Ecol. Evol. 2017, 8, 86–95. [Google Scholar] [CrossRef]
  515. Möckel, T.; Löfgren, O.; Prentice, H.C.; Eklundh, L.; Hall, K. Airborne hyperspectral data predict Ellenberg indicator values for nutrient and moisture availability in dry grazed grasslands within a local agricultural landscape. Ecol. Indic. 2016, 66, 503–516. [Google Scholar] [CrossRef]
  516. Klinke, R.; Kuechly, H.; Frick, A.; Förster, M.; Schmidt, T.; Holtgrave, A.K.; Kleinschmit, B.; Spengler, D.; Neumann, C. Indicator-Based Soil Moisture Monitoring of Wetlands by Utilizing Sentinel and Landsat Remote Sensing Data. PFG-J. Photogramm. Remote Sens. Geoinf. Sci. 2018, 86, 71–84. [Google Scholar] [CrossRef]
  517. Peng, Y.; Kira, O.; Nguy-Robertson, A.; Suyker, A.; Arkebauer, T.; Sun, Y.; Gitelson, A.A. Gross primary production estimation in crops using solely remotely sensed data. Agron. J. 2019, 111, 2981–2990. [Google Scholar] [CrossRef]
  518. Reeves, M.C.; Zhao, M.; Running, S.W. Usefulness and limits of MODIS GPP for estimating wheat yield. Int. J. Remote Sens. 2005, 26, 1403–1421. [Google Scholar] [CrossRef]
  519. Maleki, M.; Arriga, N.; Barrios, J.M.; Wieneke, S.; Liu, Q.; Peñuelas, J.; Janssens, I.A.; Balzarolo, M. Estimation of Gross Primary Productivity (GPP) Phenology of a Short-Rotation Plantation Using Remotely Sensed Indices Derived from Sentinel-2 Images. Remote Sens. 2020, 12, 2104. [Google Scholar] [CrossRef]
  520. Celis, J.; Xiao, X.; White, P.M.; Cabral, O.M.R.; Freitas, H.C. Improved Modeling of Gross Primary Production and Transpiration of Sugarcane Plantations with Time-Series Landsat and Sentinel-2 Images. Remote Sens. 2023, 16, 46. [Google Scholar] [CrossRef]
  521. Shirkey, G.; John, R.; Chen, J.; Dahlin, K.; Abraha, M.; Sciusco, P.; Lei, C.; Reed, D.E. Fine resolution remote sensing spectra improves estimates of gross primary production of croplands. Agric. For. Meteorol. 2022, 326, 109175. [Google Scholar] [CrossRef]
  522. Cicuéndez, V.; Inclán, R.; Sánchez-Cañete, E.P.; Román-Cascón, C.; Sáenz, C.; Yagüe, C. Modeling Gross Primary Production (GPP) of a Mediterranean Grassland in Central Spain Using Sentinel-2 NDVI and Meteorological Field Information. Agronomy 2024, 14, 1243. [Google Scholar] [CrossRef]
  523. He, L.; Mostovoy, G. Cotton yield estimate using Sentinel-2 data and an ecosystem model over the southern US. Remote Sens. 2019, 11, 2000. [Google Scholar] [CrossRef]
  524. Wang, T.; Zhang, Y.; Yue, C.; Wang, Y.; Wang, X.; Lyu, G.; Wei, J.; Yang, H.; Piao, S. Progress and challenges in remotely sensed terrestrial carbon fluxes. Geo-Spat. Inf. Sci. 2024, 28, 1–21. [Google Scholar] [CrossRef]
  525. Yan, Y.; Xu, X.; Liu, X.; Wen, Y.; Ou, J. Assessing the contributions of climate change and human activities to cropland productivity by means of remote sensing. Int. J. Remote Sens. 2020, 41, 2004–2021. [Google Scholar] [CrossRef]
  526. Feng, X.; Liu, G.; Chen, J.M.; Chen, M.; Liu, J.; Ju, W.M.; Sun, R.; Zhou, W. Net primary productivity of China’s terrestrial ecosystems from a process model driven by remote sensing. J. Environ. Manag. 2007, 85, 563–573. [Google Scholar] [CrossRef] [PubMed]
  527. Li, C.; Han, W.; Peng, M. Improving the spatial and temporal estimating of daytime variation in maize net primary production using unmanned aerial vehicle-based remote sensing. Int. J. Appl. Earth Obs. Geoinf. 2021, 103, 102467. [Google Scholar] [CrossRef]
  528. Niedertscheider, M.; Kastner, T.; Fetzel, T.; Haberl, H.; Kroisleitner, C.; Plutzar, C.; Erb, K.H. Mapping and analysing cropland use intensity from a NPP perspective. Environ. Res. Lett. 2016, 11, 014008. [Google Scholar] [CrossRef]
  529. Xu, L.; Zhao, Z.; Wang, C.; Wang, H.; Ma, C. Quantitative estimation of net primary productivity by an improved tCASA model using Landsat time series data: A case study of Central Plains, China. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2025, 18, 10403–10418. [Google Scholar] [CrossRef]
  530. Li, J.; Xu, D.; Xu, Z.; Wang, Y.; Yang, J. Analysis of Spatio-Temporal Variation of Vegetation Npp and its Driving Factors in Tianshan Mountains Based on Casa Model. SSRN 2025. [Google Scholar]
  531. Baeza, S.; Paruelo, J.M. Spatial and temporal variation of human appropriation of net primary production in the Rio de la Plata grasslands. ISPRS J. Photogramm. Remote Sens. 2018, 145, 238–249. [Google Scholar] [CrossRef]
  532. Liu, Y.; Song, W. Mapping human appropriation of net primary production in agroecosystems in the Heihe River Basin, China. Agric. Ecosyst. Environ. 2022, 335, 107996. [Google Scholar] [CrossRef]
  533. Paudel, S.; Mueller, K.; Ovando-Montejo, G.; Tango, L.; Rushforth, R.; Lant, C. A dataset cataloging product-specific human appropriation of net primary production (HANPP) in US counties. Data Br. 2023, 50, 109530. [Google Scholar] [CrossRef]
  534. Matej, S.; Weidinger, F.; Kaufmann, L.; Roux, N.; Gingrich, S.; Haberl, H.; Krausmann, F.; Erb, K.H. A global land-use data cube 1992–2020 based on the Human Appropriation of Net Primary Production. Sci. Data 2025, 12, 511. [Google Scholar] [CrossRef] [PubMed]
  535. Zhao, J.; Sun, X.; Wang, M.; Li, G.; Hou, X. Crop mapping and quantitative evaluation of cultivated land use intensity in Shandong Province, 2018–2022. L. Degrad. Dev. 2024, 35, 4648–4665. [Google Scholar] [CrossRef]
  536. Cai, W.; Ullah, S.; Yan, L.; Lin, Y. Remote sensing of ecosystem water use efficiency: A review of direct and indirect estimation methods. Remote Sens. 2021, 13, 2393. [Google Scholar] [CrossRef]
  537. Yu, X.; Yin, Q.; Zhao, T.; Chen, K.; Qian, L.; Wang, W.; Hu, X.; Zhang, B. Detecting ecosystem water use efficiency responses to drought from long-term remote sensing data. Ecol. Indic. 2025, 177, 113734. [Google Scholar] [CrossRef]
  538. Rao, S.; Kisekka, I. Science of the Total Environment Carbon—Water coupling in California almond orchards: A multi-scale assessment of ecosystem water use efficiency using eddy covariance and remote sensing. Sci. Total Environ. 2025, 990, 179914. [Google Scholar]
  539. Jiang, L.; Zhang, F.; Chi, J.; Yan, P.; Bu, X.; He, Y.; Bai, T. Evaluation of pear orchard yield and water use efficiency at the field scale by assimilating remotely sensed LAI and SM into the WOFOST model. Comput. Electron. Agric. 2025, 233, 110145. [Google Scholar] [CrossRef]
  540. He, J.; Zhou, Y.; Liu, X.; Duan, W.; Pan, N. Spatiotemporal Changes in Water-Use Efficiency of China’s Terrestrial Ecosystems During 2001–2020 and the Driving Factors. Remote Sens. 2025, 17, 136. [Google Scholar] [CrossRef]
  541. Ali, A.; Kaul, H.P. Monitoring Yield and Quality of Forages and Grassland in the View of Precision Agriculture Applications—A Review. Remote Sens. 2025, 17, 279. [Google Scholar] [CrossRef]
  542. Guillevic, P.C.; Aouizerats, B.; Burger, R.; Den Besten, N.; Jackson, D.; Ridderikhoff, M.; Zajdband, A.; Houborg, R.; Franz, T.E.; Robertson, G.P.; et al. Planet’s Biomass Proxy for monitoring aboveground agricultural biomass and estimating crop yield. F. Crop. Res. 2024, 316, 109511. [Google Scholar] [CrossRef]
  543. Reinermann, S.; Boos, C.; Kaim, A.; Schucknecht, A.; Asam, S.; Annuth, S.H.; Schmitt, T.M.; Koellner, T.; Kiese, R. Grassland yield estimations-potentials and limitations of remote sensing, process-based modelling and field measurements. EGUsphere 2025. [Google Scholar]
  544. Åström, O.; Månsson, S.; Lazar, I.; Nilsson, M.; Ekelöf, J.; Oxenstierna, A.; Sopasakis, A. Predicting Intra-Field Yield Variations for Winter Wheat Using Remote Sensing and Graph Attention Networks. SSRN Electron. J. 2025, 237, 110499. [Google Scholar] [CrossRef]
  545. Muruganantham, P.; Wibowo, S.; Grandhi, S.; Samrat, N.H.; Islam, N. A Systematic Literature Review on Crop Yield Prediction with Deep Learning and Remote Sensing. Remote Sens. 2022, 14, 1990. [Google Scholar] [CrossRef]
  546. 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]
  547. Filippi, P.; Jones, E.J.; Ginns, B.J.; Whelan, B.M.; Roth, G.W.; Bishop, T.F.A. Mapping the depth-to-soil pH constraint, and the relationship with cotton and grain yield at the within-field scale. Agronomy 2019, 9, 251. [Google Scholar] [CrossRef]
  548. Li, M.; Wang, J.; Li, K.; Liu, Y.; Ochir, A.; Davaasuren, D. Assessment of grazing livestock balance on the Eastern Mongolian Plateau based on remote sensing monitoring and grassland carrying capacity evaluation. Sci. Rep. 2024, 14, 32151. [Google Scholar] [CrossRef]
  549. Ren, J.; Zhang, N.; Liu, X.; Wu, S.; Li, D. Dynamic Harvest Index Estimation of Winter Wheat Based on UAV Hyperspectral Remote Sensing Considering Crop Aboveground Biomass Change and the Grain Filling Process. Remote Sens. 2022, 14, 1955. [Google Scholar] [CrossRef]
  550. Xu, J.; Meng, J.; Quackenbush, L.J. Use of remote sensing to predict the optimal harvest date of corn. F. Crop. Res. 2019, 236, 1–13. [Google Scholar] [CrossRef]
  551. Li, H.; Luo, Y.; Xue, X.; Zhao, Y.; Zhao, H.; Li, F. A comparison of harvest index estimation methods of winter wheat based on field measurements of biophysical and spectral data. Biosyst. Eng. 2011, 109, 396–403. [Google Scholar] [CrossRef]
  552. Yue, J.; Yao, Y.; Shen, J.; Li, T.; Xu, N.; Feng, H.; Wei, Y.; Xu, X.; Lin, Y.; Guo, W.; et al. Winter wheat harvest detection via Sentinel-2 MSI images. Int. J. Remote Sens. 2025, 46, 2482–2500. [Google Scholar] [CrossRef]
  553. Paz-Kagan, T.; Zaady, E.; Salbach, C.; Schmidt, A.; Lausch, A.; Zacharias, S.; Notesco, G.; Ben-Dor, E.; Karnieli, A. Mapping the spectral soil quality index (SSQI) using airborne imaging spectroscopy. Remote Sens. 2015, 7, 15748–15781. [Google Scholar] [CrossRef]
  554. Baroudy, A.A.E.; Ali, A.M.; Mohamed, E.S.; Moghanm, F.S.; Shokr, M.S.; Savin, I.; Poddubsky, A.; Ding, Z.; Kheir, A.M.S.; Aldosari, A.A.; et al. Modeling Land Suitability for Rice Crop Using Remote Sensing and Soil Quality Indicators: The Case Study of the Nile Delta. Sustainability 2020, 12, 9653. [Google Scholar] [CrossRef]
  555. Kumar, U.S.; Kapali, B.S.C.; Nageswaran, A.; Umapathy, K.; Jangir, P.; Swetha, K.; Begum, M.A. Fusion of MobileNet and GRU: Enhancing Remote Sensing Applications for Sustainable Agriculture and Food Security. Remote Sens. Earth Syst. Sci. 2024, 8, 118–131. [Google Scholar] [CrossRef]
  556. Dedeoğlu, M.; Başayiğit, L.; Yüksel, M.; Kaya, F. Assessment of the vegetation indices on Sentinel-2A images for predicting the soil productivity potential in Bursa, Turkey. Environ. Monit. Assess. 2020, 192, 16. [Google Scholar] [CrossRef] [PubMed]
  557. Fadl, M.E.; AbdelRahman, M.A.E.; El-Desoky, A.I.; Sayed, Y.A. Assessing soil productivity potential in arid region using remote sensing vegetation indices. J. Arid Environ. 2024, 222, 105166. [Google Scholar] [CrossRef]
  558. Rieser, J.; Veste, M.; Thiel, M.; Schönbrodt-stitt, S. Coverage and rainfall response of biological soil crusts using multi-temporal sentinel-2 data in a central european temperate dry acid grassland. Remote Sens. 2021, 13, 3093. [Google Scholar] [CrossRef]
  559. Crucil, G.; Van Oost, K. Towards mapping of soil crust using multispectral imaging. Sensors 2021, 21, 1850. [Google Scholar] [CrossRef]
  560. Weber, B.; Hill, J. Remote Sensing of Biological Soil Crusts at Different Scales. In Biological Soil Crusts: An Organizing Principle in Drylands; Springer: Berlin/Heidelberg, Germany, 2016; pp. 215–234. [Google Scholar]
  561. Crucil, G. Characterizing soil Physical Crusts in Loamy Soils Using Spectral Data Processing Techniques from Proximal and Remote Sensing. Ph.D. Thesis, UCL-Université Catholique de Louvain, Ottignies-Louvain-la-Neuve, Belgique, 2022. [Google Scholar]
  562. Han, H.Q.; Yin, C.Y.; Wang, K.; Zhou, H.Y. Progress and Prospects of Research on Karst Ecosystem Services. Appl. Ecol. Environ. Res. 2025, 23, 507–529. [Google Scholar] [CrossRef]
  563. Wu, C.; Wu, Z.; Wang, Y.; Yang, Y. Effect of soil crust on the prediction of soil organic matter based on soil colour. Catena 2025, 251, 108818. [Google Scholar] [CrossRef]
  564. Beaugendre, N.; Malam Issa, O.; Choné, A.; Cerdan, O.; Desprats, J.F.; Rajot, J.L.; Sannier, C.; Valentin, C. Developing a predictive environment-based model for mapping biological soil crust patterns at the local scale in the Sahel. Catena 2017, 158, 250–265. [Google Scholar] [CrossRef]
  565. Brom, J.; Duffková, R.; Haberle, J.; Zajíček, A.; Nedbal, V.; Bernasová, T.; Křováková, K. Identification of infiltration features and hydraulic properties of soils based on crop water stress derived from remotely sensed data. Remote Sens. 2021, 13, 4127. [Google Scholar] [CrossRef]
  566. Francos, N.; Romano, N.; Nasta, P.; Zeng, Y.; Szabó, B.; Manfreda, S.; Ciraolo, G.; Mészáros, J.; Zhuang, R.; Su, B.; et al. Mapping water infiltration rate using ground and uav hyperspectral data: A case study of alento, italy. Remote Sens. 2021, 13, 2606. [Google Scholar] [CrossRef]
  567. Geng, Y.; Zhou, T.; Zhang, Z.; Cui, B.; Sun, J.; Zeng, L.; Yang, R.; Wu, N.; Liu, T.; Pan, J.; et al. Continental-scale mapping of soil pH with SAR-optical fusion based on long-term earth observation data in google earth engine. Ecol. Indic. 2024, 165, 112246. [Google Scholar] [CrossRef]
  568. Zhang, Y.; Sui, B.; Shen, H.; Wang, Z. Estimating temporal changes in soil pH in the black soil region of Northeast China using remote sensing. Comput. Electron. Agric. 2018, 154, 204–212. [Google Scholar] [CrossRef]
  569. Jia, P.; Shang, T.; Zhang, J.; Sun, Y. Inversion of soil pH during the dry and wet seasons in the Yinbei region of Ningxia, China, based on multi-source remote sensing data. Geoderma Reg. 2021, 25, e00399. [Google Scholar] [CrossRef]
  570. Webb, H.; Barnes, N.; Powell, S.; Jones, C. Does drone remote sensing accurately estimate soil pH in a spring wheat field in southwest Montana? Precis. Agric. 2021, 22, 1803–1815. [Google Scholar] [CrossRef]
  571. Wang, Z.; Wu, W.; Liu, H. Comparing Soil pH Mapping from Multi-Temporal PlanetScope and Sentinel-2 Data Across Land Use Types. Remote Sens. 2025, 17, 189. [Google Scholar] [CrossRef]
  572. Molaeinasab, A.; Tarkesh, M.; Bashari, H.; Toomanian, N.; Aghasi, B.; Jalalian, A. Spatial modeling of soil chemical properties in an arid region of Central Iran using machine learning and remote sensing data. Model. Earth Syst. Environ. 2025, 11, 152. [Google Scholar] [CrossRef]
  573. Srivastava, P.K.; Srivastava, S.; Singh, P.; Gupta, A.; Dugesar, V. Soil chemical properties estimation using hyperspectral remote sensing: A review. In Earth Observation for Monitoring and Modeling Land Use; Elsevier: Amsterdam, The Netherlands, 2025; pp. 25–43. [Google Scholar]
  574. Rogovska, N.; Blackmer, A.M. Remote sensing of soybean canopy as a tool to map high pH, calcareous soils at field scale. Precis. Agric. 2009, 10, 175–187. [Google Scholar] [CrossRef]
  575. Khanal, S.; Fulton, J.; Klopfenstein, A.; Douridas, N.; Shearer, S. Integration of high resolution remotely sensed data and machine learning techniques for spatial prediction of soil properties and corn yield. Comput. Electron. Agric. 2018, 153, 213–225. [Google Scholar] [CrossRef]
  576. Šestak, I.; Mihaljevski Boltek, L.; Mesić, M.; Zgorelec, Ž.; Perčin, A. Hyperspectral sensing of soil ph, total carbon and total nitrogen content based on linear and non-linear calibration methods. J. Cent. Eur. Agric. 2019, 20, 504–523. [Google Scholar] [CrossRef]
  577. Jain, S.; Sethia, D.; Tiwari, K.C. Developing novel spectral indices for precise estimation of soil pH and organic carbon with hyperspectral data and machine learning. Environ. Monit. Assess. 2024, 196, 1255. [Google Scholar] [CrossRef]
  578. Scudiero, E.; Corwin, D.L.; Anderson, R.G.; Yemoto, K.; Clary, W.; Wang, Z.; Skaggs, T.H. Remote sensing is a viable tool for mapping soil salinity in agricultural lands. Calif. Agric. 2017, 71, 231–238. [Google Scholar] [CrossRef]
  579. Bai, L.; Wang, C.; Zang, S.; Wu, C.; Luo, J.; Wu, Y. Mapping soil alkalinity and salinity in northern songnen plain, China with the hj-1 hyperspectral imager data and partial least squares regression. Sensors 2018, 18, 3855. [Google Scholar] [CrossRef] [PubMed]
  580. Metternicht, G.I.; Zinck, J.A. Remote sensing of soil salinity: Potentials and constraints. Remote Sens. Environ. 2003, 85, 1–20. [Google Scholar] [CrossRef]
  581. Lazzeri, G.; Milewski, R.; Förster, S.; Moretti, S.; Chabrillat, S. Early Detection of Soil Salinization by Means of Spaceborne Hyperspectral Imagery. Remote Sens. 2025, 17, 2486. [Google Scholar] [CrossRef]
  582. Gorji, T.; Sertel, E.; Tanik, A. Monitoring soil salinity via remote sensing technology under data scarce conditions: A case study from Turkey. Ecol. Indic. 2017, 74, 384–391. [Google Scholar] [CrossRef]
  583. GORJİ, T.; YILDIRIM, A.; SERTEL, E.; TANIK, A. Remote sensing approaches and mapping methods for monitoring soil salinity under different climate regimes. Int. J. Environ. Geoinform. 2019, 6, 33–49. [Google Scholar] [CrossRef]
  584. Aldabaa, A.A.A.; Weindorf, D.C.; Chakraborty, S.; Sharma, A.; Li, B. Combination of proximal and remote sensing methods for rapid soil salinity quantification. Geoderma 2015, 239, 34–46. [Google Scholar] [CrossRef]
  585. Jingwei, W.; Vincent, B.; Jinzhong, Y.; Bouarfa, S.; Vidal, A. Remote sensing monitoring of changes in soil salinity: A case study in inner Mongolia, China. Sensors 2008, 8, 7035–7049. [Google Scholar] [CrossRef]
  586. AbdelRahman, M.A.E. An Overview of Land Degradation, Desertification and Sustainable Land Management Using GIS and Remote Sensing Applications; Springer International Publishing: Cham, Switzerland, 2023; Volume 34, ISBN 0123456789. [Google Scholar]
  587. Ruf, T.; Gilcher, M.; Udelhoven, T.; Emmerling, C. Implications of bioenergy cropping for soil: Remote sensing identification of silage maize cultivation and risk assessment concerning soil erosion and compaction. Land 2021, 10, 128. [Google Scholar] [CrossRef]
  588. Liu, S.; Dong, A.; Niu, B.; Xu, F.; Xu, J.; Yin, L.; Wang, S. Maize/cover crop intercropping mitigates soil erosion and enhances yield of ridge cultivation in Chinese Mollisol region. Catena 2025, 255, 109012. [Google Scholar] [CrossRef]
  589. Fahd, S.; Waqas, M.; Zafar, Z.; Soufan, W.; Almutairi, K.F.; Tariq, A. Integration of RUSLE model with remotely sensed data over Google Earth Engine to evaluate soil erosion in Central Indus Basin. Earth Surf. Process. Landf. 2025, 50, e70019. [Google Scholar] [CrossRef]
  590. Zhang, X.; Qin, C.; Ma, S.; Liu, J.; Wang, Y.; Liu, H.; An, Z.; Ma, Y. Study on the Extraction of Topsoil-Loss Areas of Cultivated Land Based on Multi-Source Remote Sensing Data. Remote Sens. 2025, 17, 547. [Google Scholar] [CrossRef]
  591. Gammoudi, A.; Guesmi, H.; Tebini, A.; Attia, R.; Chahed, T.S.; Trabelsi, H. Soil degradation risk prediction in an arid region (Northern Tataouine, Tunisia): Using an empirical model coupling with remote sensing and GIS. Environ. Res. Commun. 2025, 7, 015019. [Google Scholar] [CrossRef]
  592. Molua, O.C.; Ukpene, A.O.; Ighrakpata, F.C.; Emagbetere, J.U.; Nwachuku, D.N. Review on Nondestructive Methods of Detecting Compacted Soils and Effects of Compacted Soil on Crop Production. Open J. Agric. Sci. 2023, 4, 1–16. [Google Scholar] [CrossRef]
  593. Bento, N.L.; Ferraz, G.A.e.S.; Santana, L.S.; Faria, R.D.O.; Rossi, G.; Bambi, G. Plant Height and Soil Compaction in Coffee Crops Based on LiDAR and RGB Sensors Carried by Remotely Piloted Aircraft. Remote Sens. 2025, 17, 1445. [Google Scholar] [CrossRef]
  594. Mikhailova, E.A.; Post, C.J.; Zurqani, H.A.; Hutton, P.C.; Nelson, D.G. Enriching Earth Science Education with Direct and Proximal Remote Sensing of Soil Using a Mobile Geospatial Application. Earth 2025, 6, 8. [Google Scholar] [CrossRef]
  595. Lambot, S.; Wu, K.; Bates, J.; Fluhrer, A.; Montzka, C.; Henrion, M.; Peichl, M.; Dill, S.; Bruecker, P.; Engel, M.; et al. Multi-Sensor Drone Fleet for Soil and Crop Analyses: Integrating Gpr, Sar, Lidar, Multispectral and Infrared Imaging. SSRN 2025. [Google Scholar]
  596. Kulkarni, S.G.; Bajwa, S.G.; Huitink, G. Investigation of the effects of soil compaction in cotton. Trans. ASABE 2010, 53, 667–674. [Google Scholar] [CrossRef]
  597. Bento, N.L.; Silva Ferraz, G.A.e.; Santana, L.S.; de Oliveira Faria, R.; da Silva Amorim, J.; de Lourdes Oliveira e Silva, M.; Silva, M.M.A.; Alonso, D.J.C. Soil compaction mapping by plant height and spectral responses of coffee in multispectral images obtained by remotely piloted aircraft system. Precis. Agric. 2024, 25, 729–750. [Google Scholar] [CrossRef]
  598. Milewski, R.; Schmid, T.; Chabrillat, S.; Jiménez, M.; Escribano, P.; Pelayo, M.; Ben-Dor, E. Analyses of the Impact of Soil Conditions and Soil Degradation on Vegetation Vitality and Crop Productivity Based on Airborne Hyperspectral VNIR–SWIR–TIR Data in a Semi-Arid Rainfed Agricultural Area (Camarena, Central Spain). Remote Sens. 2022, 14, 5131. [Google Scholar] [CrossRef]
  599. Serrano, J.; Marques, J.; Shahidian, S.; Carreira, E.; Marques da Silva, J.; Paixão, L.; Paniagua, L.L.; Moral, F.; Ferraz de Oliveira, I.; Sales-Baptista, E. Sensing and Mapping the Effects of Cow Trampling on the Soil Compaction of the Montado Mediterranean Ecosystem. Sensors 2023, 23, 888. [Google Scholar] [CrossRef]
  600. Kuemmerle, T.; Fernandez, P.D.; Baumann, M.; Burton, J. Uncovering patterns of cattle intensification across South America’s dry diagonal. Environ. Res. Lett. 2025, 20, 074004. [Google Scholar] [CrossRef]
  601. Gómez Giménez, M.; de Jong, R.; Della Peruta, R.; Keller, A.; Schaepman, M.E. Determination of grassland use intensity based on multi-temporal remote sensing data and ecological indicators. Remote Sens. Environ. 2017, 198, 126–139. [Google Scholar] [CrossRef]
  602. De Vroey, M.; Radoux, J.; Defourny, P. Grassland Mowing Detection Using Sentinel-1 Time Series: Potential and Limitations. Remote Sens. 2021, 13, 348. [Google Scholar] [CrossRef]
  603. Potočnik Buhvald, A.; Račič, M.; Immitzer, M.; Oštir, K.; Veljanovski, T. Grassland Use Intensity Classification Using Intra-Annual Sentinel-1 and -2 Time Series and Environmental Variables. Remote Sens. 2022, 14, 3387. [Google Scholar] [CrossRef]
  604. Rivas, H.; Touchais, H.; Thierion, V.; Millet, J.; Curtet, L.; Fauvel, M. Nationwide operational mapping of grassland first mowing dates combining machine learning and Sentinel-2 time series. Remote Sens. Environ. 2024, 315, 114476. [Google Scholar] [CrossRef]
  605. Bekkema, M.E.; Eleveld, M. Mapping Grassland Management Intensity Using Sentinel-2 Satellite Data. GI_Forum 2018, 1, 194–213. [Google Scholar] [CrossRef]
  606. Wu, R.; Hong, Z.; Du, W.; Shan, Y.; Ying, H.; Wu, R.; Gantumur, B. A Generalized Spatiotemporally Weighted Boosted Regression to Predict the Occurrence of Grassland Fires in the Mongolian Plateau. Remote Sens. 2025, 17, 1485. [Google Scholar] [CrossRef]
  607. Molema, T.R.; Tesfamichael, S.G.; Fundisi, E. Optical and radar remote sensing for burn scar mapping in the grassland biome. Remote Sens. Appl. Soc. Environ. 2025, 38, 101548. [Google Scholar] [CrossRef]
  608. Mofokeng, O.D.; Adelabu, S.A.; Durowoju, O.S.; Adagbasa, E.A. Grass curing-driven fire danger index in a protected mountainous grassland using fused MODIS and Sentinel-2. Int. J. Remote Sens. 2024, 45, 5359–5384. [Google Scholar] [CrossRef]
  609. Hu, X.; Jiang, F.; Qin, X.; Huang, S.; Yang, X.; Meng, F. An Optimized Smoke Segmentation Method for Forest and Grassland Fire Based on the UNet Framework. Fire 2024, 7, 68. [Google Scholar] [CrossRef]
  610. Marčiš, M.; Fraštia, M.; Lieskovský, T.; Ambroz, M.; Mikula, K. Photogrammetric Measurement of Grassland Fire Spread: Techniques and Challenges with Low-Cost Unmanned Aerial Vehicles. Drones 2024, 8, 282. [Google Scholar] [CrossRef]
  611. Watzig, C.; Schaumberger, A.; Klingler, A.; Dujakovic, A.; Atzberger, C.; Vuolo, F. Grassland cut detection based on Sentinel-2 time series to respond to the environmental and technical challenges of the Austrian fodder production for livestock feeding. Remote Sens. Environ. 2023, 292, 113577. [Google Scholar] [CrossRef]
  612. Reuß, F.; Navacchi, C.; Greimeister-Pfeil, I.; Vreugdenhil, M.; Schaumberger, A.; Klingler, A.; Mayer, K.; Wagner, W. Evaluation of limiting factors for SAR backscatter based cut detection of alpine grasslands. Sci. Remote Sens. 2024, 9, 100117. [Google Scholar] [CrossRef]
  613. Dujakovic, A.; Watzig, C.; Schaumberger, A.; Klingler, A.; Atzberger, C.; Vuolo, F. Enhancing grassland cut detection using Sentinel-2 time series through integration of Sentinel-1 SAR and weather data. Remote Sens. Appl. Soc. Environ. 2025, 37, 101453. [Google Scholar] [CrossRef]
Figure 1. In situ and RS approaches and the five characteristics of A-LUI (trait indicators of A-LUI, genesis indicators of A-LUI, functional indicators of A-LUI, structural indicators of A-LUI, taxonomic indicators of A-LUI). Trait indicators of A-LUI are the most important link between in situ and RS monitoring approaches (modified after Lausch et al. [62]).
Figure 1. In situ and RS approaches and the five characteristics of A-LUI (trait indicators of A-LUI, genesis indicators of A-LUI, functional indicators of A-LUI, structural indicators of A-LUI, taxonomic indicators of A-LUI). Trait indicators of A-LUI are the most important link between in situ and RS monitoring approaches (modified after Lausch et al. [62]).
Agriculture 15 02233 g001
Figure 2. Different RS platforms; wireless sensor networks (WSN); WSN over lysimeters, phenotyping laboratories, global change experimental facility (GCEF), drones, towers, balloons, airborne- and spaceborne RS platforms with different RS technologies (RGB/photography, multispectral, hyperspectral, thermal, laser, RADAR, acoustic, and LiDAR) to monitor indicators of land use intensity on different spatial and temporal scales (modified from Lausch et al. [63]).
Figure 2. Different RS platforms; wireless sensor networks (WSN); WSN over lysimeters, phenotyping laboratories, global change experimental facility (GCEF), drones, towers, balloons, airborne- and spaceborne RS platforms with different RS technologies (RGB/photography, multispectral, hyperspectral, thermal, laser, RADAR, acoustic, and LiDAR) to monitor indicators of land use intensity on different spatial and temporal scales (modified from Lausch et al. [63]).
Agriculture 15 02233 g002
Figure 3. Problems of spatial resolution of RS data in the detection of crop vegetation. (a) Image of a rapeseed plant at a flight altitude of 1 m with a ground resolution of 0.6 mm per pixel. (b) Image of a rapeseed plant at a flight altitude of 5 m with a ground resolution of 1.5 mm per pixel. (c) Image of a rapeseed plant at a flight altitude of 10 m with a ground resolution of 2.5 mm per pixel. (d) Image of a rapeseed plant at a flight altitude of 20 m with a ground resolution of 5 mm per pixel. (e) Image of a rapeseed plant at a flight altitude of 40 m with a ground resolution of 10 mm per pixel. (f) Image of a rapeseed plant at a flight altitude of 80 m with a ground resolution of 20 mm per pixel (from Grenzdörffer [85]).
Figure 3. Problems of spatial resolution of RS data in the detection of crop vegetation. (a) Image of a rapeseed plant at a flight altitude of 1 m with a ground resolution of 0.6 mm per pixel. (b) Image of a rapeseed plant at a flight altitude of 5 m with a ground resolution of 1.5 mm per pixel. (c) Image of a rapeseed plant at a flight altitude of 10 m with a ground resolution of 2.5 mm per pixel. (d) Image of a rapeseed plant at a flight altitude of 20 m with a ground resolution of 5 mm per pixel. (e) Image of a rapeseed plant at a flight altitude of 40 m with a ground resolution of 10 mm per pixel. (f) Image of a rapeseed plant at a flight altitude of 80 m with a ground resolution of 20 mm per pixel (from Grenzdörffer [85]).
Agriculture 15 02233 g003
Figure 4. RS indicators capture management intensity (e.g., fertilisation, irrigation), while biophysical potential (climate, soil) is accounted for through normalisation and modelling. Land cover change dynamics are treated separately via RS time series to avoid conflating intensity with productivity or conversion signals.
Figure 4. RS indicators capture management intensity (e.g., fertilisation, irrigation), while biophysical potential (climate, soil) is accounted for through normalisation and modelling. Land cover change dynamics are treated separately via RS time series to avoid conflating intensity with productivity or conversion signals.
Agriculture 15 02233 g004
Figure 5. RS monitoring of the five characteristics of A-LUI: (I) the trait indicators of A-LUI, (II) the genesis indicators of A-LUI, (III) the structural indicators of A-LUI, (IV) the taxonomic indicators of A-LUI, and (V) the functional indicators of A-LUI. (a) Chlorophyll value; (b) phosphorus value (from Picado and Romero [86]); (c) terrace detection (from Yu et al. [87]); (d) perimeter boundaries of farmland blocks (from Wang et al. [88]); (e) shape, size, and small-scale nature of the border between Saxony-Anhalt and Lower Saxony; (f) hedgerow map classifications from an aerial photography and (g) TerraSAR-X image (from Betbeder et al. [89]); (h) wall-to-wall crop type mapping using the benchmark 10-day interval composite of Landsat and Sentinel-2 time series (from Griffiths et al. [90]), types of grassland management intensity: (i) extensive, (j) intensive (from Bartold et al. [91]); (k) disease severity prediction in sugar beet using UAV multispectral data (from Günder et al. [92]); (l) mean SOC content and (m) C:N ratio maps predicted with Sentinel-1, Sentinel-2 and Landsat-8 data (from Zhou et al. [93]), (n) the spectral fingerprint of A-LUI can be determined using spectral RS data.
Figure 5. RS monitoring of the five characteristics of A-LUI: (I) the trait indicators of A-LUI, (II) the genesis indicators of A-LUI, (III) the structural indicators of A-LUI, (IV) the taxonomic indicators of A-LUI, and (V) the functional indicators of A-LUI. (a) Chlorophyll value; (b) phosphorus value (from Picado and Romero [86]); (c) terrace detection (from Yu et al. [87]); (d) perimeter boundaries of farmland blocks (from Wang et al. [88]); (e) shape, size, and small-scale nature of the border between Saxony-Anhalt and Lower Saxony; (f) hedgerow map classifications from an aerial photography and (g) TerraSAR-X image (from Betbeder et al. [89]); (h) wall-to-wall crop type mapping using the benchmark 10-day interval composite of Landsat and Sentinel-2 time series (from Griffiths et al. [90]), types of grassland management intensity: (i) extensive, (j) intensive (from Bartold et al. [91]); (k) disease severity prediction in sugar beet using UAV multispectral data (from Günder et al. [92]); (l) mean SOC content and (m) C:N ratio maps predicted with Sentinel-1, Sentinel-2 and Landsat-8 data (from Zhou et al. [93]), (n) the spectral fingerprint of A-LUI can be determined using spectral RS data.
Agriculture 15 02233 g005
Figure 6. Mapping spatial nutritional variability of (a) sugarcane, (b) foliar chlorophyll, (c) foliar nitrogen, (d) phosphorus, (e) potassium concentrations using a MicaSense RedEdge-P camera attached to a drone and LiDAR data (from Picado and Romero [86]).
Figure 6. Mapping spatial nutritional variability of (a) sugarcane, (b) foliar chlorophyll, (c) foliar nitrogen, (d) phosphorus, (e) potassium concentrations using a MicaSense RedEdge-P camera attached to a drone and LiDAR data (from Picado and Romero [86]).
Agriculture 15 02233 g006
Figure 7. Spatial distribution of chlorophyll content over the maize field for vegetative stages based on UAV-MS data: (a) early vegetation, (b) mid vegetation, (c) late vegetation, (d) early reproductive, (e) mid reproductive, (f) late reproductive (from Brewer et al. [105]).
Figure 7. Spatial distribution of chlorophyll content over the maize field for vegetative stages based on UAV-MS data: (a) early vegetation, (b) mid vegetation, (c) late vegetation, (d) early reproductive, (e) mid reproductive, (f) late reproductive (from Brewer et al. [105]).
Agriculture 15 02233 g007
Figure 8. (a) Reflectance and (b) canopy SIF maps obtained with the HyPlant airborne sensor over an agricultural research site in Klein Altendorf, Germany. Lower SIF is evident in forests (left in lower panel) and higher SIF in dense agricultural fields (middle and right in lower panel). Fluorescence emission reveals information on vegetation status which is not visible in the reflectance domain. For example, the two fields denoted as A and B display almost identical reflectance (b), whereas their fluorescence emission is very different (a,b) (from Mohammed et al. [72]).
Figure 8. (a) Reflectance and (b) canopy SIF maps obtained with the HyPlant airborne sensor over an agricultural research site in Klein Altendorf, Germany. Lower SIF is evident in forests (left in lower panel) and higher SIF in dense agricultural fields (middle and right in lower panel). Fluorescence emission reveals information on vegetation status which is not visible in the reflectance domain. For example, the two fields denoted as A and B display almost identical reflectance (b), whereas their fluorescence emission is very different (a,b) (from Mohammed et al. [72]).
Agriculture 15 02233 g008
Figure 9. (a) Estimated leaf N maps from airborne hyperspectral data (0.4 m spatial resolution) for the 2021, and (b) estimation leaf N from Sentinel-2 data (10 m spatial resolution) (from Wang et al. [120]).
Figure 9. (a) Estimated leaf N maps from airborne hyperspectral data (0.4 m spatial resolution) for the 2021, and (b) estimation leaf N from Sentinel-2 data (10 m spatial resolution) (from Wang et al. [120]).
Agriculture 15 02233 g009
Figure 10. (a) Location of the study site within the Oak Openings regions in Ohio, USA. (b,c) section of visible image with dull colour linear feature interpreted as drainage tile with a parallel network, (d) the UAV used to acquire image, (e,f) a section of thermal infrared images of the study site with drainage tile (from Becker et al. [136]).
Figure 10. (a) Location of the study site within the Oak Openings regions in Ohio, USA. (b,c) section of visible image with dull colour linear feature interpreted as drainage tile with a parallel network, (d) the UAV used to acquire image, (e,f) a section of thermal infrared images of the study site with drainage tile (from Becker et al. [136]).
Agriculture 15 02233 g010
Figure 11. GF-2 RS data of fused images and their corresponding labels for detection of terrace. The top row shows the GF-2 RS dataset of fused images. The bottom row represents the true labels corresponding to the GF-2 sample set of fused images (from Yu et al. [87]).
Figure 11. GF-2 RS data of fused images and their corresponding labels for detection of terrace. The top row shows the GF-2 RS dataset of fused images. The bottom row represents the true labels corresponding to the GF-2 sample set of fused images (from Yu et al. [87]).
Agriculture 15 02233 g011
Figure 12. Deforestation events (2006–2016) were identified from Landsat time series (1990–2016) by analysing mean and standard deviation of photosynthetic vegetation indices. The example demonstrates how RS enables long-term monitoring of land cover change (from Tarazona et al. [154]).
Figure 12. Deforestation events (2006–2016) were identified from Landsat time series (1990–2016) by analysing mean and standard deviation of photosynthetic vegetation indices. The example demonstrates how RS enables long-term monitoring of land cover change (from Tarazona et al. [154]).
Agriculture 15 02233 g012
Figure 13. (a) Linear threads in farmland, (b) demarcation lines, (c) boundary objects, (d) perimeter boundaries of farmland blocks. Such boundary features allow quantification of field size and shape, which are indicators of A-LUI (from Wang et al. [88]).
Figure 13. (a) Linear threads in farmland, (b) demarcation lines, (c) boundary objects, (d) perimeter boundaries of farmland blocks. Such boundary features allow quantification of field size and shape, which are indicators of A-LUI (from Wang et al. [88]).
Agriculture 15 02233 g013
Figure 14. State border between the former FRG and GDR (different farming practices) after the fall of the Wall is clearly visible due to the shape, size, and small-scale nature of the border between Saxony-Anhalt and Lower Saxony north of the Harz Mountains in the 1990s, Germany.
Figure 14. State border between the former FRG and GDR (different farming practices) after the fall of the Wall is clearly visible due to the shape, size, and small-scale nature of the border between Saxony-Anhalt and Lower Saxony north of the Harz Mountains in the 1990s, Germany.
Agriculture 15 02233 g014
Figure 15. Normalised thermographic reflectivity profile across three fields (corn, wheat, and barley) based on RADAR RS data (from Steele-Dunne et al. [167]).
Figure 15. Normalised thermographic reflectivity profile across three fields (corn, wheat, and barley) based on RADAR RS data (from Steele-Dunne et al. [167]).
Agriculture 15 02233 g015
Figure 16. Field photographs to illustrate the surface roughness conditions in different agricultural plots on the Kosi Fan. (a) shows the photograph of a stubble field, (b) harrow field, (c) ploughed field, (d) furrow field, (e) surface undulation profile extracted by processing the photographs captured for the pin-profile using a digital camera in the field (from Singh et al. [171]).
Figure 16. Field photographs to illustrate the surface roughness conditions in different agricultural plots on the Kosi Fan. (a) shows the photograph of a stubble field, (b) harrow field, (c) ploughed field, (d) furrow field, (e) surface undulation profile extracted by processing the photographs captured for the pin-profile using a digital camera in the field (from Singh et al. [171]).
Agriculture 15 02233 g016
Figure 17. (a) Image in the top left shows the location of the Kosi megafan in the Himalayan Foreland, (b) spatial distribution of surface roughness prediction from Sentinel-1, Sentinel-2, and Shuttle RADAR Topographic Mission (SRTM) data. (from Singh et al. [171]).
Figure 17. (a) Image in the top left shows the location of the Kosi megafan in the Himalayan Foreland, (b) spatial distribution of surface roughness prediction from Sentinel-1, Sentinel-2, and Shuttle RADAR Topographic Mission (SRTM) data. (from Singh et al. [171]).
Agriculture 15 02233 g017
Figure 18. Results of the wall-to-wall crop type mapping using the benchmark 10-day interval composite of Landsat and Sentinel-2 time series for Germany (from Griffiths et al. [90]).
Figure 18. Results of the wall-to-wall crop type mapping using the benchmark 10-day interval composite of Landsat and Sentinel-2 time series for Germany (from Griffiths et al. [90]).
Agriculture 15 02233 g018
Figure 19. Types of grassland management intensity at the Podlaskie study sites: (a) extensive, (b) intensive; (c) comparison of spectral curves for intensive and extensive grasslands averaged with the loess algorithm (span = 0.35, confidence interval = 0.95) based on Sentinel-2, (from Bartold et al. [91]).
Figure 19. Types of grassland management intensity at the Podlaskie study sites: (a) extensive, (b) intensive; (c) comparison of spectral curves for intensive and extensive grasslands averaged with the loess algorithm (span = 0.35, confidence interval = 0.95) based on Sentinel-2, (from Bartold et al. [91]).
Agriculture 15 02233 g019
Figure 20. Difference map of Aboveground Biomass (AGB) estimates of 18 August 2017 and 26 August 2017 derived from PlanetScope (PS) optical, Sentinel-1 SAR, and hybrid (optical plus SAR) datasets. Reddish tones indicate AG increase and blue tones indicate AGB decrease. White areas indicate low AGB variation (from Breunig et al. [195]).
Figure 20. Difference map of Aboveground Biomass (AGB) estimates of 18 August 2017 and 26 August 2017 derived from PlanetScope (PS) optical, Sentinel-1 SAR, and hybrid (optical plus SAR) datasets. Reddish tones indicate AG increase and blue tones indicate AGB decrease. White areas indicate low AGB variation (from Breunig et al. [195]).
Agriculture 15 02233 g020
Figure 21. Exemplary application of SugarViT (Vision Transformer based model for disease severity) for disease severity prediction in sugar beet using UAV multispectral data. Each prediction is completely independent of its surrounding predictions. The model shows a highly consistent prediction behaviour (from Günder et al. [92]).
Figure 21. Exemplary application of SugarViT (Vision Transformer based model for disease severity) for disease severity prediction in sugar beet using UAV multispectral data. Each prediction is completely independent of its surrounding predictions. The model shows a highly consistent prediction behaviour (from Günder et al. [92]).
Agriculture 15 02233 g021
Figure 22. Mean SOC content and C:N ratio maps (100 m resolution) predicted by 100 runs of boosted regression trees (BRT) using multiple predictors (Sentinel-1, Sentinel-2, Landsat-8, climate, and topography). Standard deviation maps indicate model uncertainty (from Zhou et al. [93]).
Figure 22. Mean SOC content and C:N ratio maps (100 m resolution) predicted by 100 runs of boosted regression trees (BRT) using multiple predictors (Sentinel-1, Sentinel-2, Landsat-8, climate, and topography). Standard deviation maps indicate model uncertainty (from Zhou et al. [93]).
Agriculture 15 02233 g022
Figure 23. The diagram illustrates the conceptual flow: from management practices (e.g., fertilisation, irrigation, tillage) through trait responses (leaf nitrogen, canopy structure, phenology, root traits, soil organic matter) to RS observables (spectral indices, SIF, SAR, thermal, LiDAR). These feed into the five proposed A-LUI indicator categories (trait, genesis, structural, taxonomic, functional), which can be validated against in situ and administrative data and aligned with policy frameworks (e.g., SDGs, CAP indicators, IPCC inventories).
Figure 23. The diagram illustrates the conceptual flow: from management practices (e.g., fertilisation, irrigation, tillage) through trait responses (leaf nitrogen, canopy structure, phenology, root traits, soil organic matter) to RS observables (spectral indices, SIF, SAR, thermal, LiDAR). These feed into the five proposed A-LUI indicator categories (trait, genesis, structural, taxonomic, functional), which can be validated against in situ and administrative data and aligned with policy frameworks (e.g., SDGs, CAP indicators, IPCC inventories).
Agriculture 15 02233 g023
Figure 24. (a) Semantic data integration for assessing A-LUI. Integration of diverse data sources (satellite data, soil data, agricultural statistics, research databases, socioeconomic information, model data) linked open data (LOD), AGROVOC (AGRO = Agriculture, VOC = Vocabulary of the FAO). (b) A schematic figure visualise this concept for practitioners: RS data (satellite imagery → crop type map), Farm records (fertiliser log, irrigation schedule), Ontology layer: Shared concepts like “crop type,” “management practice,” “season” represented as semantic links, SPARQL query box: Example query: “Select all irrigated maize fields with NDVI > 0.7 in 2023.”, Output: harmonised map or table showing linked information.
Figure 24. (a) Semantic data integration for assessing A-LUI. Integration of diverse data sources (satellite data, soil data, agricultural statistics, research databases, socioeconomic information, model data) linked open data (LOD), AGROVOC (AGRO = Agriculture, VOC = Vocabulary of the FAO). (b) A schematic figure visualise this concept for practitioners: RS data (satellite imagery → crop type map), Farm records (fertiliser log, irrigation schedule), Ontology layer: Shared concepts like “crop type,” “management practice,” “season” represented as semantic links, SPARQL query box: Example query: “Select all irrigated maize fields with NDVI > 0.7 in 2023.”, Output: harmonised map or table showing linked information.
Agriculture 15 02233 g024
Table 1. Cross-walk between established A-LUI frameworks and the proposed A-LUI indicator taxonomy, highlighting overlaps and novel contributions.
Table 1. Cross-walk between established A-LUI frameworks and the proposed A-LUI indicator taxonomy, highlighting overlaps and novel contributions.
Established LUI Framework Typical Dimensions/IndicatorsCorresponding A-LUI Indicator CategoriesConceptual OverlapNovel Contributions in This Study
Inputs (fertiliser, irrigation, energy, labour, pesticides)Input intensity, chemical/energy flowsTrait; FunctionalInputs affect plant/soil traits measurable by RS (leaf N, chlorophyll, soil moisture)Operationalisation of inputs through RS proxies (e.g., irrigation from Sentinel-1, N status from hyperspectral)
Outputs (yield, production, harvested biomass)Productivity, output per hectareFunctionalYield proxies and biomass reflect outputsExplicit RS yield models, link to SDG indicators, inclusion of uncertainty quantification
System-level impacts (biodiversity, soil quality, GHG emissions)Ecosystem services, species diversity, carbon balanceStructural; Taxonomic; FunctionalImpacts partly addressed via land cover, diversity, ecological functionsNew explicit categories: Structural (field geometry, fragmentation) and Taxonomic (crop diversity mapping via RS)
Land use/cover change intensityExpansion, abandonment, conversion ratesGenesisSometimes treated as part of ‘impacts’New focus on temporal dynamics and trajectories (e.g., crop rotations, double cropping, abandonment)
Efficiency measuresOutput per input (yield per fertiliser, water use efficiency)Trait; FunctionalImplied in productivity frameworksPotential to derive efficiency metrics from RS (e.g., biomass per water unit via evapotranspiration modelling)
Socioeconomic drivers (markets, subsidies, governance)Institutional and policy-related intensity factorsNot directly coveredSocioeconomic context linked indirectlyRS–policy linkages highlighted; positioned as future integration pathway
Cross-scale integration (local–regional–global)Aggregated indicators for monitoringAll categories + indicator appendixOften missing in prior reviewsIndicator appendix as reference tool; bridging table (challenges → solutions) for operational uptake
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

Lausch, A.; Bumberger, J.; Jung, A.; Pause, M.; Selsam, P.; Zhou, T.; Herzog, F. Monitoring Agricultural Land Use Intensity with Remote Sensing and Traits. Agriculture 2025, 15, 2233. https://doi.org/10.3390/agriculture15212233

AMA Style

Lausch A, Bumberger J, Jung A, Pause M, Selsam P, Zhou T, Herzog F. Monitoring Agricultural Land Use Intensity with Remote Sensing and Traits. Agriculture. 2025; 15(21):2233. https://doi.org/10.3390/agriculture15212233

Chicago/Turabian Style

Lausch, Angela, Jan Bumberger, András Jung, Marion Pause, Peter Selsam, Tao Zhou, and Felix Herzog. 2025. "Monitoring Agricultural Land Use Intensity with Remote Sensing and Traits" Agriculture 15, no. 21: 2233. https://doi.org/10.3390/agriculture15212233

APA Style

Lausch, A., Bumberger, J., Jung, A., Pause, M., Selsam, P., Zhou, T., & Herzog, F. (2025). Monitoring Agricultural Land Use Intensity with Remote Sensing and Traits. Agriculture, 15(21), 2233. https://doi.org/10.3390/agriculture15212233

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