Next Article in Journal
The Latest Desertification Process and Its Driving Force in Alxa League from 2000 to 2020
Next Article in Special Issue
Utility of Leaf Area Index for Monitoring Phenology of Russian Forests
Previous Article in Journal
An Overview of Coastline Extraction from Remote Sensing Data
Previous Article in Special Issue
Accuracy Assessment of Atmospheric Correction of KMSS-2 Meteor-M #2.2 Data over Northern Eurasia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advancing Skyborne Technologies and High-Resolution Satellites for Pasture Monitoring and Improved Management: A Review

by
Michael Gbenga Ogungbuyi
1,*,
Caroline Mohammed
1,
Iffat Ara
1,
Andrew M. Fischer
2 and
Matthew Tom Harrison
1
1
Tasmanian Institute of Agriculture, University of Tasmania, Launceston, TAS 7248, Australia
2
Institute for Marine and Antarctic Studies, University of Tasmania, Launceston, TAS 7248, Australia
*
Author to whom correspondence should be addressed.
Remote Sens. 2023, 15(19), 4866; https://doi.org/10.3390/rs15194866
Submission received: 31 July 2023 / Revised: 22 September 2023 / Accepted: 6 October 2023 / Published: 8 October 2023

Abstract

:
The timely and accurate quantification of grassland biomass is a prerequisite for sustainable grazing management. With advances in artificial intelligence, the launch of new satellites, and perceived efficiency gains in the time and cost of the quantification of remote methods, there has been growing interest in using satellite imagery and machine learning to quantify pastures at the field scale. Here, we systematically reviewed 214 journal articles published between 1991 to 2021 to determine how vegetation indices derived from satellite imagery impacted the type and quantification of pasture indicators. We reveal that previous studies have been limited by highly spatiotemporal satellite imagery and prognostic analytics. While the number of studies on pasture classification, degradation, productivity, and management has increased exponentially over the last five years, the majority of vegetation parameters have been derived from satellite imagery using simple linear regression approaches, which, as a corollary, often result in site-specific parameterization that become spurious when extrapolated to new sites or production systems. Few studies have successfully invoked machine learning as retrievals to understand the relationship between image patterns and accurately quantify the biophysical variables, although many studies have purported to do so. Satellite imagery has contributed to the ability to quantify pasture indicators but has faced the barrier of monitoring at the paddock/field scale (20 hectares or less) due to (1) low sensor (coarse pixel) resolution, (2) infrequent satellite passes, with visibility in many locations often constrained by cloud cover, and (3) the prohibitive cost of accessing fine-resolution imagery. These issues are perhaps a reflection of historical efforts, which have been directed at the continental or global scales, rather than at the field level. Indeed, we found less than 20 studies that quantified pasture biomass at pixel resolutions of less than 50 hectares. As such, the use of remote sensing technologies by agricultural practitioners has been relatively low compared with the adoption of physical agronomic interventions (such as ‘no-till’ practices). We contend that (1) considerable opportunity for advancement may lie in fusing optical and radar imagery or hybrid imagery through the combination of optical sensors, (2) there is a greater accessibility of satellite imagery for research, teaching, and education, and (3) developers who understand the value proposition of satellite imagery to end users will collectively fast track the advancement and uptake of remote sensing applications in agriculture.

Graphical Abstract

1. Introduction

Pasture ecosystems transverse more than 40% of earth land surfaces [1], supporting a broad range of biodiversity, conservation, and environmental sustainability [2,3] while making substantial contributions to global carbon removal [4,5]. Globally, the livestock industry directly supports the livelihoods of over 1 billion households, particularly in developing countries, where the pastoral system underpins food security [6,7], often because grasslands are the least expensive form of feed and one of the few ways that extensive land areas can be used for agri-food production [5].
Herein, we define “pastures” as uncultivated grasslands or rangelands (shrublands, prairies, woodlands, meadows, steppe, and savannas (Figure 1)) subjected to seasonal native and domesticated livestock grazing. Rangelands are not subject to intensive management except for seasonal grazing [8,9], often due to their low productivity as determined by seasonal weather [10]. Other pasture types (e.g., sown or exotic species) may originate through the human cultivation of cleared land or conversion from natural grassland. Natural grassland converted to pasture may exclude farm inputs [9], while intensively managed pastures often are subject to management interventions (mowing, synthetic fertiliser, irrigation, pasture species renovation, fencing, etc) to enhance productivity to provide financial income [11,12,13]. Grasslands may also serve to produce livestock supplementary feeds, such as hay, silage, and grain [9,14]. Often, pastures are defined based on their global and regional relevance (Figure 1).
Many direct and indirect factors can result in the degradation of pasture ecosystems. These may include overgrazing, the inherent soil structure, adverse climate conditions, competing land-use activities or incursions by noxious weeds and/or feral flora and fauna [16,17,18,19,20,21,22,23,24]. Decisions to optimise management practices [25] require the support of the efficient and accurate monitoring of production (i.e., quantity), quality, species composition, and availability [4]. Global climate change, including the elevation of temperature and CO2, will affect pastures, altering the species competition dynamics due to changes in the optimal growth rate [5,26]. The mechanism by which plant functional types (e.g., C3, C4) adapt to environmental stresses or ecological disturbances has been classified into two main features they possess, i.e., structural and functional characteristics [9]. Pasture composition, functioning, and structure can be vulnerable to the harmful effects of climate change and anthropogenic activities (i.e., overgrazing), and when a critical threshold [27,28,29] is reached, may hit degradation tipping points.
Avenues for monitoring pasture sustainability indicators fall into three broad categories: field techniques, laboratory/greenhouse/allometric analyses, and proximal sensors.
Field techniques to quantify pasture biomass may include (1) visual methods (i.e., field walking and boot height) to estimate grass height, (2) using scientific equipment such as a Rising Plate Meter (RPM) or C-Dax system (a tow-behind device, see www.c-dax.com, accessed on 6 April 2023) to measure the height and density of pastures [3,30], and (3) destructive harvests through quadrat sampling [26]. While such methods provide data on the ground, previous studies have suggested that such methods cannot adequately account for intra-paddock or intra-seasonal variation [31,32,33].
Laboratory, greenhouse and/or allometric analyses facilitate the direct assessment of morphological parameters such as the leaf area index (LAI) and aboveground biomass (AGB) [34,35,36,37]. Forage productivity and quality through the essential structural contents of feed values (e.g., crude protein, green and dry matter, ash, neutral detergent fibre, etc.) can be easily estimated by statistically upscaling [14,38,39,40,41,42,43]. Although such approaches are accurate and suitable for measuring pasture quality, they are destructive [43,44] and cannot be scaled to larger areas.
Proximal sensing is carried out by equipment such as field spectroradiometers (FieldSpec), ultrasonic sensors, and sonars, with the intent of estimating morphological parameters, e.g., AGB, LAI, sward height, soil moisture, etc., [32,41,45,46,47] from multispectral wavelengths (i.e., red, blue, green, and near-infrared) [39,48] or hyperspectral reflectance bands (i.e., 10–20 nm). However, the influence of soil and ground reflectance (albedo) can interfere with spectral distinction [41] and introduce errors due to instrumental instability [47]. In addition, intra-paddock pasture composition variability due to grazing management (i.e., stocking rates) may not be adequately quantified with this approach [32,49]. In summary, pasture monitoring using direct or proximal approaches can effectively calibrate ground-based pastures, e.g., biomass estimation and retrievals of biophysical parameters for validation purposes. However, such methods can be time-consuming and/or labour-intensive and are often unsuitable for application over a large area [50].
Like other techniques, remote sensing is susceptible to problems associated with technology limitations (i.e., errors from cloud cover, noise, atmospheric and geometric correction, and radiometric resolution) [51]. Global studies have highlighted the possibilities of using remote sensing to monitor pasture cover and biomass while reducing error rates [52,53,54]. Most previous work has, however, focused on global, regional, or sub-regional monitoring [55,56,57]; much less attention has been paid to pasture remote sensing at the farm or paddock level, likely because most remote sensing applications have been at larger scales [33,58]. The emergence of newer satellite constellations (e.g., Sentinel and PlanetScope [59]) could be expected to premise innovation in pasture monitoring at a level that can be carried out at more frequent intervals (1–5 days).
The primary objectives of this paper are to (i) review existing satellite and UAS applications in pasture monitoring; (ii) investigate existing approaches for the management of pasture traits, productivity, botanical composition, and pasture degradation; and (iii) explore barriers to the adoption of satellite driven technology by end users.

2. Methods

We included peer-reviewed papers published from January 1991 to February 2021 and excluded conference proceedings and documents written in languages other than English. We first searched the Scopus database interrogated with the term “remote sensing,” which returned 254,392 documents (see Appendix A for the flowchart). We then searched terminology used to describe vegetation under management, i.e., “(pasture* OR grassland* OR rangeland*) management”, which returned 31,938 documents. We then combined queries (#1 AND #2), resulting in 1582 documents. We introduced grazing (i.e., graz*) to streamline this third list to select only articles that describe vegetation under a grazing regime, obtaining 633 papers. After eliminating conference papers and proceedings, we obtained 262 articles. We previewed the articles using the search strings described in Table 1, and by reading the abstracts to eliminate unrelated papers, reduced the number of papers to 214. Articles that made the final round (i.e., 214) were grouped into original research and review papers. From these documents, the following information was extracted:
  • The geographical location/site of a study.
  • The type of sensor used (i.e., optical, multispectral, hyperspectral, SAR).
  • Whether a single sensor was used or a combination of sensors together (i.e., fusion).
  • The scale with which pasture was monitored (i.e., ≤5 ha, ≥10 ha, ≤50 ha and ≥100 ha);
  • The approach for retrieving vegetation parameters for estimating pasture indicators and how this was validated. Information on the adoption of remote sensing approaches by end users.
  • Whether environmental (climate and anthropogenic) variables and machine learning were considered.
We exported results using the Research Information System (RIS) tag format by creating a custom CSV file to format and analyse data. We defined “UAS” as unmanned aerial systems remotely controlled or programmed to fly autonomously with onboard high-resolution sensor(s). In contrast, satellites orbit the earth with onboard sensor(s) often lower than UAS in spatial resolution.

3. Results

The major global grasslands where authors use remote sensing technology to study pasture conditions are shown in Figure 1. For example, the Banni grassland in India [60], the temperate [61,62] and Mongolian steppe [28,63,64] in eastern Asia, the prairies of the central United States [45,49,65,66,67,68], the meadow of North Tibet [69,70,71] and alpine in China [72,73], the tropical grassland of Brazil [74,75], the savanna of Africa [7,76,77], the Greek island of Samothraki in Europe [20,78] and the southern tablelands of Australia [79,80,81] are places of interest where human-induced activities have impacted pasture ecosystems.

3.1. Spatial and Temporal Dimensions of Reviewed Papers

The review process revealed 199 articles from 46 countries (Figure 2), with the United States having the most studies (n = 38; being primarily related to management, production, species composition, and degradation) followed closely by China (n = 36). Asia had a higher proportion of relevant papers (i.e., 24.6%) than other continents due to the publications from China; Australia (n = 16), South Africa (n = 12), Argentina (n = 8), Canada (n = 7), New Zealand (n = 5), Germany (n = 5), Sweden (n = 4), Uruguay (n = 4), Brazil (n = 4), and France (n = 4) had relevant publications.
The number of remote sensing studies was low in 1991, and has increased exponentially. Studies in earlier years focused on management and production, while the proportion of studies on pasture degradation, productivity, and management increased significantly in later years. A lack of publicly available satellites and UAS may have caused fewer studies in the early 1990s. Figure 3 shows the temporal pattern of studies when grouped by management traits, pasture production, species classification and degradation.

3.2. Remotely Sensed Environmental Parameters Applied to Pasture Monitoring

Two primary drivers, namely, anthropogenic and climate/weather, influence pasture mapping (production, composition, and degradation) globally (Figure 4a). Anthropogenic factors (referenced by 75 studies) are further categorised to include stocking rate, stocking density, grazing intensity, grazing system, livestock weights, mowing, soil and fire management, land use, pasture treatment (i.e., fertilizer, herbicide, nitrogen, etc.) and irrigation [2,7,49,65,66,80,81,82,83,84,85,86,87,88,89,90]. These variables are quantified through field measurements (and/or ancillary data) and correlated against remote sensing data [50,70].
Assessing climate or weather’s influence on pastures (i.e., mapping pasture phenology) involves (see Figure 4a) correlating climate and weather data with vegetation parameters (i.e., aboveground biomass, ground cover, and canopy cover) [6,35,49,64,66,80,91,92,93,94,95]. Historical time series satellites (Landsat, AVHRR, MODIS) (medium to low sensors) [2,17,19,70,73,84,91,92,94,96,97,98,99,100,101,102,103] are used mainly. Other methods include simulation and modelling approaches to distinguish between human activities and climate [19,84,104,105,106,107]. Examples of such modelling approaches include annual unharnessed net primary productivity (NPP) from livestock grazing intensity using a defoliation formulation model (DFM) [107], terrestrial ecosystem model (TEM) (potential) and Carnegie Ames Stanford Approach (CASA) [19].
A total of 54 studies (Figure 4a) used climate and weather data to predict climate’s effects on pastures; authors have used climatic data (over at least ten years) and historical time series satellites. Studies focusing on less than ten years were those using a monthly seasonal approach (154 studies) to monitor pastures; we categorised these studies here as “medium to low resolution” (Figure 4b) [49,64,66,80,108,109,110,111,112].
In general, the studies reviewed were conducted using low-spatial-resolution satellite instruments. A few studies used high-spatial-resolution datasets (i.e., Sentinel-2 (four studies), PlanetScope (1 study), QuickBird (1 study)) to understand climate and weather effects on pasture composition [108,113], pasture biomass [3,74,114] and pasture quality [50]. One study invoked a very high temporal time-lapse camera to study the phenology of pasture species at the paddock scale compared with that from the landscape scale using MODIS and Landsat instruments [80]. Only 18 studies considered the adaptive management of pastures with remote sensing strategies [17,21,28,34,67,73,80,94,114,115,116,117,118,119,120,121,122,123] (Figure 4b).
Regarding climate data, authors have primarily focused on temperature and rainfall; few have used only temperature [8,90] or rainfall data [91,96,100,106,124,125,126,127] to correlate with remote sensing data. Many studies have attempted to establish a correlation between rainfall and the growing season (mostly early and mid-growing season) [35,64,66,70,98,103,128]. Some studies have showed that temperature correlated positively to pasture growth according to the climatic zone (e.g., temperature contributed to growth rates in the desert steppe of Inner Mongolia [64], suggesting the steppe possesses more resilience in this region). Other studies have aimed to elucidate the effects of the climate on soil carbon stocks [113,129,130] and/or the soil water content [63,81]. A significant portion of studies used weather and climate data from meteorological stations (Figure 4b); very few were derived (e.g., groundwater content) from sensors as a proxy and compared with ground measurements (i.e., wet and dry pasture biomass) [76].

3.3. Remote Sensing Technologies Used for Pasture Monitoring

3.3.1. Description of Remote Sensing Technologies Used

A total of 18 sensors from satellites and UAS were reviewed (Figure 5a). Our results show that contemporary scientific capabilities in monitoring pasture dynamics in the past decade are gaining momentum, from satellite and UAS sensors to aerial stereoscopic imagery. Figure 5b categorised sensors using combinations of two optical instruments (OO) or optical and radar instruments (OR) [45,58,87,131]. A greater number of studies used satellites than UAS sensors (Figure 6), and fewer studies used SAR (Synthetic Aperture Radar). The main objective for combining sensors [45,118,132] is to address cloud contamination, especially in places where cloud poses significant challenges (i.e., tropical rainforest, mountain regions, polar and monsoon areas) [32,74,99,133,134,135], with multi-temporal sensors approach [99,136,137] or by using SAR imagery [7,132,138,139]. Other objectives include comparing model performances between sensors [7,34,45,132,140,141] and when greater detail is needed for field measurements and species discrimination [114,142,143]. Fused data of 30 cm resolution from UAS and PlanetScope imagery achieved a higher correlation of 87% compared with ground measurement for estimating pastures at the field level (10 ha) than those obtained from Planet (65%) data [114].
Moderate Resolution Imaging Spectroradiometer (MODIS) (Terra and Aqua) and Landsat instruments were the most used for studies (Figure 6; Table 2 and Table 3). The use of MODIS is enabled by daily revisit, 16-day composite, three spatial and global resolutions (250 m, 500 m, and 1 km), and 36 multispectral bands for wider applications. Long-term data continuity, moderate-resolution imaging, multispectral capabilities, and open data policy (Table 2) are among the factors that have aided Landsat’s utility. In general, the use of sensors by practitioners tends to follow their release, accessibility, and applications (Table 2 and Table 3).
Figure 7 shows the areas of coverage and scale of focus for the current monitoring of pasture with satellite sensors. MODIS and Landsat sensors are used primarily to support the regional and global monitoring of pastures at scales ≥100 ha [45,67,103,115,126,144,145]. Figure 8 is an example of hyper-spatial paddock monitoring. Time-series analysis showed that eight studies utilised daily remote sensing data (Figure 9) [58,114,146,147], five focused on weekly [3,126,148,149,150], while others considered monthly [8,151] and yearly data [92].

3.3.2. Definition of Pasture Feature Terminologies as Used in the Review

Pasture management traits are the desired indicators for conserving, restoring, and maintaining grassland conditions [12] (Table 1). Pasture production refers to the quantitative parameters that express pasture’s dry matter content (kg DM/ha), height, and growth stages (see Table 1). Pasture degradation refers to decreased sward productivity (carrying capacity) due to anthropogenic and environmental activities on pasture ecosystems. Botanical/species composition refers to ground cover types expressed as canopy architecture.
Table 2. The descriptive characteristics of the satellite instruments used in the review. (*) European Remote Sensing Satellite-1 (ERS-1).
Table 2. The descriptive characteristics of the satellite instruments used in the review. (*) European Remote Sensing Satellite-1 (ERS-1).
Satellite InstrumentVersionAltitude (km)Launch YearRevisit (Day)Spatial Resolution (m)Spectral Bands Red Edge InclusionMain FocusReferences
MODIS 7051999/20021250/5000/100036(2, 5, 29)NilRegional and global daily application. (MOD 17 model)[6,71,90,123,124,126,155,156]
Landsat5 to 870519721615/30/10011NilRegional and global seasonal coverage.[78,144,152,154,157,158,159,160,161,162,163]
Sentinel-2 78620155–1010/20/6013–22YesFlexible resolution (revisit spatial) and red-edge inclusion. [3,32,50,76,148,164]
SPOT2 to 76941990–20141 to 32/85NilVegetation instrument and stereo capability.[31,149,150,152,165]
AVHRR18331998–2018111005NilDaily global application archive.[91,166,167]
Sentinel-1 69320146 to 12Depend on acquisition mode.3 (0.12–0.50 nm) Provide global free C-band SAR data. Unique acquisition mode.[132,168,169]
RapidEye1 and 26301998–200816.55YesVery high daily global imagery.[140]
QuickBird 48220011–3.50.61/2.44NilVery high daily global imagery.[78]
Worldview1 to 46172007–2016<10.31/3029YesMore bands for global distinctive imaging.[170,171]
IKONOS 68119991–3, 141/44 Very high imaging and stereo capability.[157,172]
Hyperion 70520001630hyperspectral Narrow bands[173,174]
ERS-1 * 7821991 10/30 C-band SAR data and polarization. [153,175]
Formosat2 888200412/85Nil [176]
PlaneScope 461 135YesDaily fine global imaging.[173,177,178]
HySpiri 2018560hyperspectral Narrow bands for characterization.[177,179]
ALOS1 and 26282006–201414, 462.5/10L-band SAR data and 4 optical bands.NilOptical and SAR imaging possibilities.[7,180]
Venus 7202017/200523/5.312Yes High spatial and spectral application.[177]
Table 3. Characteristics of major UAS used in the review. The generic name is the instrument’s name, while the traditional name is the company’s brand name.
Table 3. Characteristics of major UAS used in the review. The generic name is the instrument’s name, while the traditional name is the company’s brand name.
Generic NameTraditional NameSensor Spatial ResolutionFocus Reference
UASPhantomMultispectral<1 mPasture biomass[58]
UASUAS LiDARLiDAR sensor40 mBiomass estimation and species classification[181]
UASPhantom and SequoiaMultispectral1.5 cm and 3.7 cmClassifying fractional cover [116]
UASHexa Copter SystemMultispectral10 cm Pasture biomass productivity[182]
UAS + PlanetScope (fused)MicaSenseMultispectral30 cmAboveground net production[114]
UASMicro MCAMultispectral30 mPasture quality[183]
UASAisaFENIXHyperspectral (VIs-SWIR1 mPasture nutrient[142]
UASHySpexHyperspectralDepend on altitudePasture species (classification)[131]
Airborne laser scanningRiegl LMS-Q680 sensorLiDAR; reflectance, echo width NDSMDepend on altitudePasture mapping[184]
UASHymapHyperspectral5 mPasture species (classification)[52]
Aircraft mounted + calibrate Landsat 5 (TMS)Very-large-scale aerial (VLSA)Multispectral of Landsat1 mm (VLSA), Landsat 30 m Pasture cover from Landsat calibration[143]

3.4. Approaches for Pasture Quantification

3.4.1. Pasture Production

Studies have used pasture heights [7,116,132,146,150,162,184,185,186] LAI [32,45,49,66,135,150,164,187], fractional cover (fCOVER) [188], above-ground net primary production (ANPP) (unit mass per unit area per unit time) [8,34,96,111,189,190], fraction of photosynthetic active radiation (fPAR) [191] as quantitative parameters to express pasture production. ANPP and fPAR are mainly derived from Landsat and MODIS time-series products to quantify the managed ecosystem productivity, making them less applicable compared to LAI and pasture heights. Studies have compared pasture biomass with LAI [32,45,49,119,188,191] and height [132,150,162]; hence, the goal is to use LAI and pasture heights as proxies in estimating pasture biomass.
Vegetation indices (VI) are the most adopted retrieval scheme with empirical approaches (Table 4) to estimate the pasture height or biophysical parameters (LAI, fCOVER, ANPP, and fPAR) (Table 4) and relate them with pasture biomass, where the normalized difference vegetation index (NDVI) [45,69,133,166,192] accounts for 83% of this method. Gillan et al. [193] correlated the canopy height (R2 = 78%) with the ground biomass to infer pasture biomass utilisation at the field scale. Next to NDVI are the enhanced vegetation index (EVI) and soil-adjusted vegetation index (EVI), used with other indices to provide complementary information about their sensitivity to sparse and dense vegetation [74,144,150,181,183]. Index-based retrievals significantly rely on the visible and NIR bands and SWIR for those that require soil water content and dry biomass estimation [149]. A mathematical transformation function (e.g., power and logarithm) is used to normalise data (i.e., expand or compress the index value) to minimise the saturation effects of vegetation indices [64,194].
A physical-based approach using the radiative transfer model (RTM) PROSAIL has been used to retrieve LAI as a vegetation canopy when combined with multispectral satellites [32,135,164,187] and often parametrized and optimised with ML algorithms [45,135,164]. Some studies [32,135] have used proximal hyperspectral data resampled to a satellite multispectral dataset (i.e., Sentinel-2) to constrain the assumption of the homogeneous canopy of the RTM and enhance the accuracy of the model. For example, [32] established a correlation coefficient of 50% between modelled (PROSAIL + resampled data) and in situ biomass. In most cases, LAI correlated better with referenced biomass data than NDVI or other indices [32,188,191]. Furthermore, both the perpendicular vegetation index (PVI) and SAVI derived from Landsat have been found to correlate with the referenced LAI (R = 50%) more than NDVI and other indices [66], confirming the site specificity of LAI-based modelling. Measured LAI is converted to biomass through a linear relationship and compared with the satellite spectral index [32,135].
Other physical-based approaches like the light use efficiency (LUE) (the amount of absorbed photosynthetically active radiation (APAR) that is converted into biomass and expressed as the net primary production (NPP)) model has been used to estimate available pasture biomass (Equation (1) shows the linear association between NPP and LUE).
N P P = A P A R × L U E
where NPP is the available biomass, expressed as the net primary production (NPP), and APAR is the absorbed photosynthetically active radiation that plants can utilise to produce biomass. MODIS-ANPP products are converted to biomass using biomass-to-carbon conversion [8,90,167]. Liu et al. [35] fused UAS and PlanetScope, while [190] used MODIS to model ANPP as a function of APAR derived from NDVI and light use efficiency (LUE). Similarly, the LUE model has been used to estimate other variables like the green canopy cover, vegetation density, fractional vegetation cover, and fAPAR [8].
Table 5 shows how ML models were integrated with remote sensing for retrievals. Random forest (RF) is the most widely applied ML algorithm due to its capability for regression and classification problems. Generally, ML is used to parametrize index-based retrievals [45,58,74,132,192] more than to retrieve spectral features [3,7]. The ML-based modelling of biomass is more accurate than NDVI [3,74]. Chen et al. [3] established a non-linear relationship between NDVI and in situ biomass. Raab et al. [132] used RF/Support vector machine (SVM)/Multi linear regression (MLR) to parametrize 77 vegetation indices derived from Senitnel-2 to estimate total standing dry matter (TSDM) at an accuracy of R2 = 45%.
In cloudy situations, authors have used three-dimensional photogrammetric point cloud modelling to assess grassland heights (i.e., between 1–20 cm) with the visible camera from UAS [116,184,185,193]. Gillan et al. [193] found a correlation of R2 = 78% between aerial imagery and in situ estimates and an average utilisation of 20% from imagery against the 18% of ground-based imagery at a scale of 150 ha. Furthermore, LIDAR has also been used to quantify biomass from different plant species using a 3-D approach at a field level (6.7 ha) with 77% accuracy [181].
Studies have used SAR imagery to complement optical applications using the backscatter signal of microwaves to estimate pasture biomass [7,45,132,153,195] and height [7]. The accuracy of models increases when SAR data are combined with optical imagery compared to a single application. For example, the combination of Sentinel-2, Sentinel-1, and Landsat improved the estimation of biomass by 30% compared to each of the sensors alone [45], and RMSE was significantly lowered when [195] fused Worldview-3 and Sentinel-1. Similarly, the L-band of ALOS (PAR-SAR-2), which is capable of penetrating the canopy structure, was combined with the C-band of Sentinel-1 and 13 spectral bands of Sentinel-2 to map and discriminate pasture heights (short, medium, and tall) from soil inundated with the vegetation canopy using RF [7]. The model’s overall accuracy improved (R2 = 86%) by integrating the three sensors rather than individual contributions.
Furthermore, the wavelet principal component analysis (WPCA) used for SAR and optical image data extraction (based on relevant features) was used to improve the fusion between ERS and Worldview, leading to the higher accuracy of the model [195]. They [195] reported a strong correlation (R2 = 79%) between the backscatter coefficient of Sentinel-1 and ground biomass from rangeland rehabilitated from mining activities. Similarly, [175] established a strong correlation (59–84%) between ERS backscatter coefficients and targeted wet grassland biomass by applying a linear inversion algorithm to the image data. In contrast, [132] found no significant contribution of Sentinel-1 data when combined with Sentinel-2 in estimating biomass and pasture height in Germany.
Table 4. Summary of the main vegetation indices and associated regression algorithms commonly used in the studies.
Table 4. Summary of the main vegetation indices and associated regression algorithms commonly used in the studies.
Vegetation IndicesModelStudies FocusSensorReference
Ratio vegetation index (RVI), enhanced vegetation index (EVI), NDVILogarithmic regressionAboveground biomassMODIS[64,189,194,196]
EVI, LAI,Linear regression modelAboveground biomassWorldview, Sentinel-1, Sentinel-2, Landsat[45,196]
Vegetation indicesSparse partial least-square regressionAboveground biomassSentinel-2, HySpiri,[179,197]
Pasture qualityUAS,[120,142]
AVHRR[91]
NDVIPower regressionPasture biomass, forage dry biomassMODIS,[91]
LAI derived from satelliteRadiative transfer modelPasture biomass prediction at the paddock levelSentinel-2[32]
NDVI derived from fused satellite sensorsLinear regression modelAboveground net primary production (i.e., carbon stock) (ANPP) estimated from Absorbed photosynthetically active radiation (APAR) at paddock levelFusion of Landsat/MODIS[34]
NDVI derived from fused satellite sensors + UASLinear regression + Light use efficiency modelAboveground net primary production (ANPP) estimated from Absorbed photosynthetically active radiation (APAR)Fusion of UAS/PlanetScope[114]
To compare NDVI and FVC derived from UVA (multispectral image)Exponential function, linear function, logarithmic function, polynomial function and power functionEstimate carbon yield canopy cover for individual plant and acrossMultispectral camera (i.e., SpecTerra)[130]
NDVI + cellulose absorption index derived from satellitesLinear unmixing approach and multiple linear regressionFVC, non-photosynthetic vegetation cover and bare soilHyperion and MODIS[91,141]
Table 5. The summary of the major machine learning and other retrieval methods used to estimate pasture biomass in the studies.
Table 5. The summary of the major machine learning and other retrieval methods used to estimate pasture biomass in the studies.
Methods/Biophysical/Spectral ParametersMachine Learning/ModelApproachSensorGround ApproachAchievementReference
LAI derived from satelliteRadiative transfer model + artificial neural network as retrievalPasture biomass Sentinel-2 [135]
NDVI derived from UASStatistical (GAM) + Machine Learning (RF)Pasture biomass prediction at the paddock levelMultispectral cameraGround calibration with RPM27% (GAM) and 22% (RF)[58]
NDVI and spectral variables derived from satellite imageryArtificial neural networkPasture biomass prediction at the paddock levelSentinel-2Calibration with C-Dax and RPM 51% (ANN) and 39% (NDVI)[3]
LAI + soil leaf canopy (SLC) derived from satelliteRF + Radiative transfer model (RTM)LAI and aboveground biomass (AGB)Sentinel-2Field samplingRMSE of 0.4. [164]
VIs (NDVI, EVI and Land surface water index) derived from satellite SVM, RF and Multiple linear regression (MLR)Estimate LAI and aboveground biomass (AGB)Sentinel-2, Sentinel-1
Landsat
Field sampling (destructive)30% improvement by combining sensors [45]
Surface reflectance data (Landsat 8 + MODIS) compared to NDVI, EVI +SAVIGaussian Process Regression (GPR)Estimation of aboveground biomassLandsat 8 and MODISField sampling (destructive)GPR outperformed the three VIs R2 = 0.64 and RMSE = 48.13 g/m2[198]
Spectral reflectance ANNQuantifying abovegroundLandsat 8Field sampling (destructive) [174]
Fractional of Absorbed Photosynthesis Active Radiation (FAPAR) derived from RSDecision Tree (Machine Learning) Estimation of herbaceous yield in a (savanna ecosystem) Traditional FAPAR + meteorological dataML + FAPAR + climate data performed better than FAPAR model only and/or climate variables.[99]
LAI + NDVI + Fractional vegetation cover (FVC) derived from satelliteK-NNMapping grazing and mowingSPOTField measurement (spectrometer)82%[188]

3.4.2. Botanical Composition

Some studies have pursued the key objective of finding a suitable instrument to discriminate vegetation canopy. Multi and hyper-spectral sensors ≤ 30 m (Table 5 and Table 6) are the most deployed to discriminate vegetation canopies with either image-based [33,52,199,200] or object-based image analysis (OBIA) [62,163,169,172,184]. Classifiers derived from hyperspectral sensors are more accurate than multispectral instruments [177,200], while multisource instruments are more accurate than single sensors [113,169,177]. Sibanda et al. [177] established a higher spectral accuracy of 92% from HySpiri than from Landsat 8 (75%), Sentinel 2 (82%), and Venus (83%). Likewise, the study from [113] concludes that IKONOS, QuickBird, and Worldview sensors with finer spatiotemporal resolutions are more sensitive to discriminating grassland from shrubs and trees than Landsat imagery. Studies that used ML algorithms as classifiers (mainly MLC, RF, SVM, and k-Nearest Neighbour (KNN) algorithms) [62,172,199,200] improved their classification accuracy more than traditional methods. For example, the accuracy reached 98% and the Kappa coefficient ≥ 90% when ancillary data were added using SVM and RF as classifiers [62]. OBIA enables the mapping of vegetation/species classes and integrations of geometric, textural, and spatial data (i.e., ancillary data) in addition to the primary spectral information to improve accuracy [62,200]. Hence, ML with OBIA can capture the environmental and management variables more accurately than pixel-based algorithms.
More studies have used supervised than non-supervised classification to use novel sampling techniques to build spectral signatures from field areas of interest. Studies have used botanal sampling protocols [139], dominant pasture species [138,161], the percentage of pasture species [33], and the height of pasture species [7,133] to build spectral features. The authors used phenological stages and early growing seasons to improve accuracy using single imagery [49,86,131,201,202]. Wakulinśka and Marcinkowska [133] reported a better classification accuracy from a multi-temporal study than a single-date one. Mapping species in the early season reduces canopy complexity and provides insight into their phenology.
The available studies on the combination of optical and radar sensors to discriminate pastures show that the spectral features derived from Sentinel-2 outperformed the backscatter and dual-polarised features of Sentinel-1 [139] when subjected to similar ML models (SVM and RF). However, merging Sentinel-2 and Sentinel-1 produced a higher accuracy (i.e., RF = 93% and SVM = 89%). Like optical sensors, Sentinel-1 data have been used to discriminate between C3, C4, and mixed C3/C4 pastures, using RF to achieve 68% accuracy [138]. The accuracy level increased to 73% on those including textural features (i.e., leaf area, plant height, size, and orientation) derived from Grey Level Co-occurrence Metrics analysis (GLCM). A study used UAS-LiDAR with a 3 cm accuracy level and a maximum 100 m measurement to detect shrub encroachment and classify 2000 habitat types with 75% accuracy, using RF as a classifier [181].
Table 6. Summary of the main classification algorithms for pasture composition used in the studies.
Table 6. Summary of the main classification algorithms for pasture composition used in the studies.
ClassifierMethodsSensorAccuracyReference
SVM + PCAPixel-basedSentinel-280% (overall)[133]
RFPixel-basedSentinel-2, Sentinel-1, ALOS86% (overall)[7]
SVM + RFObject-image basedLandsat
Kernel + SVM Non-linear performed better (0.55 ≤ R2CV ≤ 0.78; 6.68% ≤ nRMSECV ≤ 26.47%)[142]
Decision treeObject-based classificationIKONOS83%[172]
SVM
Decision tree SPOT
SVMlinear regression/classificationLandsat [136]
K-NN SPOTKappa index= 0.82
Maximum Likelihood Classifier (MLC)Object-image basedLandsat, SPOTLandsat = 60.1%, SPOT = 65.5%,[31]
MultivariateHierarchical clusteringLandsat [203]
RFPixel-based classificationSentinel-1A, Sentinel-276%, 62%, 75%[138]
RF, SVM, KNNPixel-based classificationSentinel-1, Sentinel-2KNN 0.89, RF 0.96, SVM 0.96[139]
MultivariateWhere several treatments are needed [204]
Fuzzy/KNNPixel-basedHyMap98% and 64%[52]

3.4.3. Pasture Management Traits

The main goal is to assess pasture quality using remote sensing as proxies to quantify its management traits [38,46,120,132,140,142,158,170,183,205,206,207]. Nitrogen availability [140,170,202], soil water condition [67,182,207], irrigation [168], mowing [188,208], livestock distribution [188], soil nutrients [77,142], and fertilizer treatment [207] are the major management traits that have been examined by authors and expressed as pasture quality indicators. Pasture quality has been linked with the aggregation of livestock (animal units) to areas with a rich concentration of nitrogen as a proxy for the abundance of vegetation greenness in mapping the spatial distribution of grazing animals [140,208]. Some studies have used vegetation indices by selecting bands (red, red-edge, NIR, SWIR) of interest with linear regression models to relate them with management indicators [140,170,176,202,208]. Agricultural inputs, such as the irrigation date and LAI, were retrieved from FORMOSAT-2 using spectral indices and integrated into crop models to support water management for grazed pastures in France [176].
Other biophysical variables like LAI and fCOVER derived from SPOT imagery have been used with the KNN algorithm to map grazing landscapes to support mowing management [188]. LAI shows a higher correlation of 82% with the sampled data compared to fCOVER. Higher-resolution sensors and/or a combination of multiple sensors have improved model accuracy, especially in a complex field for discriminating grasslands treated with fertilizer (i.e., nutrients) than using one sensor. Sibanda et al. [207] reported an accuracy of 81% for Sentinel-2 and 76% for Landsat (OLI), which were resampled from hyperspectral data in discriminating grasslands treated with fertilizer using sparse partial-least-square regression (SPLSR). The hyperspectral data on its own yielded 92% accuracy. Similarly, HyspIRI data are more accurate (R2 = 69%) than Sentinel-2 (R2 = 58%) using wave bands and VIs with SPLSR in predicting burning, mowing, and fertilizer application [179]. However, [140] reported that the accuracy of Sentinel-2 (R2 = 92%) is higher than RapidEye (R2 = 53%) in predicting nitrogen concentration levels from simple ratio (SR) and NDVI with RF, due to the three red-edge bands present in Sentinel-2 compared to one red-edge band in RapdEye.
ML models improve the discrimination of grasslands based on management indicators rather than linear regression. For example, [142] reported that RF achieved the best accuracy (R2 = 78%) in predicting 77% of the macro and micronutrients derived from hyperspectral UAS (spatial resolution ~3.5–11 nm); SVM achieved 86% accuracy for predicting 22% of the nutrients compared to the squares (PLSR) and kernel (PLSR) algorithms. Ancillary data like GPS provide information about livestock distribution and have been found to improve mode accuracy [140,170,208].

3.4.4. Pasture Degradation

The focus is to correlate key anthropogenic activities (i.e., grazing management) that predispose grasslands to decline with remote sensing products as proxies by relating them to biophysical variables (i.e., fPAR, AGB, fCOVER, ANPP). A significant number of studies have used Landsat and MODIS land surface reflectance products rather than finer satellite imagery to express the productivity of grasslands (fPAR) [191,209] (fCOVER) [20,96,137,167,189,190], (AGB, ANPP) [105,109,129,130,141,144,174,186,210,211], and ecosystems beyond the scope of biomass production. NDVI is the most used proxy for estimating biophysical variables. For example, stocking rate data were compared with yearly AVHRR-NDVI and rainfall data to account for overgrazing on rangelands [212]. Soil-based indices are used after NDVI to understand non-vegetation in mapping landscapes [83,144,213]. Haggen et al. [144] used the soil-adjusted total vegetation index (SATVI) from red and SWIR bands to map fCOVER, while [213] estimated the pasture productivity decline from grazing intensity and fire regime in a semi-arid environment using the derived soil tillage index (STI) from MODIS. The study of [213] showed that SWIR calculated from STI (B6 and B7) was more accurate (R2 = 67%) in mapping drier vegetation compared to NDVI and other indices. LAI was found to be more accurate in estimating fPAR than NDVI [191]. Pasture degradation indicators (grazing intensity/pressure) are often correlated with environmental variables (soil and survey data, meteorological, GPS) to understand the drivers, and are also used with (VIs) as predictor variables [2,20,71,96,105,129,130,141]. Studies have adopted mapping the land cover and land use (LULC) [83,121,199] to show the spatial and temporal variability of an area of interest.

4. Adoption of the Remote Sensing Information by End Users

Table 7 illustrates an overview of the current remote sensing application for pasture monitoring and end users’ level of adoption from this review. Studies have shown that farmers, governments, scientists, and spatial consultants are the main stakeholders in the workflow of remote sensing technology (Table 7). For example, the Queensland Government Australia developed an online “FORAGE” (http://www.longpaddock.qld.gov.au/forage/; accessed on 7 February 2021) system to support grazing management and provide site-specific information [214,215]. The customised FORAGE system has provided pasture and climate parameters on land condition and stocking management to 1700 users.
Eastwood et al. [216] acknowledged the low adoption of the remote sensing of pasture monitoring despite increased research and development (R&D) in the past decade. They suggested that there is a need for vendors/researchers to properly understand the “value proposition” of the end users and integrate this into the workflow of the remote sensing technology. An earlier representative study reported by [216] provides an empirical analysis of the current approach to pasture monitoring from interview and survey perspectives. The survey involved 500 New Zealand dairy farmers on the methods used to derive pasture measurement. Fifty-two percent used the visual approach, 45% used a technology-based scheme (RPM, C-Dax), and 3% used neither. Further technology analysis suggested that 32% depend on RPM, 11% use C-Dax, 1% use satellite, and 1% use the contractor. Therefore, although decision support tools are essential, the value of the premium that end users (e.g., farmers) place on pasture monitoring is not entirely sure; hence, the value proposition seems ambiguous to persuade non-users to consider the technology [216].

5. Discussion

This review provides a systematic analysis of published studies on the methods of remote sensing and their usefulness to pasture monitoring in major global grassland ecosystems (Figure 1). All regions and continents of the world are covered (Figure 2). Still, however, less attention has been received from Southeast Asia, the northern part of Latin America (except Mexico), the Middle East (except Iran and Syria), and Africa, with most of the studies coming from South Africa and Ethiopia.

5.1. Trend in the Remote Sensing of Pasture Management Traits, Species Composition, Pasture Production, and Pasture Degradation

Figure 4 shows a deficient proportion of studies from earlier years. Studies have centred on management, and fewer on production, while coverage on botanical composition and degradation was not in the spotlight. There was an increase with time in all topics, especially with botanical composition [221] and pasture degradation [127]. Studies on species/botanical composition may have gained more prominence recently because higher-resolution satellites and UASs for discriminating vegetation canopy are increasingly available for precision agriculture. Furthermore, issues bordering pasture production and degradation due to anthropogenic and climate activities have become prominent in the scientific literature.

5.2. Assessing the Current Remote Sensing of Pasture Monitoring

Despite the widescale coverage of studies and the current utility of satellite sensors (Figure 5a), a high proportion of this effort focuses on regional, continental, and global scales (Figure 8), with less emphasis on field-based monitoring. The number of studies that have focused on fields within 50 ha is less than 20. Higher-resolution multispectral satellites are not free but are also only constrained to a few bands (mainly visible and NIR), except Worldview and Venus, with 29 and 12 bands, respectively (Table 2). Therefore, the publicly available optical satellites MODIS, Landsat, and the recently launched Sentinel-2 have played a central role following their specifications in monitoring vegetation dynamics. Sentinel-2 arrivals in 2015 were thought to address cloud constraints for optical applications, especially with a 5-day revisit and 10 m resolution fleet. This review shows that apart from the over-emphasis on medium to coarse resolutions (Landsat and MODIS), which limited field-based monitoring, the arrival of Sentinel-2 has not resolved missing data due to cloud contamination, especially in places known for persistent cloud cover (i.e., tropical rainforest, mountain regions, and polar and monsoon areas). Researchers have used different approaches to resolve cloud contamination, such as cloud removing algorithms (e.g., CFmask) to mask cloudy pixels [21,45,67,134,136,222], multi-temporal satellite data [7,34,45,132,140,141], and conducting field campaigns in cloud-free days [164], and the stacking of satellite scenes [21,133]. More specifically, researchers and practitioners have used photogrammetry UAS cameral (visible) equipped with a 3-dimensional point cloud [116,184,193] and LIDAR sensors [181] on demand to capture near-real-time imagery as an alternative to satellite applications.
Additionally, the number of studies that have utilised daily and weekly satellite imagery for analysis is less than 20 (Section 3.4.3), meaning that the current revisit would not support/sustain operational pasture management. Intensively grown pastures are dynamic and require more frequent imagery between 5–7 days to capture sward regrowth depending on environmental conditions. The current satellite fleets with daily revisit are not available for public utility, thus limiting this application for R&D. Leveraging radar capability, the all-weather satellite data (i.e., dual-polarisation, backscatter, with C, L, and X bands), especially the free and open-source Sentinel-1 data (R&D) usage in this review, were relatively low (Figure 6 and Figure 7). In most studies involving pasture production (estimation of biomass and height) and species classification [7,45,139,175,195] except a few [132], the integration of SAR imagery has improved model accuracy more than the performance of either the optical or radar data alone. For example, the fusion of ERS and Worldview imagery using the WPCA method to extract relevant image features in a suitable rangeland environment (i.e., rangeland rehabilitated from mining activities) significantly improved the model performance (R2 = 79%) [195]. However, with cloud containment, spectral information due to surface reflectance (especially the red-edge and NIR bands) from optical data is more accurate for assessing pasture biomass and discriminating species than dual-polarised features and the backscatter coefficient of Sentinel-1 [132,139]. Hence, optical hyper or multispectral sensors in fair weather conditions offer more accuracy in distinguishing vegetation species than SAR data because of their spectral responses along multiple bands. At the same time, the microwave is not sensitive to chlorophyll content but to the structure and volume of vegetation. Therefore, we conclude that the accuracy of SAR modelling depends on the knowledge domain applied to suit the biophysical variables and target environment.
Generally, the combination of instruments significantly provides a platform to constrain remote sensing trade-offs in an integrated way to fix specific errors or limitations associated with sensors and the target environment. The saturation of biomass (in sparse vegetation and/or peak growing season) associated with vegetation indices [45,223], soil background and topography influence on SAR sensitivity [45], homogeneous canopy associated with LAI [32,135] or with 3-D point cloud photogrammetric mapping [116,193], and spatial, temporal, and radiometric resolution drawback can be addressed using appropriate modelling involving hyper-temporal, multispectral, visible and SAR to improve the accuracy of the model. The SAR backscatter is not sensitive to soil background when the vegetation canopy is low. At the same time, the visible and NIR bands of optical instruments enable the absorption of more radiation than soil, resulting in a higher reflectance for denser canopy areas and lower reflectance values for bare soil. Proximal hyperspectral data were resampled to Sentinel-2 surface reflectance and combined with RTM PROSAIL to estimate LAI, thereby confounding the homogeneous assumption related to RTM [32]. Furthermore, using 3-D point cloud photogrammetric to estimate pasture height and biomass from vegetation volume can be confounded with trees, shrubs, and other land use types, making this approach prone to error. Hence, multispectral bands (i.e., NIR) are included to map land cover or mask the non-pasture community [43,82,83,181].

5.3. Assessing the Approaches Used in the Remote Sensing of Pasture Monitoring

The retrieval of biophysical variables has been significantly restricted to empirical methods using vegetation indices, with NDVI being the most used index to understand pasture production, species classification, management indicators, and the degradation of pastures. Likewise, the physically based retrieval schemes to assess the following biophysical variables—LAI, fPAR, and fCOVER, ANPP—using LUE and RTM are driven by sites and parameters that restrict their generalisation and repeatability. Consequently, their comparison and relationship with the field (destructive and non-destructive field samplings) data has reached a milestone in addressing the problems associated with VIs and physically based modelling approaches, while at the same time revealing the potential areas where more research efforts are needed.
Using ML approaches with careful integration of appropriate satellite sensors in addition to environmental data, the modelling of pasture biomass from the selection of VIs has achieved better accuracy and a higher level of prediction (i.e., an increase from 1500 kg DM/ha to above 3000 kg DM/ha) before reaching saturation [45,223], with red-edge, NIR, and SWIR bands as the main contributors. In contrast, red and NIR bands’ usages (NDVI) have continued to trigger a debate on the effect of the saturation, soil background influence, sensitivity to vegetation types, and heterogeneity of canopy-to-model calibration, which have caused researchers to develop more indices and parametrize with ML [45,58,74,132,192] algorithms. However, NDVI has performed poorly compared to spectrally driven ML retrieval (R2 = 60% and R2 = 78%), especially when dealing with total standing dry matter [3,74,132]. Therefore, owing to the emphasis on ML-driven index-based retrieval, which is restricted to a few bands and constrained to sites, more research is needed to retrieve a detailed characterization of vegetation properties based on reflectance values, using spectral features to understand the relationship between pixels. The size of field data used for validations in this review is relatively small (test data reveal ≤ 120) compared to what is needed (see ~1000 [3]) for training to capture image patterns and improve model calibration.
Apart from the problem of generalisation associated with LAI, fPAR, fCOVER, and ANPP, the biophysical variables are mostly computed from medium to coarse sensors (Landsat and MODIS), which makes them widely applicable (i.e., daily, weekly, and monthly composite global data) but more challenging to use for field monitoring and management. Retrieving these variables from higher-spatial-resolution (1–5 m) sensors will significantly facilitate the monitoring of ≤50 ha fields. Indeed, the current hyperspectral UASs (i.e., HyMap, HySpiri, and AisaFENIX with 3.5–11 nm resolution and ~450 bands) have shown great potential in discriminating species and characterizing macro and micronutrients in mixed heterogeneous pastures, which indicates that the availability of these tools (i.e., upscaling to satellites) will significantly enable pasture management at the field scale. Issues relating to costs and logistics would possibly be resolved through disruptive technology and public partnership. The near-future hyperspectral satellites launched by the German (EnMap) and European Space Agency-(ESA) (FLEX) would help determine the cost and logistics, since pilot studies have shown promising results [133,224,225,226].

5.4. Adaptive Pasture Management and Factors That Influence the Monitoring of Pastures with Remote Sensing

Anthropogenic variables and prevailing environmental factors significantly condition pastures. Therefore, adaptive management that focuses on goal-oriented outcomes using suitable remote sensing tools is highly recommended to improve the sustainability and resilience of pastures and grazing systems over time. Remote sensing products must be appropriately quantified regarding what they represent on the ground. Adaptive pasture management integrates anthropogenic, environmental, and climate variables and remote sensing to provide insight into intensively grazed pasture dynamics, thus making pastures and land management sustainable. This review shows that a combination of different remote sensing strategies, e.g., aircraft imagery and Landsat [121], Phenocam camera [80], and UAS and satellite [114], can be used to understand the temporal and spatial variability of pastures to seasonal and climate change in establishing a framework for adaptive management. For example, a very high temporal time-lapse camera has been used to study the phenology of pasture species at the paddock scale compared with that from the landscape scale using MODIS and Landsat instruments [80]. Consequently, remote sensing products have been used with anthropogenic variables (i.e., stocking rate, grazing survey data, fire, GPS (livestock distribution)) and environmental/climate data (rainfall, temperature, soil) as predictors to improve model accuracy significantly [3,176,212]. However, only 18 studies considered the adaptive management of pastures with remote sensing in this review (Figure 4b). Therefore, more research is encouraged to demonstrate how adaptive management principles with remote sensing tools can support the sustainability of pasture management.

5.5. Analysing End Users’ Perception and Adoption of the Remote Sensing Products and Technology

Studies that have considered remote sensing technology’s adoption by end users have been few in number. The published dates (Table 7) of the available studies show more efforts in the earlier years than in the last few years. The small proportion of studies on the adoption of the remote sensing of pasture monitoring show that remote sensing products are not at the level of adoption by end users. This review identifies two setbacks to the adoption of the technology. In theory, satellite launchers expect direct benefits of the products for all stakeholders; in practice, the spatial resolutions of the current satellites benefit regional, national, and global applications, as revealed by this review. Currently, the publicly available Sentinel-2 and Sentinel-1 are being under-utilised (Figure 5a). End users may be unlikely to be persuaded to adopt remote sensing technology not at the farm level, which would support management decisions. We recommend future studies to consider monitoring pasture at the paddock level.
The second barrier is the value proposition that needs to be understood by the researchers. Existing knowledge suggests that end users (i.e., farmers) think more of value proposition over the conventional methods (i.e., RPM, C-Dax, visual monitoring) before adopting the technology [87]. From the perspective of service providers, vendors (researchers, consultants, etc.) view remote sensing as input data with other accompanying spatial skills (geographical information system (GIS), information and communication technology (ICT), etc.) in providing end users (i.e., government, range managers, commercial farm enterprise, herders) with customer service that meets specified objectives. Such objectives include (a) providing information that supports the stocking rate and carrying capacity and (b) providing a monitoring system that can reduce degradation and conserve extensive grasslands [27]. End users (e.g., government) view this approach as a knowledge-based conservation strategy. For instance, [70] pointed out that the principal motivation for enacting conservation policies and creating political awareness in some countries is to reduce pasture degradation [75]. For example, China is re-enacting a legislative framework that will prohibit the institutional over-use of grassland that has degraded the country’s green land cover due to rapid industrialisation [22]. Summarily, commercial enterprises, satellite launchers, government agencies (Table 8), and service providers (i.e., Earth observation system (EOS), Cibo Labs, AgroInsider, SPACETM, DataFarming, GeoGraze, pasture.io, etc.) entering satellite markets in established countries (e.g., Australia, New Zealand, USA, European countries, etc.) could be the drivers of digital remote sensing of pasture monitoring, and its adoption by end users. High-tech companies such as Microsoft (Microsoft Planetary Computer), Google (Google Earth Engine), Amazon (Amazon Web Services), and Oracle (Oracle Cloud Infrastructure) with cloud computing services are providing applications to support digital agriculture. Social awareness, knowledge, skill, and well-defined research objectives are essential milestones to bring end users into the workflow.
Currently, ESA (i.e., Sen2Agri) and other global agencies provide a wide range of services to researchers and practitioners aiming to foster (R&D) to make Copernicus programs accessible to the worldwide community.

6. Conclusions

In summary, this review revealed the following trends and research opportunities.
This review simplified the remote sensing of managed global grasslands into four broad areas: management indicators, pasture production, species/botanical classification, and the degradation of pastures from anthropogenic and environmental variables.
In this review, less attention is received in Southeast Asia, the northern part of Latin America (except Mexico), the Middle East (except Iran and Syria), and Africa (except South Africa and Ethiopia).
Low-resolution multispectral sensors (e.g., MODIS and Landsat) are the most used due to availability and low cost. The higher resolution multispectral satellites are not free but are also constrained to a few bands (mainly visible and NIR), except Worldview and Venus, which have 29 and 12 bands.
SAR imagery, especially Sentinel 1 (publicly available), tended to be under-utilised. In particular, SAR data were not applied for mapping management traits and pasture degradation. The utility and accuracy of SAR modelling depend on the knowledge domain used to suit the biophysical variables and target environment.
The hyperspectral sensors used in this review were mainly applied for pasture composition due to the level of detail required.
Integrating multiple remote sensing tends to fix specific errors or limitations associated with sensors and the target environment. However, only some studies have combined sensors (SAR, multi and hyper-spectral images).
Less than 20 studies considered study areas that were less than 50 ha. The number of studies that used daily (i.e., 8) and weekly (i.e., 5) time-series remote sensing products is few, thus, making operational and automation a drawback.
Many studies that used machine learning approaches parameterized the empirical methods by selecting bands, thereby constraining this process to specific sites and parameters. Only a few studies used the characterization of vegetation properties based on reflectance values using spectral features to understand the relationship between pixels.
The size of field data used in most of the studies for validations is relatively small (test data reveal ≤ 120), thereby constraining remote sensing products regarding the robustness to capture image patterns and improve model calibration (for machine learning applications).
A few studies (18 studies) considered the adaptive management of pastures, which involved integrating remote sensing products with management and environmental data. It is recommended that future research efforts consider the integration of management and environmental data with remote sensing products for validation purposes and to make pasture management more sustainable.
This review identified that social awareness, knowledge, skill, and well-defined research objectives are essential milestones to bring end users into the workflow. We provided a list of agencies providing remote sensing services that can make the future of global monitoring of pastures more sustainable.
The remote sensing of pasture monitoring with satellites and UAS to derive biomass, LAI, fPAR, fCOVER, ANPP, and quantify physical quantity like pasture heights, discriminate vegetation canopy, manage pasture quality indicators (i.e., soil nitrogen, irrigation, soil water content, fertilizer application, mowing, etc.) and maintain pasture ecosystem from degradation has evolved. In this review, we provided a synthesis of how remote sensing can combine with modelling tools to facilitate the goal of digital agricultural sustainability.

Author Contributions

Conceptualization, methodology, formal analysis, investigation, writing—original draft preparation and editing, M.G.O.; visualization, M.G.O. and M.T.H.; review and editing, C.M.; review and editing, I.A.; review and editing, A.M.F. and M.T.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded through the University of Tasmania, Tasmanian Institute of Agriculture.

Data Availability Statement

Data used for analysis were extracted from list of referenced. Meta data is available upon request from the corresponding author.

Acknowledgments

M.G.O. acknowledges the funding received from the Australian Government and the University of Tasmania. Special acknowledgment to Bethany Melville for providing helpful feedback on the original draft of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. Flowchart describing the systematic literature process.
Figure A1. Flowchart describing the systematic literature process.
Remotesensing 15 04866 g0a1

References

  1. Rendel, J.; Mackay, A.; Smale, P.; Manderson, A.; Scobie, D. Optimisation of the resource of land-based livestock systems to advance sustainable agriculture: A farm-level analysis. Agriculture 2020, 10, 331. [Google Scholar] [CrossRef]
  2. Rigge, M.; Homer, C.; Shi, H.; Wylie, B. Departures of Rangeland Fractional Component Cover and Land Cover from Landsat-Based Ecological Potential in Wyoming, USA. Rangel. Ecol. Manag. 2020, 73, 856–870. [Google Scholar] [CrossRef]
  3. Chen, Y.; Guerschman, J.; Shendryk, Y.; Henry, D.; Harrison, M.T. Estimating pasture biomass using sentinel-2 imagery and machine learning. Remote Sens. 2021, 13, 603. [Google Scholar] [CrossRef]
  4. Sándor, R.; Ehrhardt, F.; Grace, P.; Recous, S.; Smith, P.; Snow, V.; Soussana, J.-F.; Basso, B.; Bhatia, A.; Brilli, L.; et al. Ensemble modelling of carbon fluxes in grasslands and croplands. Field Crop. Res. 2020, 252, 107791. [Google Scholar] [CrossRef]
  5. Harrison, M.T.; Cullen, B.R.; Mayberry, D.E.; Cowie, A.L.; Bilotto, F.; Badgery, W.B.; Liu, K.; Davison, T.; Christie, K.M.; Muleke, A.; et al. Carbon myopia: The urgent need for integrated social, economic and environmental action in the livestock sector. Glob. Chang. Biol. 2021, 27, 5726–5761. [Google Scholar] [CrossRef]
  6. Stanimirova, R.; Arévalo, P.; Kaufmann, R.K.; Maus, V.; Lesiv, M.; Havlík, P.; Friedl, M.A. Sensitivity of Global Pasturelands to Climate Variation. Earth’s Futur. 2019, 7, 1353–1366. [Google Scholar] [CrossRef]
  7. Spagnuolo, O.S.B.; Jarvey, J.C.; Battaglia, M.J.; Laubach, Z.M.; Miller, M.E.; Holekamp, K.E.; Bourgeau-Chavez, L.L. Mapping Kenyan Grassland Heights across large spatial scales with combined optical and radar satellite imagery. Remote Sens. 2020, 12, 1086. [Google Scholar] [CrossRef]
  8. Grigera, G.; Oesterheld, M.; Pacín, F. Monitoring forage production for farmers’ decision making. Agric. Syst. 2007, 94, 637–648. [Google Scholar] [CrossRef]
  9. Pfadenhauer, J.S.; Klötzli, F.A. Global Vegetation; Springer International: Berlin/Heidelberg, Germany, 2020; ISBN 9783030498603. [Google Scholar]
  10. Rawnsley, R.P.; Smith, A.P.; Christie, K.M.; Harrison, M.T.; Eckard, R.J. Current and future direction of nitrogen fertiliser use in Australian grazing systems. Crop Pasture Sci. 2019, 70, 1034–1043. [Google Scholar]
  11. Kurtz, D.B.; Schellberg, J.; Braun, M. Ground and satellite based assessment of rangeland management in sub-tropical Argentina. Appl. Geogr. 2010, 30, 210–220. [Google Scholar] [CrossRef]
  12. Allen, V.G.; Batello, C.; Berretta, E.J.; Hodgson, J.; Kothmann, M.; Li, X.; Mcivor, J.; Milne, J.; Morris, C.; Peeters, A.; et al. An international terminology for grazing lands and grazing animals. Grass Forage Sci. 2011, 66, 2–28. [Google Scholar] [CrossRef]
  13. Christie, K.M.; Rawnsley, R.P.; Harrison, M.T.; Eckard, R.J. Using a modelling approach to evaluate two options for improving animal nitrogen use efficiency and reducing nitrous oxide emissions on dairy farms in southern Australia. Anim. Prod. Sci. 2014, 54, 1960–1970. [Google Scholar] [CrossRef]
  14. Phelan, D.C.; Harrison, M.T.; McLean, G.; Cox, H.; Pembleton, K.G.; Dean, G.J.; Parsons, D.; do Amaral Richter, M.E.; Pengilley, G.; Hinton, S.J.; et al. Advancing a farmer decision support tool for agronomic decisions on rainfed and irrigated wheat cropping in Tasmania. Agric. Syst. 2018, 167, 113–124. [Google Scholar] [CrossRef]
  15. Bai, Z.G.; Dent, D.L.; Olsson, L.; Schaepman, M.E. Global Assessment of Land Degradation and Improvement: 1. Identification by Remote Sensing; ISRIC-World Soil Information: Wageningen, The Netherlands, 2008. [Google Scholar]
  16. Stellmes, M.; Udelhoven, T.; Röder, A.; Sonnenschein, R.; Hill, J. Dryland observation at local and regional scale—Comparison of Landsat TM/ETM+ and NOAA AVHRR time series. Remote Sens. Environ. 2010, 114, 2111–2125. [Google Scholar] [CrossRef]
  17. McClaran, M.P.; Wei, H. Recent drought phase in a 73-year record at two spatial scales: Implications for livestock production on rangelands in the Southwestern United States. Agric. For. Meteorol. 2014, 197, 40–51. [Google Scholar] [CrossRef]
  18. Zucca, C.; Wu, W.; Dessena, L.; Mulas, M. Assessing the Effectiveness of Land Restoration Interventions in Dry Lands by Multitemporal Remote Sensing—A Case Study in Ouled DLIM (Marrakech, Morocco). Land Degrad. Dev. 2015, 26, 80–91. [Google Scholar] [CrossRef]
  19. Huang, K.; Zhang, Y.; Zhu, J.; Liu, Y.; Zu, J.; Zhang, J. The influences of climate change and human activities on vegetation dynamics in the Qinghai-Tibet plateau. Remote Sens. 2016, 8, 876. [Google Scholar] [CrossRef]
  20. Fetzel, T.; Petridis, P.; Noll, D.; Singh, S.J.; Fischer-Kowalski, M. Reaching a socio-ecological tipping point: Overgrazing on the Greek island of Samothraki and the role of European agricultural policies. Land Use Policy 2018, 76, 21–28. [Google Scholar] [CrossRef]
  21. Jansen, V.S.; Kolden, C.A.; Schmalz, H.J. The development of near real-time biomass and cover estimates for adaptive rangeland management using Landsat 7 and Landsat 8 surface reflectance products. Remote Sens. 2018, 10, 1057. [Google Scholar] [CrossRef]
  22. Zhou, W.; Li, J.; Yue, T. Grassland Degradation Restoration and Constructing Green Ecological Protective Screen. In Remote Sensing Monitoring and Evaluation of Degraded Grassland in China; Springer Geography; Springer: Singapore, 2020; pp. 125–138. [Google Scholar] [CrossRef]
  23. Li, R.; Han, R.; Yu, Q.; Qi, S.; Guo, L. Spatial heterogeneous of ecological vulnerability in arid and semi-arid area: A case of the Ningxia Hui autonomous region, China. Sustainability 2020, 12, 4401. [Google Scholar] [CrossRef]
  24. Ara, I.; Turner, L.; Harrison, M.T.; Monjardino, M.; deVoil, P.; Rodriguez, D. Application, adoption and opportunities for improving decision support systems in irrigated agriculture: A review. Agric. Water Manag. 2021, 257, 107161. [Google Scholar] [CrossRef]
  25. Reinermann, S.; Asam, S.; Kuenzer, C. Remote sensing of grassland production and management—A review. Remote Sens. 2020, 12, 1949. [Google Scholar] [CrossRef]
  26. Langworthy, A.D.; Rawnsley, R.P.; Freeman, M.J.; Pembleton, K.G.; Corkrey, R.; Harrison, M.T.; Lane, P.A.; Henry, D.A. Potential of summer-active temperate (C3) perennial forages to mitigate the detrimental effects of supraoptimal temperatures on summer home-grown feed production in south-eastern Australian dairying regions. Crop Pasture Sci. 2018, 69, 808–820. [Google Scholar] [CrossRef]
  27. Hodgkinson, K.C.; Wang, D. Preventing rangeland degradation: A shared problem for Australia and China. Rangel. J. 2020, 42, 323–328. [Google Scholar] [CrossRef]
  28. Fernández-Giménez, M.E.; Venable, N.H.; Angerer, J.; Fassnacht, S.R.; Reid, R.S.; Khishigbayar, J. Exploring linked ecological and cultural tipping points in Mongolia. Anthropocene 2017, 17, 46–69. [Google Scholar] [CrossRef]
  29. Harrison, M.T.; Roggero, P.P.; Zavattaro, L. Simple, efficient and robust techniques for automatic multi-objective function parameterisation: Case studies of local and global optimisation using APSIM. Environ. Model. Softw. 2019, 117, 109–133. [Google Scholar] [CrossRef]
  30. Trotter, M.G.; Lamb, D.W.; Donald, G.E.; Schneider, D.A. Evaluating an active optical sensor for quantifying and mapping green herbage mass and growth in a perennial grass pasture. Crop Pasture Sci. 2010, 61, 389–398. [Google Scholar] [CrossRef]
  31. Edirisinghe, A.; Clark, D.; Waugh, D. Spatio-temporal modelling of biomass of intensively grazed perennial dairy pastures using multispectral remote sensing. Int. J. Appl. Earth Obs. Geoinf. 2012, 16, 5–16. [Google Scholar] [CrossRef]
  32. Punalekar, S.M.; Verhoef, A.; Quaife, T.L.; Humphries, D.; Bermingham, L.; Reynolds, C.K. Application of Sentinel-2A data for pasture biomass monitoring using a physically based radiative transfer model. Remote Sens. Environ. 2018, 218, 207–220. [Google Scholar] [CrossRef]
  33. Ara, I.; Harrison, M.T.; Whitehead, J.; Waldner, F.; Bridle, K.; Gilfedder, L.; Marques Da Silva, J.; Marques, F.; Rawnsley, R. Modelling seasonal pasture growth and botanical composition at the paddock scale with satellite imagery. In Silico Plants 2021, 3, diaa013. [Google Scholar] [CrossRef]
  34. Gaffney, R.; Porensky, L.M.; Gao, F.; Irisarri, J.G.; Durante, M.; Derner, J.D.; Augustine, D.J. Using MODIS imagery, climate and soil data to estimate pasture growth rates on farms in the south-west of Western Australia. Remote Sens. 2018, 10, 1474. [Google Scholar] [CrossRef]
  35. Liu, Q.; Yang, Z.; Han, F.; Shi, H.; Wang, Z.; Chen, X. Ecological environment assessment in world natural heritage site based on remote-sensing data. A case study from the Bayinbuluke. Sustainability 2019, 11, 6385. [Google Scholar] [CrossRef]
  36. Butterfield, H.S.; Malmström, C.M. The effects of phenology on indirect measures of aboveground biomass in annual grasses. Int. J. Remote Sens. 2009, 30, 3133–3146. [Google Scholar] [CrossRef]
  37. Flynn, E.S.; Dougherty, C.T.; Wendroth, O. Assessment of pasture biomass with the normalized difference vegetation index from active ground-based sensors. Agron. J. 2008, 100, 114–121. [Google Scholar] [CrossRef]
  38. Starks, P.J.; Zhao, D.; Brown, M.A. Estimation of nitrogen concentration and in vitro dry matter digestibility of herbage of warm-season grass pastures from canopy hyperspectral reflectance measurements. Grass Forage Sci. 2008, 63, 168–178. [Google Scholar] [CrossRef]
  39. Starks, P.J.; Zhao, D.; Phillips, W.A.; Coleman, S.W. Development of canopy reflectance algorithms for real-time prediction of bermudagrass pasture biomass and nutritive values. Crop Sci. 2006, 46, 927–934. [Google Scholar] [CrossRef]
  40. Pullanagari, R.R.; Yule, I.; King, W.; Dalley, D.; Dynes, R. The use of optical sensors to estimate pasture quality. Int. J. Smart Sens. Intell. Syst. 2011, 4, 125–137. [Google Scholar] [CrossRef]
  41. Thulin, S.; Hill, M.J.; Held, A.; Jones, S.; Woodgate, P. Hyperspectral determination of feed quality constituents in temperate pastures: Effect of processing methods on predictive relationships from partial least squares regression. Int. J. Appl. Earth Obs. Geoinf. 2012, 19, 322–334. [Google Scholar] [CrossRef]
  42. Pullanagari, R.R.; Yule, I.J.; Hedley, M.J.; Tuohy, M.P.; Dynes, R.A.; King, W.M. Multi-spectral radiometry to estimate pasture quality components. Precis. Agric. 2012, 13, 442–456. [Google Scholar] [CrossRef]
  43. Serrano, J.; Shahidian, S.; Marques Da Silva, J.; Sales-Baptista, E.; Ferraz De Oliveira, I.; Lopes De Castro, J.; Pereira, A.; Cancela De Abreu, M.; Machado, E.; Carvalho, M.D. Tree influence on soil and pasture: Contribution of proximal sensing to pasture productivity and quality estimation in montado ecosystems. Int. J. Remote Sens. 2018, 39, 4801–4829. [Google Scholar] [CrossRef]
  44. Guo, X.; Wilmshurst, J.F.; Li, Z. Comparison of laboratory and field remote sensing methods to measure forage quality. Int. J. Environ. Res. Public Health 2010, 7, 3513–3530. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, J.; Xiao, X.; Bajgain, R.; Starks, P.; Steiner, J.; Doughty, R.B.; Chang, Q. Estimating leaf area index and aboveground biomass of grazing pastures using Sentinel-1, Sentinel-2 and Landsat images. ISPRS J. Photogramm. Remote Sens. 2019, 154, 189–201. [Google Scholar] [CrossRef]
  46. Wachendorf, M.; Fricke, T.; Möckel, T. Remote sensing as a tool to assess botanical composition, structure, quantity and quality of temperate grasslands. Grass Forage Sci. 2018, 73, 1–14. [Google Scholar] [CrossRef]
  47. Moeckel, T.; Safari, H.; Reddersen, B.; Fricke, T.; Wachendorf, M. Fusion of ultrasonic and spectral sensor data for improving the estimation of biomass in grasslands with heterogeneous sward structure. Remote Sens. 2017, 9, 98. [Google Scholar] [CrossRef]
  48. Starks, P.J.; Zhao, D.; Phillips, W.A.; Coleman, S.W. Herbage mass, nutritive value and canopy spectral reflectance of bermudagrass pastures. Grass Forage Sci. 2006, 61, 101–111. [Google Scholar] [CrossRef]
  49. Wagle, P.; Gowda, P.H.; Northup, B.K.; Starks, P.J.; Neel, J.P.S. Response of tallgrass prairie to management in the U.S. Southern great plains: Site descriptions, management practices, and eddy covariance instrumentation for a Long-Term Experiment. Remote Sens. 2019, 11, 1988. [Google Scholar] [CrossRef]
  50. Serrano, J.; Shahidian, S.; da Silva, J.M.; Paixão, L.; Carreira, E.; Carmona-Cabezas, R.; Nogales-Bueno, J.; Rato, A.E. Evaluation of near infrared spectroscopy (NIRS) and remote sensing (RS) for estimating pasture quality in Mediterranean montado ecosystem. Appl. Sci. 2020, 10, 4463. [Google Scholar] [CrossRef]
  51. Röder, A.; Udelhoven, T.; Hill, J.; del Barrio, G.; Tsiourlis, G. Trend analysis of Landsat-TM and -ETM+ imagery to monitor grazing impact in a rangeland ecosystem in Northern Greece. Remote Sens. Environ. 2008, 112, 2863–2875. [Google Scholar] [CrossRef]
  52. Oldeland, J.; Dorigo, W.; Lieckfeld, L.; Lucieer, A.; Jürgens, N. Combining vegetation indices, constrained ordination and fuzzy classification for mapping semi-natural vegetation units from hyperspectral imagery. Remote Sens. Environ. 2010, 114, 1155–1166. [Google Scholar] [CrossRef]
  53. Kakembo, V.; Ndou, N. Relating vegetation condition to grazing management systems in the central Keiskamma Catchment, Eastern Cape Province, South Africa. Land Degrad. Dev. 2019, 30, 1052–1060. [Google Scholar] [CrossRef]
  54. Dorigo, W.; Richter, R.; Baret, F.; Bamler, R.; Wagner, W. Enhanced Automated Canopy Characterization from Hyperspectral Data by a Novel Two Step Radiative Transfer Model Inversion Approach. Remote Sens. 2009, 1, 1139–1170. [Google Scholar] [CrossRef]
  55. Boch, S.; Bedolla, A.; Ecker, K.T.; Ginzler, C.; Graf, U.; Küchler, H.; Küchler, M.; Nobis, M.P.; Holderegger, R.; Bergamini, A. Threatened and specialist species suffer from increased wood cover and productivity in Swiss steppes. Flora Morphol. Distrib. Funct. Ecol. Plants 2019, 258, 151444. [Google Scholar] [CrossRef]
  56. Redhead, J.; Cuevas-Gonzales, M.; Smith, G.; Gerard, F.; Pywell, R. Assessing the effectiveness of scrub management at the landscape scale using rapid field assessment and remote sensing. J. Environ. Manag. 2012, 97, 102–108. [Google Scholar] [CrossRef] [PubMed]
  57. Verrelst, J.; Malenovský, Z.; Van der Tol, C.; Camps-Valls, G.; Gastellu-Etchegorry, J.-P.; Lewis, P.; North, P.; Moreno, J. Quantifying Vegetation Biophysical Variables from Imaging Spectroscopy Data: A Review on Retrieval Methods. Surv. Geophys. 2019, 40, 589–629. [Google Scholar] [CrossRef] [PubMed]
  58. De Rosa, D.; Basso, B.; Fasiolo, M.; Friedl, J.; Fulkerson, B.; Grace, P.R.; Rowlings, D.W. Predicting pasture biomass using a statistical model and machine learning algorithm implemented with remotely sensed imagery. Comput. Electron. Agric. 2021, 180, 105880. [Google Scholar] [CrossRef]
  59. Schellberg, J.; Hill, M.J.; Gerhards, R.; Rothmund, M.; Braun, M. Precision agriculture on grassland: Applications, perspectives and constraints. Eur. J. Agron. 2008, 29, 59–71. [Google Scholar] [CrossRef]
  60. Jadhav, R.N.; Kimothi, M.M.; Kandya, A.K. Grassland mapping/monitoring of Banni, Kachchh (Gujarat) using remotely-sensed data. Int. J. Remote Sens. 1993, 14, 3093–3103. [Google Scholar] [CrossRef]
  61. Na, Y.; Li, J.; Hoshino, B.; Bao, S.; Qin, F.; Myagmartseren, P. Effects of different grazing systems on aboveground biomass and plant species dominance in typical Chinese and Mongolian steppes. Sustainability 2018, 10, 4753. [Google Scholar] [CrossRef]
  62. Xu, D.; Chen, B.; Shen, B.; Wang, X.; Yan, Y.; Xu, L.; Xin, X. The Classification of Grassland Types Based on Object-Based Image Analysis with Multisource Data. Rangel. Ecol. Manag. 2019, 72, 318–326. [Google Scholar] [CrossRef]
  63. Kimura, R.; Moriyama, M. Use of a satellite-based aridity index to monitor decreased soil water content and grass growth in grasslands of north-east Asia. Remote Sens. 2020, 12, 3556. [Google Scholar] [CrossRef]
  64. Wang, X.; Dong, J.; Baoyin, T.; Bao, Y. Estimation and climate factor contribution of aboveground biomass in Inner Mongolia’s typical/desert steppes. Sustainability 2019, 11, 6559. [Google Scholar] [CrossRef]
  65. Jansen, V.S.; Kolden, C.A.; Taylor, R.V.; Newingham, B.A. Quantifying livestock effects on bunchgrass vegetation with Landsat ETM+ data across a single growing season. Int. J. Remote Sens. 2016, 37, 150–175. [Google Scholar] [CrossRef]
  66. Xu, D.; Koper, N.; Guo, X. Quantifying the influences of grazing, climate and their interactions on grasslands using Landsat TM images. Grassl. Sci. 2018, 64, 118–127. [Google Scholar] [CrossRef]
  67. Reeves, M.C.; Hanberry, B.B.; Wilmer, H.; Kaplan, N.E.; Lauenroth, W.K. An Assessment of Production Trends on the Great Plains from 1984 to 2017. Rangel. Ecol. Manag. 2020, 78, 165–179. [Google Scholar] [CrossRef]
  68. Guo, X.; Price, K.P.; Stiles, J. Grasslands Discriminant Analysis Using Landsat TM Single and Multitemporal Data. Photogramm. Eng. Remote Sens. 2003, 69, 1255–1262. [Google Scholar] [CrossRef]
  69. Chu, D. Aboveground biomass estimates of grassland in the north tibet using modis remote sensing approaches. Appl. Ecol. Environ. Res. 2020, 18, 7655–7672. [Google Scholar] [CrossRef]
  70. Wei, D.; Zhao, H.; Zhang, J.; Qi, Y.; Wang, X. Human activities alter response of alpine grasslands on Tibetan Plateau to climate change. J. Environ. Manag. 2020, 262, 110335. [Google Scholar] [CrossRef]
  71. Lyu, X.; Li, X.; Gong, J.; Wang, H.; Dang, D.; Dou, H.; Li, S.; Liu, S. Comprehensive grassland degradation monitoring by remote sensing in Xilinhot, Inner Mongolia, China. Sustainability 2020, 12, 3682. [Google Scholar] [CrossRef]
  72. Ma, Q.; Chai, L.; Hou, F.; Chang, S.; Ma, Y.; Tsunekawa, A.; Cheng, Y. Quantifying grazing intensity using remote sensing in alpine meadows on Qinghai-Tibetan Plateau. Sustainability 2019, 11, 417. [Google Scholar] [CrossRef]
  73. Dong, S.; Shang, Z.; Gao, J.; Boone, R.B. Enhancing sustainability of grassland ecosystems through ecological restoration and grazing management in an era of climate change on Qinghai-Tibetan Plateau. Agric. Ecosyst. Environ. 2020, 287, 106684. [Google Scholar] [CrossRef]
  74. Bretas, I.L.; Valente, D.S.M.; Silva, F.F.; Chizzotti, M.L.; Paulino, M.F.; D’Áurea, A.P.; Paciullo, D.S.C.; Pedreira, B.C.; Chizzotti, F.H.M. Prediction of aboveground biomass and dry-matter content in brachiaria pastures by combining meteorological data and satellite imagery. Grass Forage Sci. 2021, 76, 340–352. [Google Scholar] [CrossRef]
  75. Hott, M.C.; Carvalho, L.M.T.; Antunes, M.A.H.; Resende, J.C.; Rocha, W.S.D. Analysis of grassland degradation in Zona da Mata, MG, Brazil, based on NDVI time series data with the integration of phenological metrics. Remote Sens. 2019, 11, 2956. [Google Scholar] [CrossRef]
  76. Sibanda, M.; Onisimo, M.; Dube, T.; Mabhaudhi, T. Quantitative assessment of grassland foliar moisture parameters as an inference on rangeland condition in the mesic rangelands of southern Africa. Int. J. Remote Sens. 2021, 42, 1474–1491. [Google Scholar] [CrossRef]
  77. Sandhage-Hofmann, A.; Löffler, J.; Kotzé, E.; Weijers, S.; Wingate, V.; Wundram, D.; Weihermüller, L.; Pape, R.; du Preez, C.C.; Amelung, W. Woody encroachment and related soil properties in different tenure-based management systems of semiarid rangelands. Geoderma 2020, 372, 114399. [Google Scholar] [CrossRef]
  78. Röder, A.; Kuemmerle, T.; Hill, J.; Papanastasis, V.P.; Tsiourlis, G.M. Adaptation of a grazing gradient concept to heterogeneous Mediterranean rangelands using cost surface modelling. Ecol. Modell. 2007, 204, 387–398. [Google Scholar] [CrossRef]
  79. Dostine, P.L.; Woinarski, J.C.Z.; Mackey, B.; Nix, H. Patterns of grassland productivity, composition and seed abundance, and the diet of the flock bronzewing pigeon Phaps histrionica at one site in northern Australia over a period of marked seasonal change. Wildl. Res. 2014, 41, 343–355. [Google Scholar] [CrossRef]
  80. Watson, C.J.; Restrepo-Coupe, N.; Huete, A.R. Multi-scale phenology of temperate grasslands: Improving monitoring and management with near-surface phenocams. Front. Environ. Sci. 2019, 7, 14. [Google Scholar] [CrossRef]
  81. Donald, G.E.; Scott, J.M.; Vickery, P.J. Satellite derived evidence of whole farmlet and paddock responses to management and climate. Anim. Prod. Sci. 2013, 53, 699–710. [Google Scholar] [CrossRef]
  82. Gallego, F.; Paruelo, J.M.; Baeza, S.; Altesor, A. Distinct ecosystem types respond differentially to grazing exclosure. Austral Ecol. 2020, 45, 548–556. [Google Scholar] [CrossRef]
  83. Castro, M.; Ameray, A.; Castro, J.P. A new approach to quantify grazing pressure under mediterranean pastoral systems using GIS and remote sensing. Int. J. Remote Sens. 2020, 41, 5371–5387. [Google Scholar] [CrossRef]
  84. Fenetahun, Y.; Xu, X.W.; Wang, Y.D. Analysis of eco-environmental vulnerability: Implication for bush encroachment and livestock population dynamics of the teltele rangeland, southern, ethiopia. Appl. Ecol. Environ. Res. 2020, 18, 7255–7278. [Google Scholar] [CrossRef]
  85. Wall, A.J.; Asher, G.W.; Netzer, M.S.; Johnson, M.G.H.; O’neill, K.T.; Littlejohn, R.P.; Cox, N. Farmed red deer home range, habitat use and daily movement patterns in a southland, New Zealand, tussock grassland over calving and lactation. Anim. Prod. Sci. 2018, 59, 549–563. [Google Scholar] [CrossRef]
  86. Schucknecht, A.; Meroni, M.; Kayitakire, F.; Boureima, A. Phenology-based biomass estimation to support rangeland management in semi-arid environments. Remote Sens. 2017, 9, 463. [Google Scholar] [CrossRef]
  87. Hill, M.J.; Donald, G.E.; Vickery, P.J.; Furnival, E.P. Integration of satellite remote sensing, simple bioclimatic models and GIS for assessment of pastoral development for a commercial grazing enterprise. Aust. J. Exp. Agric. 1996, 36, 309–321. [Google Scholar] [CrossRef]
  88. Doan, T.; Guo, X. Understanding Bison Carrying Capacity Estimation in Northern Great Plains Using Remote Sensing and GIS. Can. J. Remote Sens. 2019, 45, 139–162. [Google Scholar] [CrossRef]
  89. Goodrich, D.C.; Wei, H.; Burns, I.S.; Guertin, D.P.; Spaeth, K.; Hernandez, M.; Holifield-Collins, C.; Kautz, M.; Heilman, P.; Levick, L.R.; et al. Evaluation of Conservation Effects Assessment Project Grazing Lands conservation practices on the Cienega Creek watershed in southeast Arizona with AGWA/RHEM modeling tools. J. Soil Water Conserv. 2020, 75, 304–318. [Google Scholar] [CrossRef]
  90. de Leeuw, J.; Rizayeva, A.; Namazov, E.; Bayramov, E.; Marshall, M.T.; Etzold, J.; Neudert, R. Application of the MODIS MOD 17 Net Primary Production product in grassland carrying capacity assessment. Int. J. Appl. Earth Obs. Geoinf. 2019, 78, 66–76. [Google Scholar] [CrossRef]
  91. Geerken, R.; Ilaiwi, M. Assessment of rangeland degradation and development of a strategy for rehabilitation. Remote Sens. Environ. 2004, 90, 490–504. [Google Scholar] [CrossRef]
  92. Jones, M.O.; Allred, B.W.; Naugle, D.E.; Maestas, J.D.; Donnelly, P.; Metz, L.J.; Karl, J.; Smith, R.; Bestelmeyer, B.; Boyd, C.; et al. Innovation in rangeland monitoring: Annual, 30 m, plant functional type percent cover maps for U.S. rangelands, 1984–2017. Ecosphere 2018, 9, e02430. [Google Scholar] [CrossRef]
  93. Mekuyie, M.; Jordaan, A.; Melka, Y. Land-use and land-cover changes and their drivers in rangeland-dependent pastoral communities in the southern Afar Region of Ethiopia. Afr. J. Range Forage Sci. 2018, 35, 33–43. [Google Scholar] [CrossRef]
  94. Khishigbayar, J.; Fernández-Giménez, M.E.; Angerer, J.P.; Reid, R.S.; Chantsallkham, J.; Baasandorj, Y.; Zumberelmaa, D. Mongolian rangelands at a tipping point? Biomass and cover are stable but composition shifts and richness declines after 20years of grazing and increasing temperatures. J. Arid. Environ. 2015, 115, 100–112. [Google Scholar] [CrossRef]
  95. Liu, M.; Dries, L.; Heijman, W.; Huang, J.; Zhu, X.; Hu, Y.; Chen, H. The Impact of Ecological Construction Programs on Grassland Conservation in Inner Mongolia, China. Land Degrad. Dev. 2018, 29, 326–336. [Google Scholar] [CrossRef]
  96. An, N.; Price, K.P.; Blair, J.M. Estimating above-ground net primary productivity of the tallgrass prairie ecosystem of the Central Great Plains using AVHRR NDVI. Int. J. Remote Sens. 2013, 34, 3717–3735. [Google Scholar] [CrossRef]
  97. An, R.; Wang, H.-L.; Feng, X.-Z.; Wu, H.; Wang, Z.; Wang, Y.; Shen, X.-J.; Lu, C.-H.; Quaye-Ballard, J.A.; Chen, Y.-H.; et al. Monitoring rangeland degradation using a novel local NPP scaling based scheme over the “Three-River Headwaters” region, hinterland of the Qinghai-Tibetan Plateau. Quat. Int. 2017, 444, 97–114. [Google Scholar] [CrossRef]
  98. Feng, Y.; Wu, J.; Zhang, J.; Zhang, X.; Song, C. Identifying the relative contributions of climate and grazing to both direction and magnitude of Alpine grassland productivity dynamics from 1993 to 2011 on the Northern Tibetan Plateau. Remote Sens. 2017, 9, 136. [Google Scholar] [CrossRef]
  99. Diouf, A.A.; Hiernaux, P.; Brandt, M.; Faye, G.; Djaby, B.; Diop, M.B.; Ndione, J.A.; Tychon, B. Do agrometeorological data improve optical satellite-based estimations of the herbaceous yield in Sahelian semi-arid ecosystems? Remote Sens. 2016, 8, 668. [Google Scholar] [CrossRef]
  100. Yang, X.; Guo, X.; Fitzsimmons, M. Assessing light to moderate grazing effects on grassland production using satellite imagery. Int. J. Remote Sens. 2012, 33, 5087–5104. [Google Scholar] [CrossRef]
  101. Nightingale, J.M.; Phinn, S.R. Assessment of relationships between precipitation and satellite derived vegetation condition within South Australia. Aust. Geogr. Stud. 2003, 41, 180–195. [Google Scholar] [CrossRef]
  102. De Keersmaecker, W.; van Rooijen, N.; Lhermitte, S.; Tits, L.; Schaminée, J.; Coppin, P.; Honnay, O.; Somers, B. Species-rich semi-natural grasslands have a higher resistance but a lower resilience than intensively managed agricultural grasslands in response to climate anomalies. J. Appl. Ecol. 2016, 53, 430–439. [Google Scholar] [CrossRef]
  103. Shrestha, S.; Rahimzadeh-Bajgiran, P.; De Urioste-Stone, S. Probing recent environmental changes and resident perceptions in Upper Himalaya, Nepal. Remote Sens. Appl. Soc. Environ. 2020, 18, 100315. [Google Scholar] [CrossRef]
  104. Feng, X.M.; Zhao, Y.S. Grazing intensity monitoring in northern China steppe: Integrating CENTURY model and MODIS data. Ecol. Indic. 2011, 11, 175–182. [Google Scholar] [CrossRef]
  105. Fenetahun, Y.; Yong-Dong, W.; You, Y.; Xinwen, X. Dynamics of forage and land cover changes in Teltele district of Borana rangelands, southern Ethiopia: Using geospatial and field survey data. BMC Ecol. 2020, 20, 55. [Google Scholar] [CrossRef] [PubMed]
  106. Akinyemi, F.O.; Kgomo, M.O. Vegetation dynamics in African drylands: An assessment based on the Vegetation Degradation Index in an agro-pastoral region of Botswana. Reg. Environ. Chang. 2019, 19, 2027–2039. [Google Scholar] [CrossRef]
  107. Ye, H.; Huang, X.-T.; Luo, G.-P.; Wang, J.-B.; Zhang, M.; Wang, X.-X. Improving remote sensing-based net primary production estimation in the grazed land with defoliation formulation model. J. Mt. Sci. 2019, 16, 323–336. [Google Scholar] [CrossRef]
  108. Shoko, C.; Mutanga, O.; Dube, T. Remotely sensed C3 and C4 grass species aboveground biomass variability in response to seasonal climate and topography. Afr. J. Ecol. 2019, 57, 477–489. [Google Scholar] [CrossRef]
  109. Cao, Y.; Wu, J.; Zhang, X.; Niu, B.; Li, M.; Zhang, Y.; Wang, X.; Wang, Z. Dynamic forage-livestock balance analysis in alpine grasslands on the Northern Tibetan Plateau. J. Environ. Manag. 2019, 238, 352–359. [Google Scholar] [CrossRef]
  110. Li, C.; de Jong, R.; Schmid, B.; Wulf, H.; Schaepman, M.E. Spatial variation of human influences on grassland biomass on the Qinghai-Tibetan plateau. Sci. Total Environ. 2019, 665, 678–689. [Google Scholar] [CrossRef]
  111. Donald, G.E.; Gherardi, S.G.; Edirisinghe, A.; Gittins, S.P.; Henry, D.A.; Mata, G. Using MODIS imagery, climate and soil data to estimate pasture growth rates on farms in the south-west of Western Australia. Anim. Prod. Sci. 2010, 50, 611–615. [Google Scholar] [CrossRef]
  112. Addimando, N.; Nana, E.; Bocchiola, D. Modeling pasture dynamics in a mediterranean environment: Case study in Sardinia, Italy. J. Irrig. Drain. Eng. 2015, 141, 04014063. [Google Scholar] [CrossRef]
  113. Marston, C.G.; Aplin, P.; Wilkinson, D.M.; Field, R.; O’Regan, H.J. Scrubbing up: Multi-scale investigation of woody encroachment in a Southern African savannah. Remote Sens. 2017, 9, 419. [Google Scholar] [CrossRef]
  114. Liu, H.; Dahlgren, R.A.; Larsen, R.E.; Devine, S.M.; Roche, L.M.; O’ Geen, A.T.; Wong, A.J.Y.; Covello, S.; Jin, Y. Estimating rangeland forage production using remote sensing data from a Small Unmanned Aerial System (sUAS) and planetscope satellite. Remote Sens. 2019, 11, 595. [Google Scholar] [CrossRef]
  115. Dieguez, F.J.; Pereira, M. Uruguayan native grasslands net aerial primary production model and its application on safe stocking rate concept. Ecol. Modell. 2020, 430, 109060. [Google Scholar] [CrossRef]
  116. Gillan, J.K.; Karl, J.W.; van Leeuwen, W.J.D. Integrating drone imagery with existing rangeland monitoring programs. Environ. Monit. Assess. 2020, 192, 269. [Google Scholar] [CrossRef] [PubMed]
  117. Yu, R.; Evans, A.J.; Malleson, N. An agent-based model for assessing grazing strategies and institutional arrangements in Zeku, China. Agric. Syst. 2019, 171, 135–142. [Google Scholar] [CrossRef]
  118. Zhang, B.; Zhang, L.; Xie, D.; Yin, X.; Liu, C.; Liu, G. Application of synthetic NDVI time series blended from landsat and MODIS data for grassland biomass estimation. Remote Sens. 2016, 8, 10. [Google Scholar] [CrossRef]
  119. Sanderson, M.A.; Liebig, M.A.; Hendrickson, J.R.; Kronberg, S.L.; Toledo, D.; Derner, J.D.; Reeves, J.L. Long-term agroecosystem research on northern great plains mixed-grass prairie near mandan, north dakota. Can. J. Plant Sci. 2015, 95, 1101–1116. [Google Scholar] [CrossRef]
  120. Pellissier, P.A.; Ollinger, S.V.; Lepine, L.C.; Palace, M.W.; McDowell, W.H. Remote sensing of foliar nitrogen in cultivated grasslands of human dominated landscapes. Remote Sens. Environ. 2015, 167, 88–97. [Google Scholar] [CrossRef]
  121. Fassnacht, F.E.; Li, L.; Fritz, A. Mapping degraded grassland on the Eastern Tibetan Plateau with multi-temporal Landsat 8 data—Where do the severely degraded areas occur? Int. J. Appl. Earth Obs. Geoinf. 2015, 42, 115–127. [Google Scholar] [CrossRef]
  122. Wylie, B.K.; Boyte, S.P.; Major, D.J. Ecosystem performance monitoring of rangelands by integrating modeling and remote sensing. Rangel. Ecol. Manag. 2012, 65, 241–252. [Google Scholar] [CrossRef]
  123. Hudson, T.D.; Reeves, M.C.; Hall, S.A.; Yorgey, G.G.; Neibergs, J.S. Big landscapes meet big data: Informing grazing management in a variable and changing world. Rangelands 2021, 43, 17–28. [Google Scholar] [CrossRef]
  124. Dingaan, M.N.V.; Tsubo, M. Improved assessment of pasture availability in semi-arid grassland of South Africa. Environ. Monit. Assess. 2019, 191, 733. [Google Scholar] [CrossRef] [PubMed]
  125. Jafari, R.; Bashari, H.; Tarkesh, M. Discriminating and monitoring rangeland condition classes with MODIS NDVI and EVI indices in Iranian arid and semi-arid lands. Arid. Land Res. Manag. 2017, 31, 94–110. [Google Scholar] [CrossRef]
  126. Rigge, M.; Smart, A.; Wylie, B.; Gilmanov, T.; Johnson, P. Linking phenology and biomass productivity in south dakota mixed-grass prairie. Rangel. Ecol. Manag. 2013, 66, 579–587. [Google Scholar] [CrossRef]
  127. Minor, T.B.; Lancaster, J.; Wade, T.G.; Wickham, J.D.; Whitford, W.; Jones, K.B. Evaluating change in rangeland condition using multitemporal AVHRR data and geographic information system analysis. Environ. Monit. Assess. 1999, 59, 211–223. [Google Scholar] [CrossRef]
  128. Duan, C.; Shi, P.; Song, M.; Zhang, X.; Zong, N.; Zhou, C. Land use and land cover change in the Kailash Sacred Landscape of China. Sustainability 2019, 11, 1788. [Google Scholar] [CrossRef]
  129. Elmore, A.J.; Asner, G.P. Effects of grazing intensity on soil carbon stocks following deforestation of a Hawaiian dry tropical forest. Glob. Chang. Biol. 2006, 12, 1761–1772. [Google Scholar] [CrossRef]
  130. Liu, N.; Harper, R.J.; Handcock, R.N.; Evans, B.; Sochacki, S.J.; Dell, B.; Walden, L.L.; Liu, S. Seasonal timing for estimating carbon mitigation in revegetation of abandoned agricultural land with high spatial resolution remote sensing. Remote Sens. 2017, 9, 545. [Google Scholar] [CrossRef]
  131. Möckel, T.; Dalmayne, J.; Schmid, B.C.; Prentice, H.C.; Hall, K. Airborne hyperspectral data predict fine-scale plant species diversity in grazed dry grasslands. Remote Sens. 2016, 8, 133. [Google Scholar] [CrossRef]
  132. Raab, C.; Riesch, F.; Tonn, B.; Barrett, B.; Meißner, M.; Balkenhol, N.; Isselstein, J. Target-oriented habitat and wildlife management: Estimating forage quantity and quality of semi-natural grasslands with Sentinel-1 and Sentinel-2 data. Remote Sens. Ecol. Conserv. 2020, 6, 381–398. [Google Scholar] [CrossRef]
  133. Wakulinśka, M.; Marcinkowska-Ochtyra, A. Multi-temporal sentinel-2 data in classification of mountain vegetation. Remote Sens. 2020, 12, 2696. [Google Scholar] [CrossRef]
  134. Bayle, A.; Carlson, B.Z.; Thierion, V.; Isenmann, M.; Choler, P. Improved mapping of mountain shrublands using the sentinel-2 red-edge band. Remote Sens. 2019, 11, 2807. [Google Scholar] [CrossRef]
  135. Klingler, A.; Schaumberger, A.; Vuolo, F.; Kalmár, L.B.; Pötsch, E.M. Comparison of Direct and Indirect Determination of Leaf Area Index in Permanent Grassland. PFG-J. Photogramm. Remote Sens. Geoinf. Sci. 2020, 88, 369–378. [Google Scholar] [CrossRef]
  136. Rufin, P.; Müller, H.; Pflugmacher, D.; Hostert, P. Land use intensity trajectories on Amazonian pastures derived from Landsat time series. Int. J. Appl. Earth Obs. Geoinf. 2015, 41, 1–10. [Google Scholar] [CrossRef]
  137. Wang, J.; Li, A.; Bian, J. Simulation of the grazing effects on grassland aboveground net primary production using DNDC model combined with time-series remote sensing data-a case study in Zoige plateau, China. Remote Sens. 2016, 8, 168. [Google Scholar] [CrossRef]
  138. Crabbe Richard, A.; Lamb David, W.; Edwards, C. Discriminating between C3, C4, and Mixed C3/C4 Pasture Grasses of a Grazed Landscape Using Multi-Temporal Sentinel-1a Data Richard. Remote Sens. 2019, 11, 253. [Google Scholar] [CrossRef]
  139. Crabbe, R.A.; Lamb, D.; Edwards, C. Discrimination of species composition types of a grazed pasture landscape using Sentinel-1 and Sentinel-2 data. Int. J. Appl. Earth Obs. Geoinf. 2020, 84, 101978. [Google Scholar] [CrossRef]
  140. Chabalala, Y.; Adam, E.; Oumar, Z.; Ramoelo, A. Exploiting the capabilities of Sentinel-2 and RapidEye for predicting grass nitrogen across different grass communities in a protected area. Appl. Geomatics 2020, 12, 379–395. [Google Scholar] [CrossRef]
  141. Guerschman, J.P.; Hill, M.J.; Renzullo, L.J.; Barrett, D.J.; Marks, A.S.; Botha, E.J. Estimating fractional cover of photosynthetic vegetation, non-photosynthetic vegetation and bare soil in the Australian tropical savanna region upscaling the EO-1 Hyperion and MODIS sensors. Remote Sens. Environ. 2009, 113, 928–945. [Google Scholar] [CrossRef]
  142. Pullanagari, R.R.; Kereszturi, G.; Yule, I.J. Mapping of macro and micro nutrients of mixed pastures using airborne AisaFENIX hyperspectral imagery. ISPRS J. Photogramm. Remote Sens. 2016, 117, 1–10. [Google Scholar] [CrossRef]
  143. Sivanpillai, R.; Booth, D.T. Characterizing rangeland vegetation using Landsat and 1-mm VLSA data in central Wyoming (USA). Agrofor. Syst. 2008, 73, 55–64. [Google Scholar] [CrossRef]
  144. Hagen, S.C.; Heilman, P.; Marsett, R.; Torbick, N.; Salas, W.; Van Ravensway, J.; Qi, J. Mapping total vegetation cover across western rangelands with moderate-resolution imaging spectroradiometer data. Rangel. Ecol. Manag. 2012, 65, 456–467. [Google Scholar] [CrossRef]
  145. Mueller, T.; Olson, K.A.; Fuller, T.K.; Schaller, G.B.; Murray, M.G.; Leimgruber, P. In search of forage: Predicting dynamic habitats of Mongolian gazelles using satellite-based estimates of vegetation productivity. J. Appl. Ecol. 2008, 45, 649–658. [Google Scholar] [CrossRef]
  146. Tiscornia, G.; Baethgen, W.; Ruggia, A.; Do Carmo, M.; Ceccato, P. Can we monitor height of native grasslands in Uruguay with earth observation? Remote Sens. 2019, 11, 1801. [Google Scholar] [CrossRef]
  147. Hunt, E.R., Jr.; Everitt, J.H.; Ritchie, J.C.; Moran, M.S.; Booth, D.T.; Anderson, G.L.; Clark, P.E.; Seyfried, M.S. Applications and Research Using Remote Sensing for Rangeland Management. Photogramm. Eng. Remote Sens. 2003, 69, 675–693. [Google Scholar] [CrossRef]
  148. Meshesha, D.T.; Ahmed, M.M.; Abdi, D.Y.; Haregeweyn, N. Prediction of grass biomass from satellite imagery in Somali regional state, eastern Ethiopia. Heliyon 2020, 6, e05272. [Google Scholar] [CrossRef]
  149. Hanna, M.M.; Steyn-Ross, D.A.; Steyn-Ross, M. Estimating biomass for New Zealand pasture using optical remote sensing techniques. Geocarto Int. 1999, 14, 89–94. [Google Scholar] [CrossRef]
  150. Yang, X.; Guo, X. Investigating vegetation biophysical and spectral parameters for detecting light to moderate grazing effects: A case study in mixed grass prairie. Cent. Eur. J. Geosci. 2011, 3, 336–348. [Google Scholar] [CrossRef]
  151. Lopes, M.; Fauvel, M.; Ouin, A.; Girard, S. Spectro-temporal heterogeneity measures from dense high spatial resolution satellite image time series: Application to grassland species diversity estimation. Remote Sens. 2017, 9, 993. [Google Scholar] [CrossRef]
  152. Maynard, C.L.; Lawrence, R.L.; Nielsen, G.A.; Decker, G. Ecological site descriptions and remotely sensed imagery as a tool for rangeland evaluation. Can. J. Remote Sens. 2007, 33, 109–115. [Google Scholar] [CrossRef]
  153. Smith, A.M.; Major, D.J.; McNeil, R.L.; Willms, W.D.; Brisco, B.; Brown, R.J. Complementarity of radar and visible-infrared sensors in assessing rangeland condition. Remote Sens. Environ. 1995, 52, 173–180. [Google Scholar] [CrossRef]
  154. Aragón, R.; Oesterheld, M. Linking vegetation heterogeneity and functional attributes of temperate grasslands through remote sensing. Appl. Veg. Sci. 2008, 11, 117–130. [Google Scholar] [CrossRef]
  155. Jin, Y.; Yang, X.; Li, Z.; Qin, Z.; Zhang, H.; Xu, B. Remote sensing estimation of forage mass and spatiotemporal change analysis in the Beijing-Tianjin sandstorm source region, China. Int. J. Remote Sens. 2019, 40, 2212–2226. [Google Scholar] [CrossRef]
  156. Lebed, L.; Qi, J.; Heilman, P. An ecological assessment of pasturelands in the Balkhash area of Kazakhstan with remote sensing and models. Environ. Res. Lett. 2012, 7, 025203. [Google Scholar] [CrossRef]
  157. Sant, E.D.; Simonds, G.E.; Ramsey, R.D.; Larsen, R.T. Assessment of sagebrush cover using remote sensing at multiple spatial and temporal scales. Ecol. Indic. 2014, 43, 297–305. [Google Scholar] [CrossRef]
  158. Lal, J.B.; Gulati, A.K.; Bist, M.S. Satellite mapping of alpine pastures in the himalayas. Int. J. Remote Sens. 1991, 12, 435–443. [Google Scholar] [CrossRef]
  159. Yool, S.R.; Makaio, M.J.; Watts, J.M. Techniques for computer-assisted mapping of rangeland change. J. Range Manag. 1997, 50, 307–314. [Google Scholar] [CrossRef]
  160. Ringrose, S.; Musisi-Nkambwe, S.; Coleman, T.; Nellis, C.; Bussing, D. Use of landsat thematic mapper data to assess seasonal rangeland changes in the Southeast Kalahari, Botswana. Environ. Manag. 1998, 23, 125–138. [Google Scholar] [CrossRef]
  161. Clark, P.E.; Seyfried, M.S.; Harris, B. Intermountain plant community classification using landsat TM and SPOT HRV data. J. Range Manag. 2001, 54, 152–160. [Google Scholar] [CrossRef]
  162. Marsett, R.C.; Qi, J.; Heilman, P.; Biedenbender, S.H.; Watson, M.C.; Amer, S.; Weltz, M.; Goodrich, D.; Marsett, R. Remote sensing for grassland management in the arid Southwest. Rangel. Ecol. Manag. 2006, 59, 530–540. [Google Scholar] [CrossRef]
  163. Karl, J.W. Spatial predictions of cover attributes of rangeland ecosystems using regression kriging and remote sensing. Rangel. Ecol. Manag. 2010, 63, 335–349. [Google Scholar] [CrossRef]
  164. Schwieder, M.; Buddeberg, M.; Kowalski, K.; Pfoch, K.; Bartsch, J.; Bach, H.; Pickert, J.; Hostert, P. Estimating Grassland Parameters from Sentinel-2: A Model Comparison Study. PFG-J. Photogramm. Remote Sens. Geoinf. Sci. 2020, 88, 379–390. [Google Scholar] [CrossRef]
  165. Rahetlah, V.B.; Salgado, P.; Andrianarisoa, B.; Tillard, E.; Razafindrazaka, H.; Le Mézo, L.; Ramalanjaona, V.L. Relationship between normalized difference vegetation index (NDVI) and forage biomass yield in the Vakinankaratra region, Madagascar. Livest. Res. Rural Dev. 2014, 26, 19. [Google Scholar]
  166. Hill, M.J.; Donald, G.E.; Hyder, M.W.; Smith, R.C.G. Estimation of pasture growth rate in the south west of Western Australia from AVHRR NDVI and climate data. Remote Sens. Environ. 2004, 93, 528–545. [Google Scholar] [CrossRef]
  167. Paruelo, J.M.; Oesterheld, M.; Di Bella, C.M.; Arzadum, M.; Lafontaine, J.; Cahuepé, M.; Rebella, C.M. Estimation of primary production of subhumid rangelands from remote sensing data. Appl. Veg. Sci. 2000, 3, 189–195. [Google Scholar] [CrossRef]
  168. Ambrosone, M.; Matese, A.; Di Gennaro, S.F.; Gioli, B.; Tudoroiu, M.; Genesio, L.; Miglietta, F.; Baronti, S.; Maienza, A.; Ungaro, F.; et al. Retrieving soil moisture in rainfed and irrigated fields using Sentinel-2 observations and a modified OPTRAM approach. Int. J. Appl. Earth Obs. Geoinf. 2020, 89, 102113. [Google Scholar] [CrossRef]
  169. Ai, Z.; An, R.; Chen, Y.; Huang, L. Comparison of hyperspectral HJ-1A/HSI and multispectral Landsat 8 and Sentinel-2A imagery for estimating alpine grassland coverage in the Three-River Headwaters region. J. Appl. Remote Sens. 2019, 13, 014504. [Google Scholar] [CrossRef]
  170. Zengeya, F.M.; Mutanga, O.; Murwira, A. Linking remotely sensed forage quality estimates from worldview-2 multispectral data with cattle distribution in a savanna landscape. Int. J. Appl. Earth Obs. Geoinf. 2012, 21, 513–524. [Google Scholar] [CrossRef]
  171. Dalmayne, J.; Möckel, T.; Prentice, H.C.; Schmid, B.C.; Hall, K. Assessment of fine-scale plant species beta diversity using WorldView-2 satellite spectral dissimilarity. Ecol. Inform. 2013, 18, 1–9. [Google Scholar] [CrossRef]
  172. Fava, F.; Pulighe, G.; Monteiro, A.T. Mapping Changes in Land Cover Composition and Pattern for Comparing Mediterranean Rangeland Restoration Alternatives. Land Degrad. Dev. 2016, 27, 671–681. [Google Scholar] [CrossRef]
  173. Psomas, A.; Kneubühler, M.; Huber, S.; Itten, K.; Zimmermann, N.E. Hyperspectral remote sensing for estimating aboveground biomass and for exploring species richness patterns of Grassland habitats. Int. J. Remote Sens. 2011, 32, 9007–9031. [Google Scholar] [CrossRef]
  174. Li, F.; Zheng, J.; Wang, H.; Luo, J.; Zhao, Y.; Zhao, R. Mapping grazing intensity using remote sensing in the Xilingol steppe region, Inner Mongolia, China. Remote Sens. Lett. 2016, 7, 328–337. [Google Scholar] [CrossRef]
  175. Moreau, S.; Le Toan, T. Biomass quantification of Andean wetland forages using ERS satellite SAR data for optimizing livestock management. Remote Sens. Environ. 2003, 84, 477–492. [Google Scholar] [CrossRef]
  176. Courault, D.; Hadria, R.; Ruget, F.; Olioso, A.; Duchemin, B.; Hagolle, O.; Dedieu, G. Combined use of FORMOSAT-2 images with a crop model for biomass and water monitoring of permanent grassland in Mediterranean region. Hydrol. Earth Syst. Sci. 2010, 14, 1731–1744. [Google Scholar] [CrossRef]
  177. Sibanda, M.; Mutanga, O.; Rouget, M. Discriminating Rangeland Management Practices Using Simulated HyspIRI, Landsat 8 OLI, Sentinel 2 MSI, and VENμS Spectral Data. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2016, 9, 3957–3969. [Google Scholar] [CrossRef]
  178. Li, F.; Zhao, Y.; Zheng, J.; Luo, J.; Zhang, X. Monitoring grazing intensity: An experiment with canopy spectra applied to satellite remote sensing. J. Appl. Remote Sens. 2016, 10, 026032. [Google Scholar] [CrossRef]
  179. Sibanda, M.; Mutanga, O.; Rouget, M. Comparing the spectral settings of the new generation broad and narrow band sensors in estimating biomass of native grasses grown under different management practices. GIScience Remote Sens. 2016, 53, 614–633. [Google Scholar] [CrossRef]
  180. Fadaei, H. A total ratio of vegetation index (TRVI) for shrubs sparse cover delineating in open woodland. J. Rangel. Sci. 2018, 8, 176–185. [Google Scholar]
  181. Madsen, B.; Treier, U.A.; Zlinszky, A.; Lucieer, A.; Normand, S. Detecting shrub encroachment in seminatural grasslands using UAS LiDAR. Ecol. Evol. 2020, 10, 4876–4902. [Google Scholar] [CrossRef]
  182. Vogel, S.; Gebbers, R.; Oertel, M.; Kramer, E. Evaluating soil-borne causes of biomass variability in Grassland by remote and proximal sensing. Sensors 2019, 19, 4593. [Google Scholar] [CrossRef]
  183. Gao, R.; Kong, Q.; Wang, H.; Su, Z. Diagnostic Feed Values of Natural Grasslands Based on Multispectral Images Acquired by Small Unmanned Aerial Vehicle. Rangel. Ecol. Manag. 2019, 72, 916–922. [Google Scholar] [CrossRef]
  184. Zlinszky, A.; Schroiff, A.; Kania, A.; Deák, B.; Mücke, W.; Vári, Á.; Székely, B.; Pfeifer, N. Categorizing grassland vegetation with full-waveform airborne laser scanning: A feasibility study for detecting natura 2000 habitat types. Remote Sens. 2014, 6, 8056–8087. [Google Scholar] [CrossRef]
  185. Gillan, J.K.; Karl, J.W.; Duniway, M.; Elaksher, A. Modeling vegetation heights from high resolution stereo aerial photography: An application for broad-scale rangeland monitoring. J. Environ. Manag. 2014, 144, 226–235. [Google Scholar] [CrossRef]
  186. Dawson, S.K.; Fisher, A.; Lucas, R.; Hutchinson, D.K.; Berney, P.; Keith, D.; Catford, J.A.; Kingsford, R.T. Remote sensing measures restoration successes, but canopy heights lag in restoring floodplain vegetation. Remote Sens. 2016, 8, 542. [Google Scholar] [CrossRef]
  187. Dusseux, P.; Zhao, Y.; Cordier, M.-O.; Corpetti, T.; Delaby, L.; Gascuel-Odoux, C.; Hubert-Moy, L. PaturMata, a model to manage grassland under climate change. Agron. Sustain. Dev. 2015, 35, 1087–1093. [Google Scholar] [CrossRef]
  188. Dusseux, P.; Vertès, F.; Corpetti, T.; Corgne, S.; Hubert-Moy, L. Agricultural practices in grasslands detected by spatial remote sensing. Environ. Monit. Assess. 2014, 186, 8249–8265. [Google Scholar] [CrossRef]
  189. Blanco, L.J.; Paruelo, J.M.; Oesterheld, M.; Biurrun, F.N. Spatial and temporal patterns of herbaceous primary production in semi-arid shrublands: A remote sensing approach. J. Veg. Sci. 2016, 27, 716–727. [Google Scholar] [CrossRef]
  190. Irisarri, J.G.N.; Oesterheld, M.; Paruelo, J.M.; Texeira, M.A. Patterns and controls of above-ground net primary production in meadows of Patagonia. A remote sensing approach. J. Veg. Sci. 2012, 23, 114–126. [Google Scholar] [CrossRef]
  191. Tsalyuk, M.; Kelly, M.; Koy, K.; Getz, W.M.; Scott Butterfield, H. Monitoring the impact of grazing on rangeland conservation easements using MODIS vegetation indices. Rangel. Ecol. Manag. 2015, 68, 173–185. [Google Scholar] [CrossRef]
  192. Karunaratne, S.; Thomson, A.; Morse-McNabb, E.; Wijesingha, J.; Stayches, D.; Copland, A.; Jacobs, J. The Fusion of Spectral and Structural Datasets Derived from an Airborne Multispectral Sensor for Estimation of Pasture Dry Matter Yield at Paddock Scale with Time Senani. Remote Sens. 2020, 12, 2017. [Google Scholar] [CrossRef]
  193. Gillan, J.K.; McClaran, M.P.; Swetnam, T.L.; Heilman, P. Estimating forage utilization with drone-based photogrammetric point clouds. Rangel. Ecol. Manag. 2019, 72, 575–585. [Google Scholar] [CrossRef]
  194. Liu, Y.; Feng, Q.; Wang, C.; Tang, Z. A risk-based model for grassland management using MODIS data: The case of Gannan region, China. Land use policy 2018, 72, 461–469. [Google Scholar] [CrossRef]
  195. Bao, N.; Li, W.; Gu, X.; Liu, Y. Biomass estimation for semiarid vegetation and mine rehabilitation using worldview-3 and sentinel-1 SAR imagery. Remote Sens. 2019, 11, 2855. [Google Scholar] [CrossRef]
  196. Mundava, C.; Schut, A.G.T.; Helmholz, P.; Stovold, R.; Donald, G.; Lamb, D.W. A novel protocol for assessment of aboveground biomass in rangeland environments. Rangel. J. 2015, 37, 157–167. [Google Scholar] [CrossRef]
  197. Shoko, C.; Mutanga, O.; Dube, T.; Slotow, R. Characterizing the spatio-temporal variations of C3 and C4 dominated grasslands aboveground biomass in the Drakensberg, South Africa. Int. J. Appl. Earth Obs. Geoinf. 2018, 68, 51–60. [Google Scholar] [CrossRef]
  198. Yin, G.; Li, A.; Wu, C.; Wang, J.; Xie, Q.; Zhang, Z.; Nan, X.; Jin, H.; Bian, J.; Lei, G. Seamless upscaling of the field-measured grassland aboveground biomass based on Gaussian process regression and gap-filled landsat 8 OLI reflectance. ISPRS Int. J. Geo-Inf. 2018, 7, 242. [Google Scholar] [CrossRef]
  199. Mansour, K.; Mutanga, O.; Everson, T.; Adam, E. Discriminating indicator grass species for rangeland degradation assessment using hyperspectral data resampled to AISA Eagle resolution. ISPRS J. Photogramm. Remote Sens. 2012, 70, 56–65. [Google Scholar] [CrossRef]
  200. Mirik, M.; Ansley, R.J. Comparison of ground-measured and image-classified mesquite (Prosopis glandulosa) canopy cover. Rangel. Ecol. Manag. 2012, 65, 85–95. [Google Scholar] [CrossRef]
  201. Espunyes, J.; Bartolomé, J.; Garel, M.; Gálvez-Cerón, A.; Aguilar, X.F.; Colom-Cadena, A.; Calleja, J.A.; Gassó, D.; Jarque, L.; Lavín, S.; et al. Seasonal diet composition of Pyrenean chamois is mainly shaped by primary production waves. PLoS ONE 2019, 14, e0210819. [Google Scholar] [CrossRef]
  202. Villamuelas, M.; Fernández, N.; Albanell, E.; Gálvez-Cerón, A.; Bartolomé, J.; Mentaberre, G.; López-Olvera, J.R.; Fernández-Aguilar, X.; Colom-Cadena, A.; López-Martín, J.M.; et al. The Enhanced Vegetation Index (EVI) as a proxy for diet quality and composition in a mountain ungulate. Ecol. Indic. 2016, 61, 658–666. [Google Scholar] [CrossRef]
  203. Han, W.; Lu, H.; Liu, G.; Wang, J.; Su, X. Quantifying Degradation Classifications on Alpine Grassland in the Lhasa River Basin, Qinghai-Tibetan Plateau. Sustainability 2019, 11, 7067. [Google Scholar] [CrossRef]
  204. Griffith, J.A.; Price, K.P.; Martinko, E.A. A multivariate analysis of biophysical parameters of tallgrass prairie among land management practices and years. Environ. Monit. Assess. 2001, 68, 249–271. [Google Scholar] [CrossRef] [PubMed]
  205. Durante, M.; Oesterheld, M.; Piñeiro, G.; Vassallo, M.M. Estimating forage quantity and quality under different stress and senescent biomass conditions via spectral reflectance. Int. J. Remote Sens. 2014, 35, 2963–2981. [Google Scholar] [CrossRef]
  206. Falldorf, T.; Strand, O.; Panzacchi, M.; Tømmervik, H. Estimating lichen volume and reindeer winter pasture quality from Landsat imagery. Remote Sens. Environ. 2014, 140, 573–579. [Google Scholar] [CrossRef]
  207. Sibanda, M.; Mutanga, O.; Rouget, M. Examining the potential of Sentinel-2 MSI spectral resolution in quantifying above ground biomass across different fertilizer treatments. ISPRS J. Photogramm. Remote Sens. 2015, 110, 55–65. [Google Scholar] [CrossRef]
  208. 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]
  209. Baeza, S.; Lezama, F.; Piñeiro, G.; Altesor, A.; Paruelo, J.M. Spatial variability of above-ground net primary production in Uruguayan grasslands: A remote sensing approach. Appl. Veg. Sci. 2010, 13, 72–85. [Google Scholar] [CrossRef]
  210. Robinson, N.P.; Jones, M.O.; Moreno, A.; Erickson, T.A.; Naugle, D.E.; Allred, B.W. Rangeland productivity partitioned to sub-pixel plant functional types. Remote Sens. 2019, 11, 1427. [Google Scholar] [CrossRef]
  211. Xu, B.; Yang, X.C.; Tao, W.G.; Miao, J.M.; Yang, Z.; Liu, H.Q.; Jin, Y.X.; Zhu, X.H.; Qin, Z.H.; Lv, H.Y.; et al. MODIS-based remote-sensing monitoring of the spatiotemporal patterns of China’s grassland vegetation growth. Int. J. Remote Sens. 2013, 34, 3867–3878. [Google Scholar] [CrossRef]
  212. Oesterheld, M.; Di Bella, C.M.; Kerdiles, H. Relation between NOAA-AVHRR satellite data and stocking rate of rangelands. Ecol. Appl. 1998, 8, 207–212. [Google Scholar] [CrossRef]
  213. Jacques, D.C.; Kergoat, L.; Hiernaux, P.; Mougin, E.; Defourny, P. Monitoring dry vegetation masses in semi-arid areas with MODIS SWIR bands. Remote Sens. Environ. 2014, 153, 40–49. [Google Scholar] [CrossRef]
  214. Zhang, B.; Carter, J. FORAGE—An online system for generating and delivering property-scale decision support information for grazing land and environmental management. Comput. Electron. Agric. 2018, 150, 302–311. [Google Scholar] [CrossRef]
  215. Higgins, S.; Schellberg, J.; Bailey, J.S. Improving productivity and increasing the efficiency of soil nutrient management on grassland farms in the UK and Ireland using precision agriculture technology. Eur. J. Agron. 2019, 106, 67–74. [Google Scholar] [CrossRef]
  216. Eastwood, C.; Dela, B.; Joanne, R. Developing an approach to assess farmer perceptions of the value of pasture assessment technologies. Grass Forage Sci. 2020, 75, 474–485. [Google Scholar] [CrossRef]
  217. Paltsyn, M.Y.; Gibbs, J.P.; Mountrakis, G. Integrating Traditional Ecological Knowledge and Remote Sensing for Monitoring Rangeland Dynamics in the Altai Mountain Region. Environ. Manag. 2019, 64, 40–51. [Google Scholar] [CrossRef]
  218. Butterfield, H.S.; Malmstrom, C.M. Experimental use of remote sensing by private range managers and its influence on management decisions. Rangel. Ecol. Manag. 2006, 59, 541–548. [Google Scholar] [CrossRef]
  219. Tiangang, L.; Quangong, C.; Jizhou, R.; Yuansu, W. A GIS-based expert system for pastoral agricultural development in Gansu Province, PR China. New Zealand J. Agric. Res. 2004, 47, 313–325. [Google Scholar] [CrossRef]
  220. Rasmussen, M.S.; James, R.; Adiyasuren, T.; Khishigsuren, P.; Naranchimeg, B.; Gankhuyag, R.; Baasanjargal, B. Supporting Mongolian pastoralists by using GIS to identify grazing limitations and opportunities from livestock census and remote sensing data. GeoJournal 1999, 47, 563–571. [Google Scholar] [CrossRef]
  221. Jacobsen, A.; Jacobsen, A.; Nielsen, A.; Ejmais, R.; Groom, G.B. Spectral identification of plant communities for mapping of semi-natural grasslands. Can. J. Remote Sens. 2000, 26, 370–383. [Google Scholar] [CrossRef]
  222. Hall, K.; Reitalu, T.; Sykes, M.T.; Prentice, H.C. Spectral heterogeneity of QuickBird satellite data is related to fine-scale plant species spatial turnover in semi-natural grasslands. Appl. Veg. Sci. 2012, 15, 145–157. [Google Scholar] [CrossRef]
  223. Mutanga, O.; Adam, E.; Cho, M.A. High density biomass estimation for wetland vegetation using WorldView-2 imagery and random forest regression algorithm. Int. J. Appl. Earth Obs. Geoinf. 2012, 18, 399–406. [Google Scholar] [CrossRef]
  224. Lehnert, L.W.; Meyer, H.; Meyer, N.; Reudenbach, C.; Bendix, J. A hyperspectral indicator system for rangeland degradation on the Tibetan Plateau: A case study towards spaceborne monitoring. Ecol. Indic. 2014, 39, 54–64. [Google Scholar] [CrossRef]
  225. Ali, I.; Cawkwell, F.; Dwyer, E.; Barrett, B.; Green, S. Satellite remote sensing of grasslands: From observation to management. J. Plant Ecol. 2016, 9, 649–671. [Google Scholar] [CrossRef]
  226. 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]
Figure 1. Global distribution of pastures and assessment using remote sensing tools. “Global assessment of land degradation and improvement 1. Identification by remote sensing” (Bai et al., 2008 [15]). Global grassland classification was embellished to the original map by the authors.
Figure 1. Global distribution of pastures and assessment using remote sensing tools. “Global assessment of land degradation and improvement 1. Identification by remote sensing” (Bai et al., 2008 [15]). Global grassland classification was embellished to the original map by the authors.
Remotesensing 15 04866 g001
Figure 2. Number of studies from each country and across continents reviewed. Note: the blue and orange colours represent the ratio of the number of studies (blue) compared to the total number of studies (orange).
Figure 2. Number of studies from each country and across continents reviewed. Note: the blue and orange colours represent the ratio of the number of studies (blue) compared to the total number of studies (orange).
Remotesensing 15 04866 g002
Figure 3. Temporal (annual) pattern of studies reviewed by their topics of coverage. Bars indicate the number of studies published each year.
Figure 3. Temporal (annual) pattern of studies reviewed by their topics of coverage. Bars indicate the number of studies published each year.
Remotesensing 15 04866 g003
Figure 4. (a) The two main drivers of pasture variability, climate, and anthropogenic (b) studies using remote sensing to understand how adaptive pasture management could be used to mitigate climate change. Rainfall and temperature variables are regarded as weather and climate data.
Figure 4. (a) The two main drivers of pasture variability, climate, and anthropogenic (b) studies using remote sensing to understand how adaptive pasture management could be used to mitigate climate change. Rainfall and temperature variables are regarded as weather and climate data.
Remotesensing 15 04866 g004aRemotesensing 15 04866 g004b
Figure 5. (a) Remote sensing instruments to study pasture conditions. (b) Number of studies according to how instruments were combined for investigation. OO = combination of optical instruments, OR= combination of optical and radar instruments, and UAS = unmanned aerial systems.
Figure 5. (a) Remote sensing instruments to study pasture conditions. (b) Number of studies according to how instruments were combined for investigation. OO = combination of optical instruments, OR= combination of optical and radar instruments, and UAS = unmanned aerial systems.
Remotesensing 15 04866 g005
Figure 6. Number of studies ranked by the topics covered: OO = combinational of optical instruments, OR= optical and radar instruments; and UAS = unmanned aerial systems.
Figure 6. Number of studies ranked by the topics covered: OO = combinational of optical instruments, OR= optical and radar instruments; and UAS = unmanned aerial systems.
Remotesensing 15 04866 g006
Figure 7. Summary of the scale of focus enabled by satellite sensors. Note: NS = “Not specified” for studies that do not provide a definite statement about the scale of coverage in the reviewed studies. Studies (i.e., 63) indicated study locations without providing details about the scale of focus [76,152,153,154].
Figure 7. Summary of the scale of focus enabled by satellite sensors. Note: NS = “Not specified” for studies that do not provide a definite statement about the scale of coverage in the reviewed studies. Studies (i.e., 63) indicated study locations without providing details about the scale of focus [76,152,153,154].
Remotesensing 15 04866 g007
Figure 8. (a) A high PlanetScope imagery quantifying pasture biomass variation at paddock level (image acquired from Planet Lab Inc.; and accessed on 6 April 2021); (b) was georeferenced from (a). Landholders can make management decisions based on pasture availability. (b) was georeferenced using the map features provided (i.e., Ngahinapouri, Waipa District, Waikato, 3882, New Zealand).
Figure 8. (a) A high PlanetScope imagery quantifying pasture biomass variation at paddock level (image acquired from Planet Lab Inc.; and accessed on 6 April 2021); (b) was georeferenced from (a). Landholders can make management decisions based on pasture availability. (b) was georeferenced using the map features provided (i.e., Ngahinapouri, Waipa District, Waikato, 3882, New Zealand).
Remotesensing 15 04866 g008
Figure 9. The frequency with which pastures were monitored via satellite imagery passes.
Figure 9. The frequency with which pastures were monitored via satellite imagery passes.
Remotesensing 15 04866 g009
Table 1. Search phrases used to refine papers reviewed.
Table 1. Search phrases used to refine papers reviewed.
Search CategoriesSearch Strings/Synonyms/Terms
Pasture Management traitsquality, fertilizer, manure, irrigation, nutrient, management, “soil condition”, “water”, “mowing”
Pasture Production quantity, height*, sward, biomass, production, productivity, yield*, growth, “growth rate”
Pasture Composition species, botanical*, classification,
Pasture Degradation decline, “grazing intensity”, “grazing pressure”, “overgrazing”, “carrying capacity”, “stocking rate”, “stocking density”, “land use”, “fractional cover”
VegetationGrassland*, rangeland*, pasture *, graz*
Remote Sensing“Earth observation”, UAS or UAV, drone, satellite*,
Remote Sensing Adoption“end-user* ”, adoption*, technology
Table 7. Summary of current remote sensing information and forms of adoption by end users.
Table 7. Summary of current remote sensing information and forms of adoption by end users.
Remote Sensing Data Main FocusEnd user/sCountry of AdoptionEconomic CostYearInferenceReference
Perspective article: (satellite)Pasture degradationGovernment, pastoralistAustralia and ChinaNil2020Researchers should partner with end users.[27]
Perspective article (satellite)Pasture biomass determination.FarmersNew ZealandNil2020Value proposition defines how farmers would adopt satellite data.[216]
UAS (Phantom)Pasture biomass/herbage utilisation.Researchers, rangers, farmersUSA$15002019Cloud-based remote sensing utilisation where spatial resolution counts.[193]
Perspective article (satellite)Pasture management focusing on precision agriculture.FarmersUnited Kingdom and IrelandNil2019Improvement in pasture quality through management (nutrients).[215]
NDVI derived from MODISPasture quality. FarmersAltai Mountain (Russia, Mongolia, China and Kazakhstan).Free2019Integrate farmers’ ground-based pasture management with satellite data.[217]
MODIS derived Enhanced vegetation index (EVI).Grassland classification.Policymakers and farmersChinaFree2018To manage the carrying capacity of sheep.[194]
Satellite imagery (Landsat)FORAGE system estimator.The general public (emphasis on range managers)Australia Free2018A web-based system prepared by the Queensland state government, Australia.[214]
Above Net Primary Production from NDVI derived from MODIS. (Satellite data and GIS).Forage productivity to manage stocking rate and the carrying capacity.Policy makers and farmersArgentinaFree2007Monthly monitoring tool within the selected farms. [8]
NDVI derived from Landsat imageryIncreased pasture productivity by eliminating noxious weeds. Pasture conservation. Farmers and range managersUSAFree2006An online password-protected decision support tool[218]
Landsat imagery and GIS system.Land cover classification and pasture management.Government, range manager.ChinaNil2004Expert system toward database inventory. [219]
ERS satellite dataEstimating pasture biomassPolicymakers and national agencyBoliviaNil2003Research was initiated to validate and support a national framework.[175]
Landsat and SPOT imagery and GIS system.Data to support pasture management framework. FarmersMongoliaSatellite imagery came with a cost.1999 [220]
Landsat and SPOT imagery and GIS system.Pasture growth and productivity through fertilizer application.Researchers, research institution (CSIRO) and Agric company.AustraliaSatellite imagery was provided through a license. 1996Research was conducted through a vendor. [87]
Table 8. Global agencies providing satellite imagery to enable grasslands monitoring on demand.
Table 8. Global agencies providing satellite imagery to enable grasslands monitoring on demand.
Name of AgencyData SourceData Archive
United States Geological Survey (USGS)Landsat, MODIS, Sentinel-2, and othershttps://earthexplorer.usgs.gov
Sen2AgriR&D on Sentinel-2 datahttp://due.esrin.esa.int/page_users.php
National Aeronautics and Space Administrative (NASA)MODIS, VIIRS, SMAP (data on vegetation dynamics)https://www.earthdata.nasa.gov
European Space Agency (ESA)Sentinel satellites (Sentinel-2 and Sentinel-1 for vegetation monitoring)https://scihub.copernicus.eu/dhus
National Oceanic and Atmospheric Administration (NOAA)AVHRRhttps://www.avl.class.noaa.gov
Food and Agriculture Organization (FAO)Geospatial datasets in agriculture and vegetationhttps://data.apps.fao.org/map/catalog/srv/eng/catalog.search#/home
Digital Earth Africa (DEA)Sentinel-2, Landsat, Sentinel-1 and othershttps://www.digitalearthafrica.org/
Digital Earth Australia (DEA)Sentinel-2, Landsat, Sentinel-1 and othershttps://www.dea.ga.gov.au/about/open-data-cube
Sentinel HubCloud API for satellite imageryhttps://www.sentinel-hub.com/
Google Earth Engine (GEE)Cloud API for most satellite imagery archive https://developers.google.com/earth-engine/datasets
Launch RAP (rangeland analysis platform)Landsat (rangeland monitor for the USA)https://rangelands.app/
FORAGELandsat (rangeland monitor for Queensland)https://www.longpaddock.qld.gov.au/forage/
Linear Imaging Self-scanning sensor-3 (LISS-3)Indian satellites (IRS-1C, IRS-1D and Resourcesat-2) for vegetation monitoringhttps://www.isro.gov.in/
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

Ogungbuyi, M.G.; Mohammed, C.; Ara, I.; Fischer, A.M.; Harrison, M.T. Advancing Skyborne Technologies and High-Resolution Satellites for Pasture Monitoring and Improved Management: A Review. Remote Sens. 2023, 15, 4866. https://doi.org/10.3390/rs15194866

AMA Style

Ogungbuyi MG, Mohammed C, Ara I, Fischer AM, Harrison MT. Advancing Skyborne Technologies and High-Resolution Satellites for Pasture Monitoring and Improved Management: A Review. Remote Sensing. 2023; 15(19):4866. https://doi.org/10.3390/rs15194866

Chicago/Turabian Style

Ogungbuyi, Michael Gbenga, Caroline Mohammed, Iffat Ara, Andrew M. Fischer, and Matthew Tom Harrison. 2023. "Advancing Skyborne Technologies and High-Resolution Satellites for Pasture Monitoring and Improved Management: A Review" Remote Sensing 15, no. 19: 4866. https://doi.org/10.3390/rs15194866

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