Next Article in Journal
A Novel Method for SAR Ship Detection Based on Eigensubspace Projection
Previous Article in Journal
Lidar-Based Aboveground Biomass Estimations for the Maya Archaeological Site of Yaxnohcah, Campeche, Mexico
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Evapotranspiration Acquired with Remote Sensing Thermal-Based Algorithms: A State-of-the-Art Review

by
Vicente García-Santos
1,*,
Juan Manuel Sánchez
2 and
Joan Cuxart
3
1
Department of Earth Physics and Thermodynamic, University of Valencia, 46100 Burjassot, Spain
2
Department of Applied Physics, University of Castilla-La Mancha, 02071 Albacete, Spain
3
Department of Physics, University of the Balearic Islands, 07122 Palma, Spain
*
Author to whom correspondence should be addressed.
Remote Sens. 2022, 14(14), 3440; https://0-doi-org.brum.beds.ac.uk/10.3390/rs14143440
Submission received: 17 May 2022 / Revised: 7 July 2022 / Accepted: 9 July 2022 / Published: 18 July 2022

Abstract

:
Almost fifty years have passed since the idea to retrieve a value for Evapotranspiration (ET) using remote sensing techniques was first considered. Numerous ET models have been proposed, validated and improved along these five decades, as the satellites and sensors onboard were enhanced. This study reviews most of the efforts in the progress towards providing a trustworthy value of ET by means of thermal remote sensing data. It starts with an in-depth reflection of the surface energy balance concept and of each of its terms, followed by the description of the approaches taken by remote sensing models to estimate ET from it in the last thirty years. This work also includes a chronological review of the modifications suggested by several researchers, as well as representative validations studies of such ET models. Present limitations of ET estimated with remote sensors onboard orbiting satellites, as well as at surface level, are raised. Current trends to face such limitations and a future perspective of the discipline are also exposed, for the reader’s inspiration.

1. Introduction

Evapotranspiration (ET) is a process that includes evaporation from the surface, such as open water bodies, soil and vegetation, and transpiration as the water released mostly by the plant leaves transported from the root system. ET is a key factor in the hydrological cycle, since it describes the mechanism and energy needed to transport the liquid water stored in the soil-watershed-canopy system to the atmosphere, converted into water vapour.
In the broader context, irrigated agriculture still accounts for 70% of freshwater withdrawals [1]. This is a serious problem noticed by the Horizon Europe program. In this framework, ET plays an important role as a controlling factor of the water cycle and energy transport among the biosphere, atmosphere and hydrosphere. The importance of ET has been evidenced from the very beginning of the Remote Sensing discipline, in application fields such as hydrology, for instance for the scheduling of field-scale irrigations and tillage [2] or in meteorology and climatology, for prediction of natural hazards such as floods and droughts [3] or in agriculture, where for vegetated surfaces, ET can be used as an indicator of plant water stress [4].
In situ ET field retrievals from ground instrumentation, based on conventional techniques (e.g., Eddy covariance, Bowen ratio, lysimetry, etc.), offer the advantage of a direct determination of the value of ET for a measuring site [5]. However, these instruments are often costly, their data require an adequate treatment which is quite time-consuming, and instrumental problems may happen [6]. A commonly-used alternative for the estimation of ET is the Penman-Monteith equation, which uses monitored variables (e.g., air temperature or humidity), very convenient with basic meteorological stations, especially in the form suggested by Food and Agriculture Organization (FAO) [7,8]. Furthermore, for an area which is spatially heterogeneous, measurements made in a single location may not be representative of the full area [9]. For a given area in a humid climate, annual ET can be as large as half of the precipitation, whereas for arid and semi-arid areas ET is almost equivalent to the total annual rainfall [10].
In summary, even if accurate ET estimates can be obtained for the instrumented sites, these values are not useful at larger scales due to the usual heterogeneity of the landscape and the complexity of the hydrological processes [2]. Conventional ground measurement techniques do not appreciate spatial distribution at large-scales, especially in regions composed of well-contrasted areas where advection may be important.
Remote sensing technology is a globally consistent and economically feasible means to estimate ET values at regional and meso-scales on the Earth’s surface [11], since the approach directly links surface radiances and the components of the surface energy balance [12,13,14]. Over the last 30 years, combined use of satellite remote sensing data from optical and thermal infrared sensors has provided substantial progress in the estimation of ET [15,16,17,18]. Based on the concept of surface energy balance and net radiation, most remote sensing models have estimated ET for application studies such as water consumption, water resources planning and management over watersheds [19] or modeling ecological processes and analyzing biophysical characteristics of landscape [20].
The present document is a review report of the most popular remote sensing Land surface temperature (LST)-based ET algorithms used during the last three decades. This article revisits and updates of the state of the art on this topic reported in previous review articles [17,21,22,23,24,25], incorporating recently published works showing modifications and improvements of the models. This review aims to combine the current state of the art on the subject under study, making it suitable for both experts and users of ET estimated from satellite. Nevertheless, reading the aforementioned review articles is highly recommended for a more complete and up-to-date idea of the subject under study.
A brief description of the surface energy budget is provided in Section 2, while some of the most popular remote-sensing approaches to estimate ET are described in Section 3, indicating their evolution with time and the representative validation studies in the last twenty years. Also advantages and disadvantages of each ET model are discussed. Section 4, it is shown an in-depth review of the limitations and current trends in remote sensing techniques to estimate ET, as well as the future prospects to be addressed by the remote sensing ET community. Finally, Section 5 resumes the main conclusions of this review study.

2. Surface Energy Balance (SEB) Theory

It is a common approach to consider that an energy balance equation can be applied to any point of the Earth Surface [26]. The Net Radiation is the flux term that warms the surface during the day and cools it at night.
Instead ET is a process related with the quantity of energy used by the surface for the water phase change. When condensation occurs at the surface (usually at night) energy is transferred to the surface. Ignoring at this point any other effects (e.g., advection, heat stored in the upper soil or the plant energetics), the surface energy balance equation can be written as:
Rn = G + H + LE
where the net radiation Rn (in W/m2), represents the total heat energy considered in the balance between incoming and outgoing short-wave (0.15 to 3 μm) and long-wave (3 to 100 μm) radiation under steady atmospheric condition, and it can be expressed as:
Rn = (1 − αs) Rs + εs εaσTa4 − εsσTs4
where αs is the surface shortwave albedo, calculated with satellite data as a combination of narrow spectral reflectance estimated from bands of several remote sensing instruments [27], however, it can also be estimated in other ways such as using a 4-component net radiometer. Rs is commonly called insolation and it is a measure of the incident solar radiation, either photosynthetically active radiation in the visible spectrum (400–700 nm) or in the total shortwave (150–3000 nm), received on a given surface area and time. There also exist several remote sensing products to estimate Rs. For instance, the geostationary satellite Geostationary Operational Environmental Satellite (GOES) [28] or polar satellites like the Moderate Resolution Imaging Spectroradiometer (MODIS) [29] or Advanced Very High Resolution Radiometer (AVHRR) [30]. Atmospheric emissivity (εa) can be estimated from vapor pressure values [31]. There are several ways to retrieve air temperature (Ta). Some of them are the remote sensing MODIS products MOD07 [32], or from synthetic atmospheric profiles [33] included in radiative transfer codes, which has been proven to be more accurate than MOD07 compared with meteorological data from a near station [34]. Also Ta can be retrieved from other datasets like tower-based measurements or reanalysis data such as European Centre for Medium-Range Weather Forecasts (ECMWF, https://www.ecmwf.int/en/forecasts/dataset/ecmwf-reanalysis-v5 (accessed on 1 July 2022). Surface emissivity (εs) and LST can be retrieved from remote sensing techniques separately [35,36,37] as well as at the same time [38,39]. σ is the Stefan-Boltzmann constant. Two last terms in Equation (2) can be substituted by downward and upward longwave radiation terms. Both longwave radiation variables can be estimated by remote sensing techniques. For the downward term there exist operational products from MODIS and GOES [40], AVHRR and Spinning Enhanced Visible and InfraRed Imager (SEVIRI) [30], while for the upward radiation term there are products and algorithms for MODIS, Advanced Spaceborne Thermal Emission and Reflection (ASTER), Thematic Mapper (TM), SEVIRI, AVHRR, Along Track Scanning Radiometer (ATSR), Advanced Along-Track Scanning Radiometer (AATSR), GOES, Geostationary Meteorological Satellite (GMS), Visible Infrared Imaging Radiometer Suite (VIIRS), and Atmospheric infrared sounder (AIRS) [41]. Reader is referred to review study of [42] for further details about Rn estimated from ground measurements to remote sensing tools.
The soil heat flux G (in W/m2), is the heat energy used for warming or cooling substrate soil volume. As of today, it is impossible to directly measure G from remote sensing data. However, in many studies G is retrieved as a pondered estimate of Rn [43], or related to other variables easily retrievable from remote sensing data, such as Leaf Area Index (LAI) [44], Normalized Difference Vegetation Index and LST (NDVI) [45], and Fraction Vegetation Cover (FVC) [46]. Reader is referred to review study of Gao et al. [47] where this issue is investigated and a variety of methods are intercompared.
The sensible heat flux H (in W/m2), is the driving mechanism through which the air above the surface warms or cool as heat is transported from or to the surface. In the SEB Equation (1), it is expressed as:
H = ρ C P T o T a r a H
where ρ is the density of air (kg/m3), CP is the specific heat of air (1004 J/kgK), Ta is the air temperature usually at a height z of 2 m above the canopy [48], To is the aerodynamic temperature near the surface and raH is the aerodynamic resistance (s/m) obtained from the Monin-Obukhov Stability Theory (MOST) [49]. To is a parameter not easily estimable, so usually in remote sensing it is considered as the radiometric Ts retrieved from TIR sensors on board orbiting satellites, after some reformulation of the raH parameter to account for such approximation [43]. This issue is addressed in the next Section 2.2.
The Latent Heat Flux LE (in W/m2), governs the loss of latent heat from surface due to evapotranspiration. It is usually retrieved as a residual from the SEB Equation (1) and it is linearly related with ET:
L E = λ ET
where λ is the latent heat of vaporization of water, which is dependent on the temperature, and usually takes a value of 2.45 × 106 J/kg at 20 °C.
The remote sensing LST-based ET models produce instantaneous (ETi) values, for which the algorithms will be described in the next section. This instantaneous estimation is of not much use for most hydrological and water resources management applications and it is customary to convert ETi retrieved from overpass times of satellites to daily (ETd) or longer time scales to make full usable application of the remote sensing data. Several authors have proposed approximation methods to derive ETd values from ETi estimated from satellite data. Some of these well-accepted methods are described below.
Jackson et al. [50] related ETi to ETd value according to the diurnal solar irradiance evolution, following the sine function equation:
ET d = 2 N π sin π t N ET i
N = 0.945 n 1 + n 2 sin 2 π D + 10 365
n 1 = 12 5.69 · 10 2 l a t 2.02 · 10 4 l a t 2 + 8.25 · 10 6 l a t 3 3.15 · 10 7 l a t 4
n 2 = 0.123   l a t 3.1 · 10 4 l a t 2 + 8 · 10 7 l a t 3 + 4.99 · 10 7 l a t 4
where t is the time of day, D day of year and lat is the geographical latitude between 60°S and 60°N. For cloud free or a relatively constant cloud cover days, the sine function (5) has produced reliable estimates of ETd [11,51].
Another method assumes that evaporative fraction (EF) remains constant during the daylight hours [52]. EF is the rate between ETi and available instantaneous energy (Rn−G)i. However, it has been found that during nighttime ET could reach as many as about 10 percent of the daily totals [53]. So, the night induces a percentage that is added to EF, and after that assumes EF constant during the 24 h of the day [52,53]. Then ETd can therefore be written as:
ET d = 1.1 R n d R n G i ET i
where Rnd is the daytime available energy, actually it is (Rn−G)d but it is generally assumed that daily G is negligible (see Section 2.2).
Allen et al. [54] considered that EF was not able to capture the impact of advection, changing wind and humidity conditions along the day. So, they defined the constant reference ET fraction (ETrF) as the ratio between ETi and reference ET over the standardized 0.5 m tall alfalfa (ETr) [7]. For horizontal areas (no significant slope terrain effects) Allen et al. proposed the next ETd approximation:
ET d = E T r F E T r = ET i E T r E T r , d
where ETr,d is the cumulative daily reference of ET [54].

2.1. Aerodynamic Resistance to Heat Transfer Based on MOST

Aerodynamic resistance to heat transfer (raH in Equation (3)), is an ecosystem parameter which is affected by the surface roughness (vegetation height and structure), wind speed (u) and atmospheric stability. Increasing the intensity of the wind or the roughness of the surface imply a decrease in the value of raH [48]. It must be adjusted according to surface characteristics and based on the calculation of momentum flux and the observed logarithmic profile of wind speed close to surfaces, it is given by:
r a H = l n z u d z 0 M ψ M l n z T d z 0 H ψ H κ 2 u
where zu and zT are the heights at which u and Ta were respectively measured, d is the zero-displacement height, z0M and z0H are the roughness length of momentum and sensible heat exchange, respectively. κ = 0.41 is the dimensionless von Karman constant, ΨM and ΨH are the Monin-Obukhov stability functions [49] for momentum and sensible heat turbulent fluxes. Usually, two levels of temperature and wind are required to compute these functions, typically at 2 m above the canopy height for the atmospheric variables, while wind is set to zero at z0M and the value of LST is taken at z0H.
Atmospheric stability correction on raH is an important issue since not considering it leads to an overestimation of ET when the canopy-air temperature difference is greater than about ±2 °C [48]. To determine the stability functions the Obukhov length [49] is defined as:
Λ = u * 3 ρ C P T a g κ H
where u* is the friction velocity and g the gravity acceleration. It must be solved with H and u* [55]. When Λ < 0, means atmospheric instability, and for Λ > 0 stable conditions. It is worth to note that both Ta and u* are external parameters which retrieval is non-dependent on remote sensing techniques [56,57,58,59,60].
In case of atmospheric thermal instability and no predominant free convection conditions (usually at daytime), ΨM and ΨH can be expressed as [61]:
ψ M = 2 l n 1 + x 2 + l n 1 + x 2 2 2 a r c t a n x + π 2
ψ H = 2 l n 1 + x 2 2
x = 1 16 z u d Λ 0.25
Under atmospheric stable conditions (prevailing at night) [62,63] proposed the same formulation for the Monin-Obukhov stability functions, as:
ψ M = ψ H = 5 z u d Λ
The effective roughness length for momentum (z0M) is a distance above the zero-plane displacement height where the wind speed is assumed to be zero when log-profile wind speed is extrapolated downward, rather than at true ground surface [64]. z0M and d are complex functions of the surface roughness (vegetation height and structure), estimated from models that consider canopy height, width, and element spacing into account [65,66]. There are also alternative methods that integrate the LAI in the estimation of z0M and d [67].
The roughness length for heat exchange (z0H) can be expressed as a function of z0M as [67]:
l n z 0 M z 0 H = κ B
where B the Stanton number (dimensionless). κ/B rate is a complex function, also called “excess resistance”, which is dependent on time of the day, weather, and the type of vegetation, but constant values for z0H/z0M of 0.1 and 0.2 are often used for dense and sparse crops, respectively [68].
In summary, surface roughness plays an important role in the estimation of H and it is worth noting it still remains a challenging issue for large scale retrieval of the turbulent heat fluxes. However, resistance Equation (11) is applied since early seventies over local and regional scales and for varied vegetation covers [48,69,70]. We will avoid further details on MOST at this point not to distract the reader from the main aim of this review. For more information about surface roughness formulation, reader is referred to reviews of [71,72] and an in-depth study by [73].

2.2. Satellite Estimation of To

Regarding To (Equation (3)), it represents the temperature of the apparent source/sink of sensible heat flux [74], and it is computed at the level of d + zoH, averaged from the weighted contribution to the aerodynamic conductance of each element composing the full canopy scene. According to [75] To is neither measured nor easily estimated, while Ts is relatively easily measured with infrared thermometers and scanners. To has been usually substituted by Ts in Equation (3) since it is very difficult to measure [69,76]. Min et al. [77] indicated that for homogeneous and isothermal surfaces the definition of aerodynamic and thermodynamic (canopy or surface radiometric) temperatures are equivalent, however, assuming To equal to Ts at canopy level could be problematic since it is known that Ts is 2–3 °C higher than To for uniform canopy covers [78], and up to 10–15 °C higher for incomplete canopy covers [79]. Ts is also influenced by the viewing angle, so if it is measured at nadir Ts tends to be larger than To due to the direct incident sunlight. Nevertheless, at larger viewing angles between 50° and 70° from nadir, Ts coincides with To [80]. Difference between Ts and To leads to errors in the estimation of H which in turn leads to errors in the estimation of LE (ET). To account for those differences, researchers commonly have parameterized, with mixed results, the H equation in order to be able to use Ts instead of To. However, there is an effect of vegetation three-dimensional (3D) structure and its interaction with directional wind speed (wind direction, angle of attack), on the generation of To, roughness length (for heat and momentum transfer), friction velocity; which in turn affect H and LE.
More recently, [81] improved a To model for deficit irrigated corn by including a turbulence mixing-row resistance that attempts to incorporate the interactions between crop rows orientation and wind direction (the angle of attack). However, the approach is only applicable to the corn canopy structure and site conditions. In terms of vegetation canopy structure effects on energy balance models, [82] found out that, in the case of the unique canopy structure and configuration of vineyards, these SEB models were not able to adequately describe the physical processes driving evapotranspiration from vineyards.
In general, research results reported in the literature show that the surface aerodynamic temperature can be locally parameterized for different vegetation surfaces, surface homogeneity and environmental and climatological conditions, with relative success. However, researchers still use surface radiometric temperature images, from satellites, instead of the appropriate aerodynamic temperature because currently there is no single robust and comprehensive To model, for the application of the one-source energy balance of land surfaces model, that works for multiple surface cover types and environmental conditions. Thus, there is a need to study the mechanisms that generate the surface aerodynamic temperature, for different vegetation/crop surface types and environmental conditions.

2.3. Energy Balance Closure Flux Terms Frequently Not Considered

The hypothesis that the four terms described above close the surface energy budget is usually not fulfilled [26], with imbalances amounting typically between 10 and 40% depending on the terrain heterogeneity and complexity [83]. The most relevant motions seem to take place at the hectometre scale [84,85].
The main reasons for this lack of closure usually mentioned are the uncertainty of the measurements, the horizontal transport by advection, the unadequate computation of heat stored in the upper soil or the canopy and the lack of consideration of the energy used in the plant physiological processes for dense canopies. Several studies using drone or satellite-derived LST fields show that the advection of heat may be important [86,87,88] and well correlated with the imbalance.
Forcing the closure of the four-term surface budget equation to obtain LE, implies that the imbalance is considered entirely as latent heat flux, for which there is no clear evidence, although some studies point in that direction [89]. Instead most authors tend to distribute the observed imbalance between the latent and the sensible heat fluxes proportionally to the observed Bowen ratio (H/LE) [90]. In the rest of this work, we will consider that the surface energy budget is closed unless stated otherwise.

3. Remote Sensing LST-Based ET Algorithms

ET algorithms based on thermal remote sensing data can be divided into three categories: (i) single-source models, (ii) VI-TS Triangle/trapezoidal models, and (iii) dual (two)-source models.

3.1. Single-Source Energy Balance Models

3.1.1. Surface Energy Balance Index (SEBI) and Simplified-SEBI (S-SEBI) Models

Originally proposed by [91], the physically-based assumption is that ET varies with the Ts for a homogeneous surface. For a given surface albedo and boundary layer conditions, there exists a temperature gradient between the surface and the atmosphere (Ts−Ta) at some crop height, that ranges from a maximum value for which ET is assumed to be zero, to a minimum value for which the potential ET (ETp) is taken, as given by the Penman-Monteith equation [7].
Figure 1 shows a complete scheme of the SEBI principles. There exists a correlation between Ts and the surface reflectance (albedo), where for low reflectance values Ts remains constant with the progressive increase of such reflectance. This phenomenon is attributed to presence of body water or well-watered surfaces, where all the available energy is used for the evaporation. From certain reflectance threshold Ts increases with the reflectance, inducing a decrease in the evaporation, in the so-called “evaporation controlled” range. From another higher reflectance threshold, Ts decreases with the increase of the reflectance, because the Rn decreases due to the high reflection of the energy received, a process known as “radiation controlled”.
Based on the previous assumptions, for a specific surface given Ts, [91] provided a value for ET by means of an interpolation expression:
ET i = 1 T s T p b l r a H T s , m i n T p b l r a H , m i n T s , m a x T p b l r a H , m a x T s , m i n T p b l r a H , m i n ET p = 1 S E B I ET p
where Tpbl is the average air potential temperature at higher elevation or at the top of the Planetary Boundary Layer (pbl) and raH,min and raHmax are the minimum and maximum aerodynamic resistance in the sensible heat flux H [49].
A simplified version of the SEBI method (S-SEBI) was proposed by [92] with the advantage that no additional meteorological data are needed if the hydrological conditions of the surface are extreme enough to observe a significant range of gradient temperatures for the corresponding reflectance spectra.
Therefore if for a given αs value the two extremes of Ts exist, that is for Hmax (Ts,max, αs) and LEmax (Ts,min, αs), [92] expressed the ET value as:
λ ET i = T s , m a x T s T s , m a x T s , m i n R n G
The method is called simplified if the wet and dry pixel conditions are observed for a specific albedo value in a satellite scene. But [92] simplified even more Equation (19) by adjusting linearly the trends of maximum and minimum Ts with the albedo (see Figure 1), according the following equations:
T s , m a x = a m a x + b m a x α s
T s , m i n = a m i n + b m i n α s
where amax and bmax are the slope and offset coefficients for the linear regression of the maximum temperature values with the albedo and amin and bmin are the corresponding coefficients for the linear regression of the minimum temperatures, also with the albedo. These four parameters must be adjusted in each individual satellite scene. Including Equations (20) and (21) in Equation (19), ET expression from S-SEBI model can be defined then as:
λ ET i = a m a x + b m a x α s T s a m a x a m i n + b m a x b m i n α s R n G
Very few modifications have been proposed in the literature for the S-SEBI model. One of them is made by [93], which in order to avoid using the LE and H fluxes, it is proposed using instead the evaporative fraction (EF), that can be extracted from the Ts-α scatterplot shape.
A recently published study [94] proposed an adaptation of EVapotranspiration Assessment from SPAce (EVASPA) tool [95] to S-SEBI model. This model, denoted as E3S (for EVASPA S-SEBI Sahel), proposes two relevant changes. The first one is related to the handling of the algorithms determining the wet and dry edges, since in some regions like the Sahel with strong seasonal contrast only one edge may be identified. This led to the introduction of twelve additional seasonally-dependent algorithms to estimate the dry and wet points, using an ensemble member weighting. Secondly a cloud edge filtering method was implemented, to avoid that pixel outliers can distort the Ts scatterplot shape. The cloud edge filtering was developed in two levels: (1) the first level filter was based on the quality flag provided by the MODIS LST products. Thus, all pixels bordering clouds and associated with an LST error >1 K were removed; (2) a second level filter consisted in extending the first level filter to cloud-bordering pixels with LST below the first LST quartile in the entire image. This quantile level selection could be site and climate-specific.
Table 1 shows a review of the validation studies carried out from the S-SEBI proposal up to date. The validation studies are listed chronologically, indicating the sensor used and the spatial resolution. The uncertainties for each study are indicated in terms of Root Mean Square Error (RMSE), or bias. The corresponding type of surface (i.e., crop, forest, plateau, etc.) is also included in Table 1. The main advantage of the S-SEBI method is that no ground-based measurements are required. Instead, all the four input parameters (αs, Ts, Rn and G) can be retrieved from satellite data (G is related with Rn, see Section 2.2). Nevertheless, a disadvantage is that extreme temperatures are mandatory and these are site-specific. Daily ET values for the S-SEBI ETi, are estimated in each study of Table 1, using Equations (5)–(10).
It is worth to note, that for Table 1, Table 2, Table 3, Table 4, Table 5, Table 6 and Table 7, sometimes the spatial resolutions for the Landsat series is 30 m. This is because authors are referring to the original scale of the VNIR sensor, not the original resolution of the thermal InfraRed (TIR) sensor of Landsat, which is 60 m (L7), 120 m (L5) and 100 m (L8−L9). The Landsat team offers to the user the TIR products of its sensors at 30 m resolution, by complementing the data from its Visible and Near InfraRed (VNIR) sensors.

3.1.2. Surface Energy Balance System (SEBS) Model

Proposed by [46] the SEBS method is based, as it is S-SEBI, SEBAL and METRIC (to be described below), on the same main concept as SEBI. The difference of SEBS from the rest is that it deals with estimating the ET fluxes at large scales, therefore considering heterogeneous surfaces.
Therefore, the assumption of constant atmospheric conditions for the whole scene is not applicable. Moreover, spatial variation of roughness height for different surfaces is allowed.
The SEBS is an extended model that defines the surface roughness length in the heat fluxes, reformulating the EF at the limiting cases. To this end in the SEBS theory the dry and wet case limits are:
Dry limit:
LE d r y = 0
H d r y = R n G
Wet limit:
LE w e t = R n G H w e t
H w e t = R n G ρ C P V P D r a H γ 1 Δ γ
where VPD is the vapor pressure deficit (difference between saturated and non-saturated vapor pressure), Δ is the slope of the curve relating temperature difference (Ts-Ta) with VPD, and γ is the is the psychrometric constant.
As an intermediate step SEBS method defines a relative EF (EFr) as:
E F r = 1 H H w e t H d r y H w e t
Once all the required parameters have been estimated, LE is expressed in SEBS as:
L E = E F r · LE w e t
Some studies presented a modification of the SEBS inner structure in terms of: (i) momentum roughness length and zero plane displacement; (ii) an empirical relationship between land surface temperature and air temperature; and (iii) designing a new input interface of meteorological field at reference, with the objective of representing land surface and climatic changes of a specific region. These studies adapted SEBS to be applicable in different countries like Taiwan (SEBS-Taiwan, [102]), China (SEBS-China, [103]) or Iran (SEBS-Iran, [104]). SEBS was also modified at local scale (SEBS-Urban, [105]) after including the anthropogenic heat flux in the SEB equation to estimate the ET of Beijing city. This study hypothesized that ET was equal to the water consumption in the study area.
An improvement of SEBS method was suggested by [106] by correcting the excess resistance (Equation (15)) with a scale factor dependent on soil moisture (SM), following:
κ B c o r r = a + 1 1 + e x p b c S M r e l κ B
where a, b and c are coefficients of the sigmoid function and SMrel is a relative soil moisture defined as:
S M r e l = 1 S M S M m i n S M m a x S M m i n
SMmax and SMmin values can be obtained using a time series analysis of the soil moisture data on annual or long-term basis [106]. Wu et al. [107] suggested the use of the Normal Differential Water Index (NDWI) instead of SMrel in Equation (30). Yi et al. [108] also recommended replacing SMrel by the inverse of the modified perpendicular drought index (MPDI).
It is also worth mentioning the study of Chen et al. [109], in which the excess resistance is internally modified to be robust for heat flux calculation for high and low canopies, since it has been proved its depends on environmental conditions, surface types, canopy structure and vegetation surface geometry. kB−1 structure was adjusted from a one-foliage layer to a multi-foliage layer, which provides the possibility of including the impact of vertical variations in foliage leaf area density, foliage shelter factor, and foliage heat transfer coefficient.
Table 2 shows a review of the validation studies carried out for the SEBS model up to date. The main advantages of the SEBS are that new limit cases in the SEB reduce uncertainty from Ts and meteorological variables, SEBS introduces a new formulation for the roughness length for heat transfer instead of using fixed values, and a priori knowledge of the actual turbulent heat fluxes is not required. However, the disadvantages are that SEBS requires too many parameters and offers relatively complex solution for the turbulent heat fluxes, which can be a source of inconveniences when data are not available.

3.1.3. Surface Energy Balance Algorithm for Land (SEBAL) Model

Proposed by [120], SEBAL is an iterative and feedback-based numerical procedure that deduces the radiation, heat and evaporation fluxes. SEBAL estimates LE with the residual of Equation (1) in a pixel by pixel satellite scene, just considering two anchor points (dry and wet points) for the full scene. At the dry point LE = 0 and H = Rn-G, and at the wet point H = 0 and LE can be considered either Rn-G or a reference ET [7].
The main consideration of SEBAL is to establish a linear regression between Ts and difference of Ts with the air temperature at the crop height z (δTa), as follows:
δ T a = c 1 T s c 0
where c1 and c0 are the empirical coefficients derived from anchored conditions of wet (δTa,wet = 0, Ts,wet) and dry points (Ts,dry), for which the temperature difference is expressed as:
δ T a , d r y = ( R n G ) r a H ρ C P
These coefficients must be estimated for each individual satellite scene, independently of the surface type composing the image. Once the regressed equation, for each Ts value assigned to each pixel of the scene is retrieved, the H is estimated with the expression:
H ( T s ) = ρ C P r a H c 1 T s c 0
A first modification of SEBAL was proposed by [121] after introducing Digital elevation model (DEM) data to account for impacts of slope aspect on incident solar radiation. But the main subject of modification of SEBAL model is related with the choice of endmembers or anchor points (i.e., “hot” and “wet” points in the satellite scene). A Modified version of SEBAL (M-SEBAL) was proposed by [122], that based the election of the anchor (dry/wet) points on the type of surface, defined by the FVC. They adjusted Equation (31) for every specific surface defined by FVC and subsequently estimated H and LE. Feng et al. [123] derived geo-mathematically four reference dry and wet limits, instead of the classical option of two anchor points.
Another modification is addressed in [124], where authors deal with the endmember pixel selection process, proposing an automated procedure by introducing an exhaustive search algorithm (ESA), in which all possible candidates are systematically considered and checked to determine if the solution is satisfied, based on two steps: (1) create a binary map of candidate pixels that are identified using a simple rule-based classifier, and (2) apply ESA to identify hot and cold pixels that meet the defined criteria. In this sense, [125] created the automated SEBAL (ASEBAL) model that allows automatic selection of the endmember pixels according to criteria based on the variation in biophysical parameter values retrieved from satellite products.
It is also worth mentioning the study of [126], where authors found more appropriate to combine the capabilities of Synthetic Aperture Radar (SAR) with optical remote sensing technique in the applications where soil moisture conditions to locate anchor pixels, since SAR is more suitable to detect soil moisture conditions.
Two interesting studies that suggested modifications of SEBAL ET model were related with changes in the SEB equation, by introducing new fluxes terms that must be accounted in certain situations. Mkhwanazi et al. [127] proposed a modified version of SEBAL model (SEBAL-A), that is capable of accounting for advection flux term, estimated with a wind function which included a daily mean or extreme temperatures parametrization, which is incorporated it into the original SEBAL daily ET sub-model. [128] also suggested enhancing the SEBAL model when it is applied in a heterogeneous urban environment (uSEBAL), after considering urban land surface parameters and including the urban anthropogenic heat flux in the SEB budget.
Finally, it is worth to note that recent open access software code versions of SEBAL are accessible. One of them is a new python version of SEBAL (pySEBAL), which incorporates an automated pixel selection procedure and is currently under development and testing at the IHE-Delft Institute [129]. In addition, a new tool based on the SEBAL algorithm, modified with a simplified version of the Calibration using Inverse Modeling at Extreme Conditions process [130] for the endmembers selection, was developed within the Google Earth Engine (geeSEBAL) environment [131].
Table 3 shows a review of the validation studies carried out by the SEBAL model up to date. The main advantages of SEBAL are the minimum ground data required, the unnecessary atmospheric correction thanks to the internal calibration. However, election of dry/wet anchor points is subjective, SEBAL applications over non-flat areas and uncertainties on Ts estimates have a great impact on H retrievals.

3.1.4. Mapping Evapotranspiration at High Resolution and with Internalized Calibration (METRIC) Model

METRIC was proposed by [54], based almost completely on SEBAL method. The main differences with SEBAL are that Rn and Ts are corrected topographically using a Digital Elevation Model. Both methods differ in the calculation of H as in METRIC the temperature gradient of the wet point (Equation (32)) is expressed as:
δ T a , d r y = ( R n G 1.05 E T r ) r a H ρ C P
The assumption that ET rate (ETr) is about 5% above that of the standard ETr in the METRIC model is assumed because, for a large population of fields, some fields will have a wet soil surface beneath a full vegetation canopy that will tend to increase the total ET rate [54].
Since METRIC is directly based on SEBAL model, all the modifications previously discussed in Section 3.1.3 can also apply to METRIC. However, it is interesting to note a recent study [138], where calibration variability in METRIC is examined through the anchor pixels selection. Following METRIC original procedure to detect “cold” and “hot” pixel in a scene, the authors detected a range of different “cold” and “hot” pixels that conformed a total of five sets of calibration pixel pairs (based on combination of maximum and minimum “cold” and “hot”, as well as values closest to the flux station). So, different slopes and intercept values were retrieved from the plot δT vs. Ts (Equation (31)). In summary, authors observed significant spatial differences in ET across the study area, which suggested that candidate calibration pixels cannot be completely quantified using only NDVI and Ts. Therefore, it can be presumed that the large differences in δT among the candidate pixels were not random noise, and that the inclusion of available energy into the calibration pixel selection process could potentially improve the calibration process and provide better estimates of ET.
Is also worth to note that METRIC results can be directly obtained at any region of the world with the new Earth Engine Evapotranspiration Flux (EEFlux) portal [139,140]. Table 4 shows a review of the validation studies carried out for the METRIC model up to date. The main advantages and drawbacks of METRIC are the same as for SEBAL, Highlighting that the main advantage over SEBAL is its application over mountainous areas.

3.2. Land Surface Temperatures–Vegetation Index Triangle/Trapezoidal ET Models

Vegetation Indices (VIs) offer important information from Ts−VI triangle/trapezoidal relationship. These diagrams can be useful for estimating regional ET values under the condition that full ranges of SM content and vegetation are present in the remote sensing scene.
The Triangle method was described by [146], and provides a biophysical justification for the combination of T and vegetation indices (VI) time series for land-cover mapping and land-cover change analysis. Figure 2a shows the simplified Ts−VI triangle diagram where the so called “dry edge” is defined by a decrease of the Ts with the increase of the VI, this edge is defined under the assumption that highest Ts for different types of surfaces (different VIs) is associated to points with unavailability of SM content. The lower envelope represents the wet edge, which is nearly constant and represents the lowest Ts for the different types of surfaces, assuming here that the ET is potential. Theoretically the two envelopes (dry and wet edges) intersect in a point of full cover vegetation and maximum ET.
Based on the Priestley-Taylor formulation [147] and fully remotely sensed data, an expression proposed by [148] allows estimating the LE according to the Ts−VI triangular relationship as:
L E = φ R n G Δ Δ + γ
where φ varies from 0 to 1.26, and it is estimated with a two-step linear interpolation. This interpolation consists in: (1) once set a φmin = 0 (driest bare soil) and φmax = 1.26 (largest VI and lowest Ts), for each intermediate VI a φmin,i is estimated by interpolation between φmin and φmax and a φmax,i is assigned to the minimum Ts value of such VI (usually 1.26), (2) Different φi can be estimated from the Ts range at each VI. VI can be either the NDVI, FVC, the Soil Adjusted Vegetation Index (SAVI), Enhanced Vegetation Index (EVI), etc.
The Ts−VI Trapezoidal method based on the proposed Water Deficit Index (WDI) by [70], an index related to the ratio of actual and potential evapotranspiration, which evaluates the evapotranspiration rates of both full-cover and partially vegetated sites. WDI based Ts−VI Trapezoidal method considers that full covered vegetation can both water-stressed and well-watered in a satellite scene, so there is no intersect in a point of full cover vegetation and the Ts−VI diagram become a trapezoid (Figure 2b). Moran et al. [70], extended the ET estimation from full to partially vegetated surface areas, considering that temperature gradient (TsTa) varies linearly with the vegetation status. So, WDI is linearly related at each VI to the maximum (dry bare soil) and minimum (probably well-watered full-cover vegetation) gradient temperature, taking values between 0 and 1 and LE is calculated similar to Equation (19) as the ratio of potential ET calculated from Penman-Monteith equation [7], LE for the trapezoidal method is expressed as:
LE = 1 W D I L E P = 1 T s T a m i n T s T a i T s T a m i n T s T a m a x L E P
Stinsen et al. [149], combined the triangle method with thermal inertia information obtained from MSG-SEVIRI sensor, by changing the surface temperature over time. Authors also applied a slightly modified version of the two-step interpolation scheme to obtain φ, based on a interpolation where the decomposition is regarded as non-linear with the argument that non-linear intersections are observed in the dry edge and in the iso-lines of equal moisture availability of the Ts−VI tringle feature. Adopting equations from this study, [150] proposed a Time-Domain Triangle Method (TDTM), based on a pixel by pixel Ts–VI feature space, obtained by exploring the temporal domains over each single image pixel, since the triangle contains the whole climatic variability of the Ts–VI. Also, in [151], an algorithm to determine quantitatively the dry and wet edges for the Ts–VI triangular space in arid and semi-arid areas, where wet pixels are not generally easily identified, was developed with an iterative process that automatically retrieve the two edges in the Ts–FVC triangle space in ten steps. The process starts dividing the range of FVC in the Ts–FVC triangle space into M intervals evenly and ends with a linear regression to obtain the dry edge.
Another interesting modification of the Ts−VI triangle method is proposed by [152] that estimate daily ET directly using the top of atmosphere (TOA) radiances without performing atmospheric correction or other processes, to avoid the uncertainty associated with the estimation of ET, attributed to satellite data products. Furthermore, a recent study [153] estimates the soil heat flux (G) in the trapezoidal space of Ts-FVC, prior to apply the original Ts−VI trapezoidal method afterwards.
Table 5 shows a review of the validation studies carried out by the Ts−VI Triangular/Trapezoidal methods up to date. The main advantage of Triangular method is that no atmospheric or ground data are required, but the drawback is that large numbers of pixels with wide range of SM and FVC are required to assure dry and wet limits, a subjective procedure. For the Trapezoidal method one advantage over triangular version is that a large number of pixels is not required since intermediate values are determined by four limiting vertices. But a drawback is that ground measured data (vapor pressure, Ta, u, stomatal resistances, etc.) are required. A major disadvantage for both Triangular/Trapezoidal methods is that they do not detect immediately the water stress, since it does not have instantaneous impact on vegetation cover.

3.3. Dual (Two)-Source Energy Balance (TSEB) Model

A two-layer model of turbulent exchange that includes the view geometry associated with directional radiometric surface temperature was developed and evaluated by [43]. The main assumption of TSEB model is that Ts estimated from remote sensing measurements is the composite of soil and canopy surfaces, and it can be expressed as:
T s = F V C T c 4 + 1 F V C T s o i l 4
where Tc is the temperature of the canopy and Tsoil is the temperature of the non-vegetated soil. Note the term FVC in this Equation (37) is a function of LAI and sensor viewing angle [44]. The original TSEB does not consider the weighting by the emissivities of each component, but some authors have claimed for its importance [132,159,160].
The key of the TSEB model is the partition of sensible heat flux into the canopy and soil layers. If we assume that there is an interaction between the fluxes of both components, due to a heating of the in-canopy air by heat transport from the soil, the resistance framework in TSEB can be considered to be in series [43,161,162,163]. A parallel configuration was also introduced [43], based on the assumption that the resistance of each source interacts independently with respect to the surface-atmosphere, although the canopy still has an effect on the radiation transmitted towards the soil and the wind attenuation through the canopy to the soil surface. Schemes in Figure 3 shows both the series and parallel configurations [164]. Each component has its own SEB equation expressed as:
R n , s o i l = H s o i l + L E s o i l + G
R n , c = H c + L E c
A variety of radiative transfer models have been implemented within TSEB to estimate Rn components. Traditional versions of TSEB estimate Rn using the physically-based radiative transfer model described in [164], first implemented within TSEB by [165]:
In this revised two-source model, the separation of net radiation between canopy and soil can be expressed as:
R n c = L n c + 1 τ s 1 α c S
R n s = L n s + τ s 1 α s S
where S is solar radiation, τ s is solar transmittance in the canopy, α s is soil albedo, α c is canopy albedo, LnS and LnC are longwave radiation for soil and canopy, respectively, and are estimated via the following expression:
L n s = exp ( κ L Ω L A I ) L s k y + 1 exp ( κ L Ω L A I ) L C L s
L n c = 1 exp ( κ L Ω L A I ) L s k y + L s 2 L c
where LSky, LS, and LC are longwave radiation from the sky, soil, and canopy, and can be calculated from air, soil, and canopy temperature, Ω is the clumping factor, and κL is an extinction coefficient. Campbell and Norman [164] described how to estimate the transmittance, albedo, and extinction coefficients for soil and canopy.
Soil and canopy sensible heat fluxes H of Equations (38) and (39), can be deduced easily from Figure 3 as:
H s o i l = ρ C P T s o i l T a r a H + r s
H c = ρ C P T c T a r a H
where the parameter rs is the resistance to the heat flow in the boundary layer immediately above the soil surface, and it can be expressed as:
r s = 1 0.0025 T s T c 1 / 3 + 0.012 u s
where us is the wind speed at a height above the soil surface where the effect of soil surface roughness is minimal (i.e., 0.05–0.1 m).
Latent heat fluxes for soil (LEs) and canopy (LEc) are calculated as residuals of Equations (44) and (45).
Another type of two-source model formulation is the so-called patch model where it is assumed that all the fluxes act vertically and that there is no interaction between soil and canopy components (i.e., a complete energy balance between the atmosphere and each element [165,166]).
There exist modifications on the TSEB model, like the Two-Source Time-Integrated Model (TSTIM) proposed by [53], lately called Atmosphere-Land Exchange Inverse (ALEXI) in [167], which relates time-differential Ts observed (1.5–5.5 h after sunrise) from geostationary satellites to the time-integrated energy balance within the surface-atmospheric boundary layer system.
ALEXI has minimal reliance on absolute (instantaneous) air or surface temperature input data, and therefore provides a relatively robust flux determination at the coarse geostationary pixel scale (5–10 km). For finer scale ET applications, ALEXI flux fields can be spatially disaggregated using higher resolution Ts information from polar orbiting systems (e.g., Landsat or MODIS) in an algorithm referred to Disaggregated ALEXI (DisALEXI) model [168].
Another modified TSEB model is the Simplified TSEB (STSEB) model proposed by [169], which suggests dividing the Rn, H and LE fluxes in soil and canopy weighted values, according to the FVC. The scheme proposed by [170], which is based on canopy conductance by means of a formulation that has sound parameterizations of transpiration and evaporation, that links soil surface temperature and soil evaporation eliminating the need of using soil moisture as input, and allowing the model predicting surface temperature and energy fluxes simultaneously. Furthermore, it can be calibrated using remotely sensed LST data.
Table 6 and Table 7 show a list of the validation studies carried out using TSEB model up to date. TSEB is obviously a very well-accepted model, based on the amount of validation studies. The main advantages of TSEB model are: angular effects are into consideration, no “resistance excess” correction is required, atmospheric and emissivity induced errors on Ts are reduced (case of ALEXI/DisALEXI model).
In general terms, the validation studies of all the ET models shown in Table 1, Table 2, Table 3, Table 4, Table 5, Table 6 and Table 7 establish the estimation error of the daily ET between 1–5 mm/d, with an average error of around 1 mm/d regardless of the sensor spatial resolution or the type of surface under study. However, it can be seen for ET models such as S-SEBI, SEBS, LST-VI Triangle/Trapezoidal or Two-Source, that at high spatial resolutions (60–100 m) the error is smaller (around 1 mm/d), and around 2 mm/d for kilometric resolutions. Regarding the one-source ET SEBAL and METRIC models, the validation studies have shown the smallest errors, with values ranging between 0.4–2.4 mm/d, although most of these studies have been carried out with high spatial resolution sensors. Finally, it cannot be stated that the error for all models has decreased over the years.
This review has shown, in line with conclusions from the recent Carlson’s essay [183], that mathematical models of natural land-living systems applied to satellite-based measurement systems, will never be able of full-describing and predicting all relevant processes related to the functioning of such systems, insomuch as these models also require knowledge of plants and soils. However, such models can be useful, in describing key driving variables under limited environment conditions for plant growth and development, and surface energy balance at spatial resolutions ranging from tens of meters, up to hundreds of kilometers.

4. Limitations, Current Trends and Future Prospects

4.1. Toward High Spatiotemporal Resolution ET Retrievals

Water resource management applications require spatial information about daily and seasonal ET at pixel scales resolving individual agricultural fields or even finer management zones within fields. Such applications include irrigation scheduling, monitoring compliance with water rights, negotiating water rights transfers, sales or leases, predicting changes in regional water use due to land-use and/or climate changes, and studying the impacts of agriculture on surrounding ecosystems and stream flow [173]. Ts is the most influential input parameter in deriving the energy fluxes, because it provides an indication of the net effect of land-atmosphere interactions, which results in the process of evapotranspiration. However, coarser-scale TIR imaging devices (e.g., MODIS) can provide Ts data for ET mapping on a near-daily basis for relatively large fields, these data being too coarse to resolve water use at the scale of individual users in most irrigation districts. On the other hand, some polar orbiting satellites (e.g., Landsat series, ASTER) are the only platforms that provide routine, global thermal imagery at scales (60–120 m) that resolve water use patterns over heterogeneous agricultural areas, but overpass frequency (16 days) is not optimal for ET monitoring. Hence, a desirable alternative technique is to derive high resolution Ts data using the fine resolution VIS/NIR/TIR bands that are available from either the same or different satellite systems, at daily frequencies of temporal resolution.
A strong inverse relationship exists between Ts and VIs (see Section 3.2.), and VIs can be derived from VIS/NIR data at high spatial resolution (60–120 m). Such a relationship can be exploited to improve the spatial resolution of TIR band imagery (at a coarser resolution of 1–10 km) than the one associated to VIS/NIR band instruments. This relationship reflects the fact that denser vegetation cover tends to be correlated with lower surface temperatures due to cooling by transpiration.
Generally, most methods first imply a coincident overpass of two (high and coarse) spatial resolution satellite imager (e.g., MODIS-Landsat/ASTER), both operating in the VIS/NIR/TIR regions. Second, the methods are based on a statistical regression between VI (e.g., NDVI, SAVI, EVI, Modified SAVI (MSAVI), Normalized Difference Moisture Index (NDMI), NDWI), etc.), which is aggregated to the original resolution of the Ts (e.g., Landsat NDVI at 30 m is aggregated to 960 m of the original MODIS Ts) and the actual Ts. The resulting regression is used to disaggregate the coarse resolution Ts into fine resolution Ts with the VI retrieved for other dates. The fundamental assumption in this method of disaggregation is that the Ts–VI relationship remains continual at multiple spatial resolutions.
Techniques and tools to disaggregate either Ts, VI or ET are varied. So, the pioneering work of [184] used the Spatial and Temporal Adaptive Reflectance Fusion Model (STARFM) [185], which uses comparisons of one or more pairs of observed Landsat/MODIS maps, collected on the same day, to predict maps at Landsat-scale on other MODIS observation dates.
To estimate from ALEXI/DisALEXI application, daily ET values at MODIS (1 km) and Landsat (30 m) scales, with continuous data retrieved from geostationary GOES sensor. Such STARFM- ALEXI/DisALEXI combined algorithm has been proven successful in successive studies [173,175,176]. Other techniques like the hybrid use of the hot edge model and artificial neural network [186], machine learning approaches [187] or data mining [188] are also valuable disaggregation alternatives. However, temperature variations due to subpixel moisture variability cannot be recovered unless they are well-correlated with vegetation features. Therefore, dominant moisture variations must be well-resolved at the native TIR resolution. These results suggest that while thermal sharpening is a valuable tool for enhancing spatial information content in TIR imagery, it does not replace the need for TIR data collection at the sub-field scale. Other differences between both sensors exist due to several effects such as acquisitions at different time, spectral resolution, atmospheric correction, viewing angle, and pixel footprint that have to be considered and resolved, usually by normalization procedures.
An actual prospect which has been recently addressed in [189,190,191,192,193], is the synergistic application of disaggregation techniques to retrieve ET values at high spatiotemporal resolution, using two different satellites, one of which is carrying a sensor without thermal bands at coarser (~ 1 km) spatial resolution, and the other carrying a VIS-NIR sensor (i.e., MODIS- Spot-5), mapping Earth surfaces at finer spatial resolution (~10–30 m) and with a revisit of few days (e.g., 5 days for Sentiel-2, combining platforms A and B).
In this context, the ESA has funded the Sentinels for Evapotranspiration (Sen-ET) project, which main objective was to develop an open-source software application for accurate, pre-operational modelling of instantaneous evapotranspiration, by means of an optimal methodology for estimating ET at tens of meters spatial scale, based on synergistic use of Sentinel 2 and Sentinel 3 satellites observations. Also, the continuing Increasing Crop Water Use Efficiency at Multiple Scales Using Sentinel Evapotranspiration (ET4FAO) project, will implement further refinements to the Sen-ET method. For further information readers are referred to Guzinski et al. [194].

4.2. Accurate and Spatial Representative Field Instrumentation

Common ET measuring systems include water balance, lysimeters, Bowen ratio, eddy covariance, or sap flow [5]. Inherent to all these methods is the reality that the instrumental uncertainties and the difficulties to set a fully consistent experimental setup to determine the energy and water budgets can lead to water use estimates far from reality.
Instrumentations like water balance, lysimeters, Bowen ratio or sap flow are point specific and could not represent the ET of medium (100 m) to coarse (1–10 km) scales estimated from satellite data. The Eddy-Covariance systems are typically representative of an area surrounding them of a few hundred meters of radius [195], but in the case of terrain heterogeneities of smaller size their representativity is questionable, as local circulations at the hectometer may advect significant amounts of water vapour [196]. New Large Aperture Scintillometers (LAS) devices provide an approach to validate ET at much larger scales. LAS estimate the sensible flux based on measurements of small fluctuations in the refractive index of air caused by temperature, humidity, and pressure induced variations in density. However, according to [6] all the above-mentioned instruments showed average typical errors of: 10% (lysimetry), 15% (Bowen Ratio), 20% (soil water balance), 25% (Eddy covariance and LAS) and 35% (sap flow).
Eddy covariance and LAS are well-accepted techniques to validate ET estimates at medium to coarse (100–1000 m) and large (up to 10 km) scales, because of a good ratio error/advantage. But accurate instrumentation is required for ET validations at different scales. In this sense Laser LAS devices [26,197] are lately being used to ensure residual ET estimations (through accurate H estimations) at high and medium scales (10–100 m), reducing the problem of the SEB non-closure induced by the Eddy correlation [198].
The use of unmanned aerial vehicles (UAVs) has given a boost to precision agriculture, offering an opportunity to provide multi-spectral imagery at the highest spatial resolution (few cm per pixel, [87]). UAVs are portable and cost-effective, and flight procedures over crop fields helps increasing the production. New VIS/NIR/TIR sensors onboard UAVs have the ability of observing the agricultural fields with great detail. UAV images can produce ortho-mosaics and 3D information products to be used in mapping structure and volume of plant canopy, which is directly correlated to plant biomass and LAI, [199]. A significant number of studies recently published [200,201,202,203,204,205,206] have taken advantage of the good performance of UAVs, applying remote sensing techniques to estimates the crop ET. It must be emphasized the care to be taken, when applying ET models to thermal sensors on board UAVs, since it is well-known that uncooled commercial thermal cameras induce very significant errors in LST, which is key in the application of energy balance models [87].

4.3. Operational ET Products

Historically, the MODIS evapotranspiration products (MO/YD16) are the first 0.5 km operational land surface ET database for the global vegetated land surface at 8-day, monthly, and annual intervals [207]. This product is based on the Penman-Monteith approach [7], not on thermal data. This non-TIR based product has been widely evaluated against ground measurements, showing errors around 0.31 mm/d, or 24%, using data from 46 eddy covariance flux towers [208]. Other important non-thermal-based datasets products are the FLUXCOM [209], WaPOR [210], or the European Centre for Medium-Range Weather Forecasts (ECMWF) (GLDAS) ET images [211]. However, some initiatives assimilate the land surface temperature in their ET products. This is the case of the ECOSTRESS ALEXI/JPL ET [212,213], or the EUropean Organisation for the Exploitation of METeorological SATellites (EUMETSAT) Satellite Application Facility on Land Surface Analysis (LSA-SAF), ET product from Meteosat Second Generation (MSG) [214]. The comparison of these operational ET products with 15 eddy covariance flux towers located in Europe showed the good performance of the EUMETSAT over the MODIS ET [208,215,216,217] products [194].
The Global Land Evaporation Amsterdam Model (GLEAM) [218] is the only global ET operational product designed to be driven by remote sensing observations derived from microwave sensors, including soil moisture and vegetation optical depth. To assimilate microwave soil moisture observations—typically sensitive to the first few cm of the soil—GLEAM uses a simple Newtonian nudging algorithm, which minimizes the computational demands [219]. All output datasets have a daily resolution and a common 0.25° global grid.
As previously commented on Section 3.1.4., METRIC ET products can also be directly downloaded, for any region of the world, with the new Earth Engine Evapotranspiration Flux (EEFlux) portal [140,220], that includes a tool that processes ET maps from input images of Landsat series 7 and 8, i.e., ET maps at the spatial resolution of 30 m every 16 days. Initiatives like online platform (OpenET, [221]) for mapping daily, monthly and yearly ET values at the scale of individual fields (i.e., 30 m). This webtool offers easily accessible satellite-based ET across the western US. Also missions like GRAPEX [222] tries to apply a multi-scale remote sensing ET toolkit for mapping crop water use and crop stress for improved irrigation scheduling and water management in vineyards in the Central Valley of California, a region of endemic periodic drought.
Despite the enhancement of operational ET products, there is still a demand of higher resolution and more accurate operational ET products. As pointed out in Section 4.1., water resource management applications require spatial information about daily and seasonal ET at pixel scales of the hundreds of meters or lower. Operational high spatiotemporal ET products are being demanded specially from the agricultural sector and this is the goal to be achieved in the near future.

4.4. New Advent of Orbiting Sensors with Improved Spatial and Temporal Features

Since its beginning nearly four decades ago, the field of Remote Sensing has experienced an exponential increase in terms of satellites, products and users. However, the advance of the satellite engineering is always a compromise between high spatial/low temporal resolution (or vice versa) of the scenes acquired by the imager onboard. Fortunately, these advances tend to converge to instrument features that attend the actual requirements of remote sensing users. So, future Hyperspectral Infrared Imager (HyspIRI) [223], will combine near and shortwave infrared imaging spectrometer with a multi-channel TIR radiometer, designed to have a spatial resolution of 60 m and a revisit time of 5 days. Another future micro-satellite is the MIcro Satellite for Thermal Infrared GRound Surface Imaging (MISTIGRI) [224], which intends to estimate daily the TIR radiance from surface at resolutions of 50 m, also with a revisit period of 5 days. Currently is orbiting and acquiring data the ninth Landsat satellite [225], which is carrying, as well as its immediate predecessor Landsat 8, two sensors on-board, Operational Land Imager (OLI) detecting the VIS-NIR spectral reflectance at 30 m resolution and TIRS-2, measuring the spectral thermal emittance at 100 m. Landsat 9 revisit time is 16 days, like its predecessors.
In parallel, work is being carried out on the development of increasingly higher resolution multispectral thermal sensors, such as the ECOsystem Spaceborne Thermal Radiometer Experiment on Space Station (ECOSTRESS) mission [226], the Land Surface Temperature Monitoring (LSTM) mission [227] or the Thermal infraRed Imaging Satellite for High-resolution Natural resource Assessment (THRISNA) mission of Centre national d’études spatiales- Indian Space Research Organisation (CNES-ISRO) [228], which will allow further progress in the applications of this parameter in the estimation of crop water requirements, among other applications.
Recently, the U.S. Geological Survey- National Aeronautics and Space Administration (USGS-NASA) consortium has launched an operational LST product for the entire Landsat time series globally, which will also apply to the new TIRS-2 sensor on board Landsat 9, thus continuing the entire Landsat legacy beyond 2030.

4.5. SEB Closure and Impact of Advection Term

All the remote sensing-based ET models described in here, presume that the Equation (1) is undoubtedly true and the SEB closure takes place as Rn is used only as the source of the energy fluxes H, G and LE. The analysis of the surface energy budget with eddy-covariance measurements has made evident that this presumed closure seldom takes place, instrumentation error aside, and there are other terms that at least at some time of the day, it must be taken seriously into consideration. Cuxart et al. [83] indicated that the four mains traditionally terms would be only balanced if no other thermal sources or sinks existed, implying no thermal advection and surface homogeneity. Therefore, once surface heterogeneity comes into the picture, thermal advection will take place and the usual 4-term energy budget will not close anymore, and the corresponding imbalance is the sum of several other terms like the heat storage in the volume, the biological thermal exchanges, the temperature tendency, the thermal advection and any other unaccounted factor [88].
In a recent study [87] it was observed from high resolution (2 m) Ts maps obtained from a TIR camera assembled in a drone, that after up-scaling the original resolution to scales of 70–200 m, the advection term can explain an important amount of the imbalance occurred at midday in a SEB station. Non-closure of SEB equation is an important issue historically left aside, but new tools and remote sensing sensors commit to ET community to study such non-closure causes in more detail.

4.6. Machine Learning ET Retrievals

There is no doubt that the Ts is the most influential indicator of the land-atmosphere evapotranspiration. However, in the last years there have been a noticeable increase of published papers [229,230,231,232,233,234,235,236,237] that fused remote sensing techniques with a high variety of machine learning (ML) methods to estimate ET values. Artificial intelligence approaches using the ML techniques like: neural network (NN), support vector machine (SVM) or random forest (RF) have tried to consider the non-linearity between ET and the variables for meteorological and land surface conditions. However, these methods can be time-consuming since in order to use properly a machine learning algorithm, it is necessary to train it with a different dataset, and once optimized, the ML algorithm is checked against a test dataset, different from the training and validation datasets. A new review study about this topic [238] is recommended to the reader for more detail.

4.7. Merging ET Models

In the last five years the number of studies related with the ET estimations with multi-model techniques have raised. The reason of such increase is because hybrid ET models may address landscape heterogeneity occurring in most ecosystems. These areas are partially covered with soil, trees, and grasses patches, so taking advantage of better performance of single source ET models in homogeneous areas and Two-source ET models in heterogeneous ones is of key importance for best ET retrievals. So, a new growing trend to work with three-source SEB models, to account for complexity of various contributions of different vegetation elements in the scene has been stated [239], results showed that three-source SEB models can be of great utility for irrigation management of patchy surface as vineyard. Some of these hybrid ET models are also based on machine learning techniques such SVM [240], also with the Bayesian model averaging (BMA) method, which is based on the basis of the posterior probability distribution of multiple candidate models [101,241]. BMA can not only predict more accurate values, but do so with smaller biases and variances, therefore this method usually outperforms individual candidate models. Finally, a recent study [119] converted the single-source SEBS into a dual-source model (called TS-SEBS) to partition ET into E and T for detailed studies of flux exchanges between soils and canopies.

5. Conclusions

Monitoring the evapotranspiration of a crop or vegetated surface, through remote sensing techniques, has been one of the goals of disciplines such as agriculture, hydrology or climatology and meteorology, for a variety of purposes such as optimizing the irrigation scheduling and the agronomic management, or for the prediction of natural hazards such as floods and droughts. Direct estimation is simply not possible, and ET is generally obtained as a residue from the energy balance equation, once the terms Rn, G and H have been estimated, assuming that other effects, such as advection, are negligible, which is often not the case. However, the original concept of using an aerodynamic temperature as a reference was substituted by the information provided by the surface radiometric temperature, more easily derived from satellite data. Although results in the literature show that the aerodynamic surface temperature can be parameterized locally for different vegetation surfaces, surface homogeneity, and environmental and climatological conditions, with relative success, the lack of a robust and comprehensive model for the application of this approach keeps most efforts focused on the use of satellite thermal images. This review article brings together and updates the main ET models, based on LST data obtained from satellite. These models fall into 3 categories: (i) one-source models, (ii) triangular/trapezoidal LST-VI relationship models, and (iii) two-source models. A comprehensive selection of validation studies for each particular model is included.
After updating the state of the art of the subject under study, no firm conclusions can be stated regarding the better performance of one model versus others. One can only extract partial findings such as the better suitability of the one-source approaches for full cover crops, whereas the two-source models can better represent ET in woody or row crops. In general terms, the validation studies of all the ET models shown in Table 1, Table 2, Table 3, Table 4, Table 5, Table 6 and Table 7 establish the estimation error of the daily ET ranging 1–5 mm/d, with a mean error of around 1 mm/d regardless of the sensor, the spatial resolution or the type of surface under study. However, it is noticed that for ET models such as S-SEBI, SEBS, LST-VI Triangle/Trapezoidal or Two-Source, the error is smaller (around 1 mm/d) at high spatial resolutions (60–100 m), and degrades up to 2 mm/d for kilometer resolutions. As for the one-source ET SEBAL and METRIC models, validation studies report the smallest errors, with values ranging from 0.4–2.4 mm/d, although most of these studies focus and are based on high spatial resolution sensors. One of the main findings in this review is that new formulations, approaches or model updates do not succeed in reducing the uncertainty in ET modeling over the years. For these reasons, decisions on the models to be implemented to derive operational ET products based on coming missions such as ESA′s LSTM or NASA′s SBG, should rather stand on reliability and robustness under a variety of environmental and surface conditions. Joint efforts initiatives such as the recent OpenET should prevail in this search. Some important challenges that the field of ET monitoring from satellite data have been identified, with special attention on the use of machine learning methodologies to improve ET estimates.

Author Contributions

Conceptualization, V.G.-S. and J.C.; investigation, V.G.-S.; resources, V.G.-S., J.C. and J.M.S.; writing—original draft preparation, V.G.-S.; writing—review and editing, J.C. and J.M.S.; supervision, J.C. and J.M.S.; project administration, V.G.-S., J.C. and J.M.S.; funding acquisition, V.G.-S., J.C. and J.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was part of the R&D projects RT12018-098963C31-R, PID2020-118797RB-I00 and PID2020-113498RB-C21 funded by MCIN/AEI/10.13039/501100011033 and PROMETEO-2021-016 project funded by the Generalitat Valenciana Government.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors want to show their enormous gratitude to all those people who selflessly shared their knowledge, publications and studies so kindly, when the authors needed it most.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. FAO. The State of the World’s Land and Water Resources for Food and Agriculture–Systems at Breaking Point (SOLAW 2021); Synthesis Report; FAO: Rome, Italy, 2021. [Google Scholar] [CrossRef]
  2. Idso, S.B.; Jackson, R.D.; Reginato, R.J. Estimating Evaporation: A Technique Adaptable to Remote Sensing. Science 1975, 189, 991–992. [Google Scholar] [CrossRef]
  3. Brutsaert, W. Catchment-Scale Evaporation and the Atmospheric Boundary Layer. Water Resour. Res. 1986, 22, 39S–45S. [Google Scholar] [CrossRef]
  4. Jackson, R.D.; Idso, S.B.; Reginato, R.J.; Pinter, P.J. Canopy Temperature as a Crop Water Stress Indicator. Water Resour. Res. 1981, 17, 1133–1138. [Google Scholar] [CrossRef]
  5. Cuxart, J.; Boone, A.A. Evapotranspiration over Land from a Boundary-Layer Meteorology Perspective. Bound. Layer Meteorol. 2020, 177, 427–459. [Google Scholar] [CrossRef]
  6. Allen, R.G.; Pereira, L.S.; Howell, T.A.; Jensen, M.E. Evapotranspiration Information Reporting: I. Factors Governing Measurement Accuracy. Agric. Water Manag. 2011, 98, 899–920. [Google Scholar] [CrossRef] [Green Version]
  7. Allen, R.G.; Pereira, L.S.; Raes, D.; Smith, M. Crop Evapotranspiration—Guidelines for Computing Crop Water Requirements—FAO Irrigation and Drainage Paper 56; FAO—Food and Agriculture Organization of the United Nations: Rome, Italy, 1998; ISBN 92-5-104219-5. [Google Scholar]
  8. Pereira, L.S.; Paredes, P.; Hunsaker, D.J.; López-Urrea, R.; Jovanovic, N. Updates and Advances to the FAO56 Crop Water Requirements Method. Agric. Water Manag. 2021, 248, 106697. [Google Scholar] [CrossRef]
  9. Zhang, L.; Lemeur, R. Evaluation of Daily Evapotranspiration Estimates from Instantaneous Measurements. Agric. For. Meteorol. 1995, 74, 139–154. [Google Scholar] [CrossRef]
  10. Engman, E.T.; Gurney, R.J. Remote Sensing in Hydrology; Chapman and Hall Ltd.: London, UK, 1991; ISBN 0412244500. [Google Scholar]
  11. Kustas, W.P.; Norman, J.M. Use of Remote Sensing for Evapotranspiration Monitoring over Land Surfaces. Hydrol. Sci. J. 1996, 41, 495–516. [Google Scholar] [CrossRef]
  12. Jackson, R.D. Evaluating Evapotranspiration at Local and Regional Scales. Proc. IEEE 1985, 73, 1086–1096. [Google Scholar] [CrossRef]
  13. Caselles, V.; Sobrino, J.A.; Coll, C. On the Use of Satellite Thermal Data for Determining Evapotranspiration in Partially Vegetated Areas. Int. J. Remote Sens. 2007, 13, 2669–2682. [Google Scholar] [CrossRef]
  14. McCabe, M.F.; Wood, E.F. Scale Influences on the Remote Estimation of Evapotranspiration Using Multiple Satellite Sensors. Remote Sens. Environ. 2006, 105, 271–285. [Google Scholar] [CrossRef]
  15. Seguin, B.; Itier, B. Using Midday Surface Temperature to Estimate Daily Evaporation from Satellite Thermal IR Data. Int. J. Remote Sens. 2007, 4, 371–383. [Google Scholar] [CrossRef]
  16. De Troch, F.P.; Troch, P.A.; Su, Z.; Lin, D.S. Application of Remote Sensing for Hydrological Modelling. In Distributed Hydrological Modelling; Springer: Dordrecht, The Netherlands, 1990; pp. 165–191. [Google Scholar] [CrossRef]
  17. Li, Z.L.; Tang, R.; Wan, Z.; Bi, Y.; Zhou, C.; Tang, B.; Yan, G.; Zhang, X. A Review of Current Methodologies for Regional Evapotranspiration Estimation from Remotely Sensed Data. Sensors 2009, 9, 3801–3853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Liou, Y.A.; Kar, S.K. Evapotranspiration Estimation with Remote Sensing and Various Surface Energy Balance Algorithms—A Review. Energies 2014, 7, 2821–2849. [Google Scholar] [CrossRef] [Green Version]
  19. Schultz, G.A.; Engman, E.T. Present Use and Future Perspectives of Remote Sensing in Hydrology and Water Management. In Remote Sensing and Hydrology 2000. Selected Papers from a Conference Held at Santa Fe, NM, USA, 2–7 April 2000; IAHS Press: Wallingford, UK, 2001; pp. 545–551. [Google Scholar]
  20. Quattrochi, D.A.; Luvall, J.C. Thermal Infrared Remote Sensing for Analysis of Landscape Ecological Processes: Methods and Applications. Landsc. Ecol. 1999, 14, 577–598. [Google Scholar] [CrossRef]
  21. Courault, D.; Seguin, B.; Olioso, A. Review on Estimation of Evapotranspiration from Remote Sensing Data: From Empirical to Numerical Modeling Approaches. Irrig. Drain. Syst. 2005, 19, 223–249. [Google Scholar] [CrossRef]
  22. Kalma, J.D.; Mcvicar, T.R.; Matthew, A.E.; Mccabe, F.; Kalma, J.D.; Mcvicar, T.R.; Mccabe, M.F. Estimating Land Surface Evaporation: A Review of Methods Using Remotely Sensed Surface Temperature Data. Surv. Geophys. 2008, 29, 421–469. [Google Scholar] [CrossRef]
  23. Kustas, W.; Anderson, M. Advances in Thermal Infrared Remote Sensing for Land Surface Modeling. Agric. For. Meteorol. 2009, 149, 2071–2081. [Google Scholar] [CrossRef]
  24. Wang, K.; Dickinson, R.E. A Review of Global Terrestrial Evapotranspiration: Observation, Modeling, Climatology, and Climatic Variability. Rev. Geophys. 2012, 50, 2005. [Google Scholar] [CrossRef]
  25. Zhang, K.; Kimball, J.S.; Running, S.W. A Review of Remote Sensing Based Actual Evapotranspiration Estimation. Wiley Interdiscip. Rev. Water 2016, 3, 834–853. [Google Scholar] [CrossRef]
  26. Mauder, M.; Foken, T.; Cuxart, J. Surface-Energy-Balance Closure over Land: A Review. Bound. Layer Meteorol. 2020, 177, 395–426. [Google Scholar] [CrossRef]
  27. Liang, S. Quantitative Remote Sensing of Land Surfaces; John Wiley & Sons: Hoboken, NJ, USA, 2005. [Google Scholar] [CrossRef]
  28. Zheng, T.; Liang, S.; Wang, K. Estimation of Incident Photosynthetically Active Radiation from GOES Visible Imagery. J. Appl. Meteorol. Climatol. 2008, 47, 853–868. [Google Scholar] [CrossRef]
  29. Liang, S.; Zheng, T.; Liu, R.; Fang, H.; Tsay, S.C.; Running, S. Estimation of Incident Photosynthetically Active Radiation from Moderate Resolution Imaging Spectrometer Data. J. Geophys. Res. Atmos. 2006, 111, D15. [Google Scholar] [CrossRef] [Green Version]
  30. Schulz, J.; Albert, P.; Behr, H.D.; Caprion, D.; Deneke, H.; Dewitte, S.; Dürr, B.; Fuchs, P.; Gratzki, A.; Hechler, P.; et al. Operational Climate Monitoring from Space: The EUMETSAT Satellite Application Facility on Climate Monitoring (CM-SAF). Atmos. Chem. Phys. 2009, 9, 1687–1709. [Google Scholar] [CrossRef] [Green Version]
  31. Brutsaert, W. On a Derivable Formula for Long-Wave Radiation from Clear Skies. Water Resour. Res. 1975, 11, 742–744. [Google Scholar] [CrossRef]
  32. Borbas, E.E.; Seemann, S.W.; Kern, A.; Moy, L.; Li, J.; Gumley, L.; Menzel, W.P. MODIS Atmospheric Profile Retrieval Algorithm Theoretical Basis Document, Collection 6; Cooperative Institute for Meteorological Satellite Studies: Madison, WI, USA, 2011. [Google Scholar]
  33. Barsi, J.A.; Schott, J.R.; Palluconi, F.D.; Hook, S.J. Validation of a Web-Based Atmospheric Correction Tool for Single Thermal Band Instruments. Earth Obs. Syst. X 2005, 5882, 58820E. [Google Scholar] [CrossRef]
  34. Pérez-Planells, L.; García-Santos, V.; Caselles, V. Comparing Different Profiles to Characterize the Atmosphere for Three MODIS TIR Bands. Atmos. Res. 2015, 161–162, 108–115. [Google Scholar] [CrossRef] [Green Version]
  35. Valor, E.; Caselles, V. Mapping Land Surface Emissivity from NDVI: Application to European, African, and South American Areas. Remote Sens. Environ. 1996, 57, 167–184. [Google Scholar] [CrossRef]
  36. Wan, Z. New Refinements and Validation of the Collection-6 MODIS Land-Surface Temperature/Emissivity Product. Remote Sens. Environ. 2014, 140, 36–45. [Google Scholar] [CrossRef]
  37. Hulley, G.; Hook, S. The ASTER Global Emissivity Database (ASTER GED); Jet Propulsion Laboratory, California Institute of Technology: Pasadena, CA, USA, 2015. [Google Scholar]
  38. Gillespie, A.; Rokugawa, S.; Matsunaga, T.; Steven Cothern, J.; Hook, S.; Kahle, A.B. A Temperature and Emissivity Separation Algorithm for Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) Images. IEEE Trans. Geosci. Remote Sens. 1998, 36, 1113–1126. [Google Scholar] [CrossRef]
  39. Hulley, G.; Malakar, N.; Freepartner, R. Moderate Resolution Imaging Spectroradiometer (MODIS) Land Surface Temperature and Emissivity Product (MxD21) Algorithm Theoretical Basis Document Collection-6; Jet Propulsion Laboratory, California Institute of Technology: Pasadena, CA, USA, 2016. [Google Scholar]
  40. Wang, W.; Liang, S. A Method for Estimating Clear-Sky Instantaneous Land-Surface Longwave Radiation with GOES Sounder and GOES-R ABI Data. IEEE Geosci. Remote Sens. Lett. 2010, 7, 708–712. [Google Scholar] [CrossRef]
  41. Liang, S. Recent Developments in Estimating Land Surface Biogeophysical Variables from Optical Remote Sensing. Prog. Phys. Geogr. Earth Environ. 2007, 31, 501–516. [Google Scholar] [CrossRef] [Green Version]
  42. Liang, S.; Liang, S.; Zhang, X.; Wang, K.; Zhang, X.; Wild, M. Review on Estimation of Land Surface Radiation and Energy Budgets from Ground Measurement, Remote Sensing and Model Simulations. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2010, 3, 225–240. [Google Scholar] [CrossRef]
  43. Norman, J.M.; Kustas, W.P.; Humes, K.S. Source Approach for Estimating Soil and Vegetation Energy Fluxes in Observations of Directional Radiometric Surface Temperature. Agric. For. Meteorol. 1995, 77, 263–293. [Google Scholar] [CrossRef]
  44. Gómez, M.; Olioso, A.; Sobrino, J.A.; Jacob, F. Retrieval of Evapotranspiration over the Alpilles/ReSeDA Experimental Site Using Airborne POLDER Sensor and a Thermal Camera. Remote Sens. Environ. 2005, 96, 399–408. [Google Scholar] [CrossRef]
  45. Chirouze, J.; Boulet, G.; Jarlan, L.; Fieuzal, R.; Rodriguez, J.C.; Ezzahar, J.; Er-Raki, S.; Bigeard, G.; Merlin, O.; Garatuza-Payan, J.; et al. Intercomparison of Four Remote-Sensing-Based Energy Balance Methods to Retrieve Surface Evapotranspiration and Water Stress of Irrigated Fields in Semi-Arid Climate. Hydrol. Earth Syst. Sci. 2014, 18, 1165–1188. [Google Scholar] [CrossRef] [Green Version]
  46. Su, Z. The Surface Energy Balance System (SEBS) for Estimation of Turbulent Heat Fluxes. Hydrol. Earth Syst. Sci. 2002, 6, 85–99. [Google Scholar] [CrossRef]
  47. Gao, Z.; Russell, E.S.; Missik, J.E.C.; Huang, M.; Chen, X.; Strickland, C.E.; Clayton, R.; Arntzen, E.; Ma, Y.; Liu, H. A Novel Approach to Evaluate Soil Heat Flux Calculation: An Analytical Review of Nine Methods. J. Geophys. Res. 2017, 122, 6934–6949. [Google Scholar] [CrossRef]
  48. Hatfield, J.L.; Perrier, A.; Jackson, R.D. Estimation of Evapotranspiration at One Time-of-Day Using Remotely Sensed Surface Temperatures. Agric. Water Manag. 1983, 7, 341–350. [Google Scholar] [CrossRef]
  49. Monin, A.S.; Obukhov, A.M. Basic Laws of Turbulent Mixing in the Surface Layer of the Atmosphere. Orig. Publ. Tr. Akad. Nauk SSSR Geophiz. Inst 1954, 24, 163–187. [Google Scholar]
  50. Jackson, R.D.; Hatfield, J.L.; Reginato, R.J.; Idso, S.B.; Pinter, P.J. Estimation of Daily Evapotranspiration from One Time-of-Day Measurements. Agric. Water Manag. 1983, 7, 351–362. [Google Scholar] [CrossRef]
  51. Colaizzi, P.D.; Evett, S.R.; Howell, T.A.; Tolk, J.A. Comparison of Five Models to Scale Daily Evapotranspiration from One-Time-of-Day Measurements. Trans. ASABE 2006, 49, 1409–1417. [Google Scholar] [CrossRef]
  52. Sugita, M.; Brutsaert, W. Daily Evaporation over a Region from Lower Boundary Layer Profiles Measured with Radiosondes. Water Resour. Res. 1991, 27, 747–752. [Google Scholar] [CrossRef]
  53. Anderson, M.C.; Norman, J.M.; Diak, G.R.; Kustas, W.P.; Mecikalski, J.R. A Two-Source Time-Integrated Model for Estimating Surface Fluxes Using Thermal Infrared Remote Sensing. Remote Sens. Environ. 1997, 60, 195–216. [Google Scholar] [CrossRef]
  54. Allen, R.G.; Tasumi, M.; Trezza, R. Satellite-Based Energy Balance for Mapping Evapotranspiration with Internalized Calibration (METRIC)-Model. J. Irrig. Drain. Eng. 2007, 133, 380–394. [Google Scholar] [CrossRef]
  55. Choudhury, B.J. Synergism of Multispectral Satellite Observations for Estimating Regional Land Surface Evaporation. Remote Sens. Environ. 1994, 49, 264–274. [Google Scholar] [CrossRef]
  56. Prihodko, L.; Goward, S.N. Estimation of Air Temperature from Remotely Sensed Surface Observations. Remote Sens. Environ. 1997, 60, 335–346. [Google Scholar] [CrossRef]
  57. Niclòs, R.; Valiente, J.A.; Barberà, M.J.; Caselles, V. Land Surface Air Temperature Retrieval From EOS-MODIS Images. IEEE Geosci. Remote Sens. Lett. 2014, 11, 1380–1384. [Google Scholar] [CrossRef] [Green Version]
  58. Liu, S.; Su, H.; Zhang, R.; Tian, J.; Wang, W. Estimating the Surface Air Temperature by Remote Sensing in Northwest China Using an Improved Advection-Energy Balance for Air Temperature Model. Remote Sens. Environ. 2016, 60, 335–346. [Google Scholar] [CrossRef] [Green Version]
  59. Liu, S.; Su, H.; Tian, J.; Zhang, R.; Wang, W.; Wu, Y. Evaluating Four Remote Sensing Methods for Estimating Surface Air Temperature on a Regional Scale. J. Appl. Meteorol. Climatol. 2017, 56, 803–814. [Google Scholar] [CrossRef]
  60. Sobrino, J.A.; Cheng, J.; Wang, C.; Bi, X.; Luan, Q.; Li, Z. Estimation of Daily and Instantaneous Near-Surface Air Temperature from MODIS Data Using Machine Learning Methods in the Jingjinji Area of China. Remote Sens. 2022, 14, 1916. [Google Scholar] [CrossRef]
  61. Paulson, C.A. The Mathematical Representation of Wind Speed and Temperature Profiles in the Unstable Atmospheric Surface Layer. J. Appl. Meteorol. Climatol. 1970, 9, 857–861. [Google Scholar] [CrossRef]
  62. Webb, E.K. Profile Relationships: The Log-Linear Range, and Extension to Strong Stability. Q. J. R. Meteorol. Soc. 1970, 96, 67–90. [Google Scholar] [CrossRef]
  63. Businger, J.A.; Wyngaard, J.C.; Izumi, Y.; Bradley, E.F. Bradley Flux-Profile Relationships in the Atmospheric Surface Layer. J. Atmos. Sci. 1971, 28, 181–189. [Google Scholar] [CrossRef]
  64. Carlson, T.N.; Dodd, J.K.; Benjamin, S.G.; Cooper, J.N. Cooper Satellite Estimation of the Surface Energy Balance, Moisture Availability and Thermal Inertia. J. Appl. Meteorol. Climatol. 1981, 20, 67–87. [Google Scholar] [CrossRef] [Green Version]
  65. Raupach, M.R. Drag and Drag Partition on Rough Surfaces. Bound. Layer Meteorol. 1992, 60, 375–395. [Google Scholar] [CrossRef]
  66. Raupach, M.R. Simplified Expressions for Vegetation Roughness Length and Zero-Plane Displacement as Functions of Canopy Height and Area Index. Bound. Layer Meteorol. 1994, 71, 211–216. [Google Scholar] [CrossRef]
  67. Colaizzi, P.D.; Evett, S.R.; Howell, T.A.; Tolk, J.A. Comparison of Aerodynamic and Radiometric Surface Temperature Using Precision Weighing Lysimeters. Remote Sens. Modeling Ecosyst. Sustain. 2004, 5544, 215–229. [Google Scholar] [CrossRef] [Green Version]
  68. Mölder, M.; Lindroth, A. Dependence of KB−1 Factor on Roughness Reynolds Number for Barley and Pasture. Agric. For. Meteorol. 2001, 106, 147–152. [Google Scholar] [CrossRef]
  69. Blad, B.L.; Rosenberg, N.J. Measurement of Crop Temperature by Leaf Thermocouple, Infrared Thermometry and Remotely Sensed Thermal Imagery 1. Agron. J. 1976, 68, 635–641. [Google Scholar] [CrossRef]
  70. Moran, M.S.; Clarke, T.R.; Inoue, Y.; Vidal, A. Estimating Crop Water Deficit Using the Relation between Surface-Air Temperature and Spectral Vegetation Index. Remote Sens. Environ. 1994, 49, 246–263. [Google Scholar] [CrossRef]
  71. Maes, W.H.; Steppe, K. Estimating Evapotranspiration and Drought Stress with Ground-Based Thermal Remote Sensing in Agriculture: A Review. J. Exp. Bot. 2012, 63, 4671–4712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Foken, T. 50 Years of the Monin–Obukhov Similarity Theory. Bound. Layer Meteorol. 2006, 119, 431–447. [Google Scholar] [CrossRef]
  73. Martí, B.; Martínez-Villagrasa, D.; Cuxart, J. Flux–Gradient Relationships Below 2 m Over a Flat Site in Complex Terrain. Bound. Layer Meteorol. 2022, 1–26. [Google Scholar] [CrossRef]
  74. Blonquist, J.M.; Norman, J.M.; Bugbee, B. Automated Measurement of Canopy Stomatal Conductance Based on Infrared Temperature. Agric. For. Meteorol. 2009, 149, 1931–1945. [Google Scholar] [CrossRef]
  75. Chehbouni, A.; lo Seen, D.; Njoku, E.G.; Monteny, B.M. Examination of the Difference between Radiative and Aerodynamic Surface Temperatures over Sparsely Vegetated Surfaces. Remote Sens. Environ. 1996, 58, 177–186. [Google Scholar] [CrossRef]
  76. Kalma, J.D.; Jupp, D.L.B. Estimating Evaporation from Pasture Using Infrared Thermometry: Evaluation of a One-Layer Resistance Model. Agric. For. Meteorol. 1990, 51, 223–246. [Google Scholar] [CrossRef]
  77. Min, W.; Chen, Z.; Sun, L.; Gao, W.; Luo, X.; Yang, T.; Pu, J.; Huang, G.; Yang, X.; Min, W.; et al. A Scheme for Pixel-Scale Aerodynamic Surface Temperature over Hilly Land. Adv. Atmos. Sci. 2004, 21, 125–131. [Google Scholar] [CrossRef]
  78. Chávez, J.L.; Gowda, P.H.; Howell, T.A.; Copeland, K.S. Radiometric Surface Temperature Calibration Effects on Satellite Based Evapotranspiration Estimation. Int. J. Remote Sens. 2009, 30, 2337–2354. [Google Scholar] [CrossRef]
  79. Chávez, J.L.; Howell, T.A.; Gowda, P.H.; Copeland, K.S.; Prueger, J.H. Surface Aerodynamic Temperature Modeling over Rainfed Cotton. Trans. ASABE 2010, 53, 759–767. [Google Scholar] [CrossRef] [Green Version]
  80. Matsushima, D.; Kondo, J. A Proper Method for Estimating Sensible Heat Flux above a Horizontal-Homogeneous Vegetation Canopy Using Radiometric Surface Observations. J. Appl. Meteorol. 1997, 36, 1696–1711. [Google Scholar] [CrossRef]
  81. Costa-Filho, E.; Chávez, J.L.; Comas, L. Determining Maize Water Stress through a Remote Sensing-Based Surface Energy Balance Approach. Irrig. Sci. 2020, 38, 501–518. [Google Scholar] [CrossRef]
  82. Alfieri, J.G.; Kustas, W.P.; Nieto, H.; Prueger, J.H.; Hipps, L.E.; McKee, L.G.; Gao, F.; Los, S. Influence of Wind Direction on the Surface Roughness of Vineyards. Irrig. Sci. 2019, 37, 359–373. [Google Scholar] [CrossRef]
  83. Cuxart, J.; Conangla, L.; Jiménez, M.A. Evaluation of the Surface Energy Budget Equation with Experimental Data and the ECMWF Model in the Ebro Valley. J. Geophys. Res. Atmos. 2015, 120, 1008–1022. [Google Scholar] [CrossRef] [Green Version]
  84. Cuxart, J.; Wrenger, B.; Martínez-Villagrasa, D.; Reuder, J.; Jonassen, M.O.; Jiménez, M.A.; Lothon, M.; Lohou, F.; Hartogensis, O.; Dünnermann, J.; et al. Estimation of the Advection Effects Induced by Surface Heterogeneities in the Surface Energy Budget. Atmos. Chem. Phys. 2016, 16, 9489–9504. [Google Scholar] [CrossRef] [Green Version]
  85. de Roo, F.; Mauder, M. The Influence of Idealized Surface Heterogeneity on Virtual Turbulent Flux Measurements. Atmos. Chem. Phys. 2018, 18, 5059–5074. [Google Scholar] [CrossRef] [Green Version]
  86. Twine, T.E.; Kustas, W.P.; Norman, J.M.; Cook, D.R.; Houser, P.R.; Meyers, T.P.; Prueger, J.H.; Starks, P.J.; Wesely, M.L. Correcting Eddy-Covariance Flux Underestimates over a Grassland. Agric. For. Meteorol. 2000, 103, 279–300. [Google Scholar] [CrossRef] [Green Version]
  87. Garcia-Santos, V.; Cuxart, J.; Jimenez, M.A.; Martinez-Villagrasa, D.; Simo, G.; Picos, R.; Caselles, V. Study of Temperature Heterogeneities at Sub-Kilometric Scales and Influence on Surface-Atmosphere Energy Interactions. IEEE Trans. Geosci. Remote Sens. 2019, 57, 640–654. [Google Scholar] [CrossRef]
  88. Dare-Idowu, O.; Brut, A.; Cuxart, J.; Tallec, T.; Rivalland, V.; Zawilski, B.; Ceschia, E.; Jarlan, L. Surface Energy Balance and Flux Partitioning of Annual Crops in Southwestern France. Agric. For. Meteorol. 2021, 308, 108529. [Google Scholar] [CrossRef]
  89. Gebler, S.; Hendricks Franssen, H.J.; Pütz, T.; Post, H.; Schmidt, M.; Vereecken, H. Actual Evapotranspiration and Precipitation Measured by Lysimeters: A Comparison with Eddy Covariance and Tipping Bucket. Hydrol. Earth Syst. Sci. 2015, 19, 2145–2161. [Google Scholar] [CrossRef] [Green Version]
  90. Bambach, N.; Kustas, W.; Alfieri, J.; Prueger, J.; Hipps, L.; Mckee, L.; Castro, S.J.; Volk, J.; Alsina, M.M.; Mcelrone, A.J. Evapotranspiration Uncertainty at Micrometeorological Scales: The Impact of the Eddy Covariance Energy Imbalance and Correction Methods. Irrig. Sci. 2022, 1, 1–17. [Google Scholar] [CrossRef]
  91. Menenti, M.; Choudhury, B.J. Exchange Processes at the Land Surface for a Range of Space and Time Scales. In Proceedings of the Exchange Processes at the Land Surface for a Range of Space and Time Scales, Yokohama, Japan, 13–16 July 1993; IAHS Press: Wallingford, UK, 1993; pp. 561–568. [Google Scholar]
  92. Roerink, G.J.; Su, Z.; Menenti, M. S-SEBI: A Simple Remote Sensing Algorithm to Estimate the Surface Energy Balance. Phys. Chem. Earth Part B 2000, 25, 147–157. [Google Scholar] [CrossRef]
  93. Zhao, D.; Zhang, W.; Liu, C. A Modified S-SEBI Algorithm to Estimate Evapotranspiration Using Landsat ETM+ Image and Meteorological Data over the Hanjiang Basin, China. In Proceedings of the 2007 IEEE International Geoscience and Remote Sensing Symposium, Barcelona, Spain, 23–28 July 2007; pp. 3253–3256. [Google Scholar]
  94. Allies, A.; Demarty, J.; Olioso, A.; Moussa, I.B.; Issoufou, H.B.A.; Velluet, C.; Bahir, M.; Maïnassara, I.; Oï, M.; Chazarin, J.P.; et al. Evapotranspiration Estimation in the Sahel Using a New Ensemble-Contextual Method. Remote Sens. 2020, 12, 380. [Google Scholar] [CrossRef] [Green Version]
  95. Gallego-Elvira, B.; Olioso, A.; Mira, M.; Reyes-Castillo, S.; Boulet, G.; Marloie, O.; Garrigues, S.; Courault, D.; Weiss, M.; Chauvelon, P.; et al. EVASPA (EVapotranspiration Assessment from SPAce) Tool: An Overview. Procedia Environ. Sci. 2013, 19, 303–310. [Google Scholar] [CrossRef]
  96. Verstraeten, W.W.; Veroustraete, F.; Feyen, J. Estimating Evapotranspiration of European Forests from NOAA-Imagery at Satellite Overpass Time: Towards an Operational Processing Chain for Integrated Optical and Thermal Sensor Data Products. Remote Sens. Environ. 2005, 96, 256–276. [Google Scholar] [CrossRef]
  97. Galleguillos, M.; Jacob, F.; Prévot, L.; French, A.; Lagacherie, P. Comparison of Two Temperature Differencing Methods to Estimate Daily Evapotranspiration over a Mediterranean Vineyard Watershed from ASTER Data. Remote Sens. Environ. 2011, 115, 1326–1340. [Google Scholar] [CrossRef]
  98. Bhattarai, N.; Shaw, S.B.; Quackenbush, L.J.; Im, J.; Niraula, R. Evaluating Five Remote Sensing Based Single-Source Surface Energy Balance Models for Estimating Daily Evapotranspiration in a Humid Subtropical Climate. Int. J. Appl. Earth Obs. Geoinf. 2016, 49, 75–86. [Google Scholar] [CrossRef]
  99. Wagle, P.; Bhattarai, N.; Gowda, P.H.; Kakani, V.G. Performance of Five Surface Energy Balance Models for Estimating Daily Evapotranspiration in High Biomass Sorghum. ISPRS J. Photogramm. Remote Sens. 2017, 128, 192–203. [Google Scholar] [CrossRef] [Green Version]
  100. Eswar, R.; Sekhar, M.; Bhattacharya, B.K. Comparison of Three Remote Sensing Based Models for the Estimation of Latent Heat Flux over India. Hydrol. Sci. J. 2017, 62, 2705–2719. [Google Scholar] [CrossRef]
  101. Sun, H.; Yang, Y.; Wu, R.; Gui, D.; Xue, J.; Liu, Y.; Yan, D. Improving Estimation of Cropland Evapotranspiration by the Bayesian Model Averaging Method with Surface Energy Balance Models. Atmosphere 2019, 10, 188. [Google Scholar] [CrossRef] [Green Version]
  102. Zhan, C.S.; Zhao, J.; Wang, H.X.; Yin, J. Quantitative Estimation of Land Surface Evapotranspiration in Taiwan Based on MODIS Data. Water Sci. Eng. 2011, 4, 237–245. [Google Scholar] [CrossRef]
  103. Zhan, C.; Yin, J.; Wang, F.; Dong, Q. Regional Estimation and Validation of Remotely Sensed Evapotranspiration in China. Catena 2015, 133, 35–42. [Google Scholar] [CrossRef]
  104. Jahangir, M.H.; Arast, M. Remote Sensing Products for Predicting Actual Evapotranspiration and Water Stress Footprints under Different Land Cover. J. Clean. Prod. 2020, 266, 121818. [Google Scholar] [CrossRef]
  105. Cong, Z.T.; Shen, Q.N.; Zhou, L.; Sun, T.; Liu, J.H. Evapotranspiration Estimation Considering Anthropogenic Heat Based on Remote Sensing in Urban Area. Sci. China Earth Sci. 2017, 60, 659–671. [Google Scholar] [CrossRef]
  106. Gokmen, M.; Vekerdy, Z.; Verhoef, A.; Verhoef, W.; Batelaan, O.; van der Tol, C. Integration of Soil Moisture in SEBS for Improving Evapotranspiration Estimation under Water Stress Conditions. Remote Sens. Environ. 2012, 121, 261–274. [Google Scholar] [CrossRef]
  107. Wu, X.; Zhou, J.; Wang, H.; Li, Y.; Zhong, B. Evaluation of Irrigation Water Use Efficiency Using Remote Sensing in the Middle Reach of the Heihe River, in the Semi-Arid Northwestern China. Hydrol. Processes 2015, 29, 2243–2257. [Google Scholar] [CrossRef]
  108. Yi, Z.; Zhao, H.; Jiang, Y.; Yan, H.; Cao, Y.; Huang, Y.; Hao, Z. Daily Evapotranspiration Estimation at the Field Scale: Using the Modified SEBS Model and HJ-1 Data in a Desert-Oasis Area, Northwestern China. Water 2018, 10, 640. [Google Scholar] [CrossRef] [Green Version]
  109. Chen, X.; Massman, W.J.; Su, Z. A Column Canopy-Air Turbulent Diffusion Method for Different Canopy Structures. J. Geophys. Res. Atmos. 2019, 124, 488–506. [Google Scholar] [CrossRef]
  110. Su, H.; Mccabe, M.F.; Wood, E.F.; Su, Z.; Prueger, J.H. Modeling Evapotranspiration during SMACEX: Comparing Two Approaches for Local-and Regional-Scale Prediction. J. Hydrometeorol. 2005, 6, 910–922. [Google Scholar] [CrossRef]
  111. Oku, Y.; Ishikawa, H.; Su, Z. Estimation of Land Surface Heat Fluxes over the Tibetan Plateau Using GMS Data. J. Appl. Meteorol. Climatol. 2007, 46, 183–195. [Google Scholar] [CrossRef] [Green Version]
  112. Su, H.; Wood, E.F.; Mccabe, M.F.; Su, Z. Evaluation of Remotely Sensed Evapotranspiration Over the CEOP EOP-1 Reference Sites. J. Meteorol. Soc. Jpn. 2007, 85, 439–459. [Google Scholar] [CrossRef] [Green Version]
  113. Yang, D.; He, C.; Lei, H. Estimation of Evapotranspiration Using a Remote Sensing Model over Agricultural Land in the North China Plain. Int. J. Remote Sens. 2010, 31, 3783–3798. [Google Scholar] [CrossRef]
  114. Tang, R.; Li, Z.L.; Jia, Y.; Li, C.; Sun, X.; Kustas, W.P.; Anderson, M.C. An Intercomparison of Three Remote Sensing-Based Energy Balance Models Using Large Aperture Scintillometer Measurements over a Wheat-Corn Production Region. Remote Sens. Environ. 2011, 115, 3187–3202. [Google Scholar] [CrossRef]
  115. Michel, D.; Jiménez, C.; Miralles, D.G.; Jung, M.; Hirschi, M.; Ershadi, A.; Martens, B.; Mccabe, M.F.; Fisher, J.B.; Mu, Q.; et al. The WACMOS-ET Project—Part 1: Tower-Scale Evaluation of Four Remote-Sensing-Based Evapotranspiration Algorithms. Hydrol. Earth Syst. Sci. 2016, 20, 803–822. [Google Scholar] [CrossRef] [Green Version]
  116. Li, G.; Jing, Y.; Wu, Y.; Zhang, F. Improvement of Two Evapotranspiration Estimation Models Using a Linear Spectral Mixture Model over a Small Agricultural Watershed. Water 2018, 10, 474. [Google Scholar] [CrossRef] [Green Version]
  117. Jamshidi, S.; Zand-Parsa, S.; Pakparvar, M.; Niyogi, D. Evaluation of Evapotranspiration over a Semiarid Region Using Multiresolution Data Sources. J. Hydrometeorol. 2019, 20, 947–964. [Google Scholar] [CrossRef]
  118. Xue, J.; Bali, K.M.; Light, S.; Hessels, T.; Kisekka, I. Evaluation of Remote Sensing-Based Evapotranspiration Models against Surface Renewal in Almonds, Tomatoes and Maize. Agric. Water Manag. 2020, 238, 106228. [Google Scholar] [CrossRef]
  119. Khand, K.; Bhattarai, N.; Taghvaeian, S.; Wagle, P.; Gowda, P.H.; Alderman, P.D. Modeling Evapotranspiration of Winter Wheat Using Contextual and Pixel-Based Surface Energy Balance Models. Trans. ASABE 2021, 64, 507–519. [Google Scholar] [CrossRef]
  120. Bastiaanssen, W.G.M.; Menenti, M.; Feddes, R.A.; Holtslag, A.A.M. A Remote Sensing Surface Energy Balance Algorithm for Land (SEBAL). 1. Formulation. J. Hydrol. 1998, 212, 198–212. [Google Scholar] [CrossRef]
  121. Allen, R.G.; Morse, A.; Tasumi, M.; Bastiaanssen, W.; Kramber, W.; Anderson, H. Evapotranspiration from Landsat (SEBAL) for Water Rights Management and Compliance with Multi-State Water Compacts. In Proceedings of the IGARSS 2001, Scanning the Present and Resolving the Future, IEEE 2001 International Geoscience and Remote Sensing Symposium, Sydney, Australia, 9–13 July 2001; IEEE: New York, NY, USA, 2002; Volume 2, pp. 830–833. [Google Scholar] [CrossRef]
  122. Long, D.; Singh, V.P. A Modified Surface Energy Balance Algorithm for Land (M-SEBAL) Based on a Trapezoidal Framework. Water Resour. Res. 2012, 48. [Google Scholar] [CrossRef]
  123. Feng, J.; Wang, Z. A Satellite-Based Energy Balance Algorithm with Reference Dry and Wet Limits. Int. J. Remote Sens. 2013, 34, 2925–2946. [Google Scholar] [CrossRef]
  124. Bhattarai, N.; Quackenbush, L.J.; Im, J.; Shaw, S.B. A New Optimized Algorithm for Automating Endmember Pixel Selection in the SEBAL and METRIC Models. Remote Sens. Environ. 2017, 196, 178–192. [Google Scholar] [CrossRef]
  125. Silva, A.M.; da Silva, R.M.; Santos, C.A.G. Automated Surface Energy Balance Algorithm for Land (ASEBAL) Based on Automating Endmember Pixel Selection for Evapotranspiration Calculation in MODIS Orbital Images. Int. J. Appl. Earth Obs. Geoinf. 2019, 79, 1–11. [Google Scholar] [CrossRef]
  126. Prakash Mohan, M.M.; Rajitha, K.; Varma, M.R.R. A Novel Approach to Select Anchor Pixels in SEBAL Model by Using Inputs from SAR Images. In Proceedings of the IGARSS 2018—2018 IEEE International Geoscience and Remote Sensing Symposium, Valencia, Spain, 22–27 July 2018; pp. 7383–7386. [Google Scholar] [CrossRef]
  127. Mkhwanazi, M.; Chávez, J.L.; Andales, A.A. SEBAL-A: A Remote Sensing ET Algorithm That Accounts for Advection with Limited Data. Part I: Development and Validation. Remote Sens. 2015, 7, 15046–15067. [Google Scholar] [CrossRef] [Green Version]
  128. Faridatul, M.I.; Wu, B.; Zhu, X.; Wang, S. Improving Remote Sensing Based Evapotranspiration Modelling in a Heterogeneous Urban Environment. J. Hydrol. 2020, 581, 124405. [Google Scholar] [CrossRef]
  129. Hessel, T.; van Opstal, J.; Trambauer, P.; Mohamed, Y.; Bastiaanssen, W. User Guide: PySEBAL for Landsat Imagery (Beta Version); IHE Delft Institute for Water Education: Delft, The Netherlands, 2017. [Google Scholar]
  130. Allen, R.G.; Burnett, B.; Kramber, W.; Huntington, J.; Kjaersgaard, J.; Kilic, A.; Kelly, C.; Trezza, R. Automated Calibration of the METRIC-Landsat Evapotranspiration Process. JAWRA J. Am. Water Resour. Assoc. 2013, 49, 563–576. [Google Scholar] [CrossRef]
  131. Laipelt, L.; Kayser, R.H.B.; Santos Fleischmann, A.; Ruhoff, A.; Bastiaanssen, W.; Erickson, T.A.; Melton, F. Long-Term Monitoring of Evapotranspiration Using the SEBAL Algorithm and Google Earth Engine Cloud Computing. ISPRS J. Photogramm. Remote Sens. 2021, 178, 81–96. [Google Scholar] [CrossRef]
  132. French, A.N.; Jacob, F.; Anderson, M.C.; Kustas, W.P.; Timmermans, W.; Gieske, A.; Su, Z.; Su, H.; McCabe, M.F.; Li, F.; et al. Surface Energy Fluxes with the Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) at the Iowa 2002 SMACEX Site (USA). Remote Sens. Environ. 2005, 99, 55–65. [Google Scholar] [CrossRef]
  133. Timmermans, W.J.; Kustas, W.P.; Anderson, M.C.; French, A.N. An Intercomparison of the Surface Energy Balance Algorithm for Land (SEBAL) and the Two-Source Energy Balance (TSEB) Modeling Schemes. Remote Sens. Environ. 2007, 108, 369–384. [Google Scholar] [CrossRef]
  134. Minacapilli, M.; Agnese, C.; Blanda, F.; Cammalleri, C.; Ciraolo, G.; D’urso, G.; Iovino, M.; Pumo, D.; Provenzano, G.; Rallo, G. Hydrology and Earth System Sciences Estimation of Actual Evapotranspiration of Mediterranean Perennial Crops by Means of Remote-Sensing Based Surface Energy Balance Models. Hydrol. Earth Syst. Sci. 2009, 13, 1061–1074. [Google Scholar] [CrossRef] [Green Version]
  135. Teixeira, A.H.d.C.; Bastiaanssen, W.G.M.; Ahmad, M.D.; Bos, M.G. Reviewing SEBAL Input Parameters for Assessing Evapotranspiration and Water Productivity for the Low-Middle São Francisco River Basin, Brazil: Part A: Calibration and Validation. Agric. For. Meteorol. 2009, 149, 462–476. [Google Scholar] [CrossRef] [Green Version]
  136. Yang, Y.; Shang, S.; Jiang, L. Remote Sensing Temporal and Spatial Patterns of Evapotranspiration and the Responses to Water Management in a Large Irrigation District of North China. Agric. For. Meteorol. 2012, 164, 112–122. [Google Scholar] [CrossRef]
  137. Rahimi, S.; Gholami Sefidkouhi, M.A.; Raeini-Sarjaz, M.; Valipour, M. Estimation of Actual Evapotranspiration by Using MODIS Images (a Case Study: Tajan Catchment). Arch. Agron. Soil Sci. 2015, 61, 695–709. [Google Scholar] [CrossRef]
  138. Dhungel, S.; Barber, M.E. Estimating Calibration Variability in Evapotranspiration Derived from a Satellite-Based Energy Balance Model. Remote Sens. 2018, 10, 1695. [Google Scholar] [CrossRef] [Green Version]
  139. Allen, R.; Irmak, A.; Trezza, R.; Hendrickx, J.M.H.; Bastiaanssen, W.; Kjaersgaard, J. Satellite-Based ET Estimation in Agriculture Using SEBAL and METRIC. Hydrol. Processes 2011, 25, 4011–4027. [Google Scholar] [CrossRef]
  140. Allen, R.G.; Morton, C.; Kamble, B.; Kilic, A.; Huntington, J.; Thau, D.; Gorelick, N.; Erickson, T.; Moore, R.; Trezza, R.; et al. EEFlux: A Landsat-Based Evapotranspiration Mapping Tool on the Google Earth Engine. In Proceedings of the 2015 ASABE/IA Irrigation Symposium: Emerging Technologies for Sustainable Irrigation—A Tribute to the Career of Terry Howell, Sr. Conference Proceedings, Long Beach, CA, USA, 10–12 November 2015; ASABE: St. Joseph, MI, USA, 2015; pp. 1–11. [Google Scholar]
  141. Choi, M.; Kustas, W.P.; Anderson, M.C.; Allen, R.G.; Li, F.; Kjaersgaard, J.H. An Intercomparison of Three Remote Sensing-Based Surface Energy Balance Algorithms over a Corn and Soybean Production Region (Iowa, U.S.) during SMACEX. Agric. For. Meteorol. 2009, 149, 2082–2097. [Google Scholar] [CrossRef]
  142. Gonzalez-Dugo, M.P.; Neale, C.M.U.; Mateos, L.; Kustas, W.P.; Prueger, J.H.; Anderson, M.C.; Li, F. A Comparison of Operational Remote Sensing-Based Models for Estimating Crop Evapotranspiration. Agric. For. Meteorol. 2009, 149, 1843–1853. [Google Scholar] [CrossRef]
  143. Paço, T.A.; Pôças, I.; Cunha, M.; Silvestre, J.C.; Santos, F.L.; Paredes, P.; Pereira, L.S. Evapotranspiration and Crop Coefficients for a Super Intensive Olive Orchard. An Application of SIMDualKc and METRIC Models Using Ground and Satellite Observations. J. Hydrol. 2014, 519, 2067–2080. [Google Scholar] [CrossRef] [Green Version]
  144. Bhattarai, N.; Quackenbush, L.J.; Dougherty, M.; Marzen, L.J. A Simple Landsat–MODIS Fusion Approach for Monitoring Seasonal Evapotranspiration at 30 m Spatial Resolution. Int. J. Remote Sens. 2015, 36, 115–143. [Google Scholar] [CrossRef]
  145. French, A.N.; Hunsaker, D.J.; Thorp, K.R. Remote Sensing of Evapotranspiration over Cotton Using the TSEB and METRIC Energy Balance Models. Remote Sens. Environ. 2015, 158, 281–294. [Google Scholar] [CrossRef]
  146. LAMBIN, E.F.; Ehrlich, D. The Surface Temperature-Vegetation Index Space for Land Cover and Land-Cover Change Analysis. Int. J. Remote Sens. 1996, 17, 463–487. [Google Scholar] [CrossRef]
  147. Priestley, C.H.B.; Taylor, R.J. On the Assessment of Surface Heat Flux and Evaporation Using Large-Scale Parameters. Mon. Weather Rev. 1972, 100, 81–92. [Google Scholar] [CrossRef]
  148. Jiang, L.; Islam, S. A Methodology for Estimation of Surface Evapotranspiration over Large Areas Using Remote Sensing Observations. Geophys. Res. Lett. 1999, 26, 2773–2776. [Google Scholar] [CrossRef] [Green Version]
  149. Stisen, S.; Sandholt, I.; Nørgaard, A.; Fensholt, R.; Jensen, K.H. Combining the Triangle Method with Thermal Inertia to Estimate Regional Evapotranspiration—Applied to MSG-SEVIRI Data in the Senegal River Basin. Remote Sens. Environ. 2008, 112, 1242–1255. [Google Scholar] [CrossRef]
  150. Minacapilli, M.; Consoli, S.; Vanella, D.; Ciraolo, G.; Motisi, A. A Time Domain Triangle Method Approach to Estimate Actual Evapotranspiration: Application in a Mediterranean Region Using MODIS and MSG-SEVIRI Products. Remote Sens. Environ. 2016, 174, 10–23. [Google Scholar] [CrossRef]
  151. Tang, R.; Li, Z.L.; Tang, B. An Application of the Ts-VI Triangle Method with Enhanced Edges Determination for Evapotranspiration Estimation from MODIS Data in Arid and Semi-Arid Regions: Implementation and Validation. Remote Sens. Environ. 2010, 114, 540–551. [Google Scholar] [CrossRef]
  152. Peng, J.; Liu, Y.; Zhao, X.; Loew, A. Estimation of Evapotranspiration from MODIS TOA Radiances in the Poyang Lake Basin, China. Hydrol. Earth Syst. Sci. 2013, 17, 1431–1444. [Google Scholar] [CrossRef] [Green Version]
  153. Shekar, S.N.C.; Nandagiri, L. A Penman-Monteith Evapotranspiration Model with Bulk Surface Conductance Derived from Remotely Sensed Spatial Contextual Information. Int. J. Remote Sens. 2019, 41, 1486–1511. [Google Scholar] [CrossRef]
  154. Jiang, L.; Islam, S. Estimation of Surface Evaporation Map over Southern Great Plains Using Remote Sensing Data. Water Resour. Res. 2001, 37, 329–340. [Google Scholar] [CrossRef] [Green Version]
  155. Nishida, K.; Nemani, R.R.; Running, S.W.; Glassy, J.M. An Operational Remote Sensing Algorithm of Land Surface Evaporation. J. Geophys. Res. Atmos. 2003, 108. [Google Scholar] [CrossRef]
  156. Batra, N.; Islam, S.; Venturini, V.; Bisht, G.; Jiang, L. Estimation and Comparison of Evapotranspiration from MODIS and AVHRR Sensors for Clear Sky Days over the Southern Great Plains. Remote Sens. Environ. 2006, 103, 1–15. [Google Scholar] [CrossRef]
  157. Jiang, L.; Islam, S.; Guo, W.; Singh Jutla, A.; Senarath, S.U.S.; Ramsay, B.H.; Eltahir, E. A Satellite-Based Daily Actual Evapotranspiration Estimation Algorithm over South Florida. Glob. Planet. Change 2009, 67, 62–77. [Google Scholar] [CrossRef]
  158. Garcia, M.; Fernández, N.; Villagarcía, L.; Domingo, F.; Puigdefábregas, J.; Sandholt, I. Accuracy of the Temperature-Vegetation Dryness Index Using MODIS under Water-Limited vs. Energy-Limited Evapotranspiration Conditions. Remote Sens. Environ. 2014, 149, 100–117. [Google Scholar] [CrossRef]
  159. Sánchez, J.M.; Caselles, V.; Niclòs, R.; Coll, C.; Kustas, W.P. Estimating Energy Balance Fluxes above a Boreal Forest from Radiometric Temperature Observations. Agric. For. Meteorol. 2009, 149, 1037–1049. [Google Scholar] [CrossRef]
  160. Häusler, M.; Conceição, N.; Tezza, L.; Sánchez, J.M.; Campagnolo, M.L.; Häusler, A.J.; Silva, J.M.N.; Warneke, T.; Heygster, G.; Ferreira, M.I. Estimation and Partitioning of Actual Daily Evapotranspiration at an Intensive Olive Grove Using the STSEB Model Based on Remote Sensing. Agric. Water Manag. 2018, 201, 188–198. [Google Scholar] [CrossRef]
  161. Kustas, W.P.; Norman, J.M. Evaluation of Soil and Vegetation Heat Flux Predictions Using a Simple Two-Source Model with Radiometric Temperatures for Partial Canopy Cover. Agric. For. Meteorol. 1999, 94, 13–29. [Google Scholar] [CrossRef]
  162. Anderson, M.C.; Norman, J.M.; Kustas, W.P.; Houborg, R.; Starks, P.J.; Agam, N. A Thermal-Based Remote Sensing Technique for Routine Mapping of Land-Surface Carbon, Water and Energy Fluxes from Field to Regional Scales. Remote Sens. Environ. 2008, 112, 4227–4241. [Google Scholar] [CrossRef]
  163. Nieto, H.; Kustas, W.P.; Torres-Rúa, A.; Alfieri, J.G.; Gao, F.; Anderson, M.C.; White, W.A.; Song, L.; Alsina, M.d.M.; Prueger, J.H.; et al. Evaluation of TSEB Turbulent Fluxes Using Different Methods for the Retrieval of Soil and Canopy Component Temperatures from UAV Thermal and Multispectral Imagery. Irrig. Sci. 2019, 37, 389–406. [Google Scholar] [CrossRef] [Green Version]
  164. Kustas, P.W.; Norman, J.M. Reply to Comments about the Basic Equations of Dual-Source Vegetation–Atmosphere Transfer Models. Agric. For. Meteorol. 1999, 94, 275–278. [Google Scholar] [CrossRef]
  165. Blyth, E.M.; Harding, R.J. Application of Aggregation Models to Surface Heat Flux from the Sahelian Tiger Bush. Agric. For. Meteorol. 1995, 72, 213–235. [Google Scholar] [CrossRef]
  166. Lhomme, J.P.; Chehbouni, A. Comments on Dual-Source Vegetation–Atmosphere Transfer Models. Agric. For. Meteorol. 1999, 94, 269–273. [Google Scholar] [CrossRef]
  167. Mecikalski, J.R.; Diak, G.R.; Anderson, M.C.; Norman, J.M. Estimating Fluxes on Continental Scales Using Remotely Sensed Data in an Atmospheric-Land Exchange Model. J. Appl. Meteorol. 1999, 38, 1352–1369. [Google Scholar] [CrossRef]
  168. Norman, J.M.; Anderson, M.C.; Kustas, W.P.; French, A.N.; Mecikalski, J.; Torn, R.; Diak, G.R.; Schmugge, T.J.; Tanner, B.C.W. Remote Sensing of Surface Energy Fluxes at 101-m Pixel Resolutions. Water Resour. Res. 2003, 39. [Google Scholar] [CrossRef] [Green Version]
  169. Sánchez, J.M.; Kustas, W.P.; Caselles, V.; Anderson, M.C. Modelling Surface Energy Fluxes over Maize Using a Two-Source Patch Model and Radiometric Soil and Canopy Temperature Observations. Remote Sens. Environ. 2008, 112, 1130–1143. [Google Scholar] [CrossRef]
  170. Gan, G.; Kang, T.; Yang, S.; Bu, J.; Feng, Z.; Gao, Y. An Optimized Two Source Energy Balance Model Based on Complementary Concept and Canopy Conductance. Remote Sens. Environ. 2019, 223, 243–256. [Google Scholar] [CrossRef]
  171. Li, F.; Kustas, W.P.; Prueger, J.H.; Neale, C.M.U.; Jackson, T.J. Utility of Remote Sensing-Based Two-Source Energy Balance Model under Low-and High-Vegetation Cover Conditions. J. Hydrometeorol. 2005, 6, 878–891. [Google Scholar] [CrossRef]
  172. Li, F.; Kustas, W.P.; Anderson, M.C.; Prueger, J.H.; Scott, R.L. Effect of Remote Sensing Spatial Resolution on Interpreting Tower-Based Flux Observations. Remote Sens. Environ. 2008, 112, 337–349. [Google Scholar] [CrossRef]
  173. Anderson, M.C.; Allen, R.G.; Morse, A.; Kustas, W.P. Use of Landsat Thermal Imagery in Monitoring Evapotranspiration and Managing Water Resources. Remote Sens. Environ. 2012, 122, 50–65. [Google Scholar] [CrossRef]
  174. Long, D.; Singh, V.P. A Two-Source Trapezoid Model for Evapotranspiration (TTME) from Satellite Imagery. Remote Sens. Environ. 2012, 121, 370–388. [Google Scholar] [CrossRef]
  175. Cammalleri, C.; Anderson, M.C.; Gao, F.; Hain, C.R.; Kustas, W.P. A Data Fusion Approach for Mapping Daily Evapotranspiration at Field Scale. Water Resour. Res. 2013, 49, 4672–4686. [Google Scholar] [CrossRef]
  176. Cammalleri, C.; Anderson, M.C.; Gao, F.; Hain, C.R.; Kustas, W.P. Mapping Daily Evapotranspiration at Field Scales over Rainfed and Irrigated Agricultural Areas Using Remote Sensing Data Fusion. Agric. For. Meteorol. 2014, 186, 1–11. [Google Scholar] [CrossRef] [Green Version]
  177. Yang, Y.; Long, D.; Guan, H.; Liang, W.; Simmons, C.; Batelaan, O. Comparison of Three Dual-Source Remote Sensing Evapotranspiration Models during the MUSOEXE-12 Campaign: Revisit of Model Physics. Water Resour. Res. 2015, 51, 3145–3165. [Google Scholar] [CrossRef]
  178. Ortega-Farías, S.; Ortega-Salazar, S.; Poblete, T.; Kilic, A.; Allen, R.; Poblete-Echeverría, C.; Ahumada-Orellana, L.; Zuñiga, M.; Sepúlveda, D. Estimation of Energy Balance Components over a Drip-Irrigated Olive Orchard Using Thermal and Multispectral Cameras Placed on a Helicopter-Based Unmanned Aerial Vehicle (UAV). Remote Sens. 2016, 8, 638. [Google Scholar] [CrossRef] [Green Version]
  179. Semmens, K.A.; Anderson, M.C.; Kustas, W.P.; Gao, F.; Alfieri, J.G.; McKee, L.; Prueger, J.H.; Hain, C.R.; Cammalleri, C.; Yang, Y.; et al. Monitoring Daily Evapotranspiration over Two California Vineyards Using Landsat 8 in a Multi-Sensor Data Fusion Approach. Remote Sens. Environ. 2016, 185, 155–170. [Google Scholar] [CrossRef] [Green Version]
  180. Song, L.; Liu, S.; Kustas, W.P.; Zhou, J.; Xu, Z.; Xia, T.; Li, M. Application of Remote Sensing-Based Two-Source Energy Balance Model for Mapping Field Surface Fluxes with Composite and Component Surface Temperatures. Agric. For. Meteorol. 2016, 230, 8–19. [Google Scholar] [CrossRef] [Green Version]
  181. Yang, Y.; Anderson, M.C.; Gao, F.; Hain, C.R.; Semmens, K.A.; Kustas, W.P.; Noormets, A.; Wynne, R.H.; Thomas, V.A.; Sun, G. Daily Landsat-Scale Evapotranspiration Estimation over a Forested Landscape in North Carolina, USA, Using Multi-Satellite Data Fusion. Hydrol. Earth Syst. Sci. 2017, 21, 1017–1037. [Google Scholar] [CrossRef] [Green Version]
  182. Feng, J.; Wang, W.; Xu, F.; Sun, S. Estimating Surface Heat and Water Vapor Fluxes by Combining Two-Source Energy Balance Model and Back-Propagation Neural Network. Sci. Total Environ. 2020, 729, 138724. [Google Scholar] [CrossRef]
  183. Carlson, T.N. Limitations and the Value of Land Surface Models and Their Role in Remote Sensing. Remote Sens. Lett. 2022, 1–5. [Google Scholar] [CrossRef]
  184. Anderson, M.C.; Kustas, W.P.; Norman, J.M.; Hain, C.R.; Mecikalski, J.R.; Schultz, L.; González-Dugo, M.P.; Cammalleri, C.; D’Urso, G.; Pimstein, A.; et al. Mapping Daily Evapotranspiration at Field to Continental Scales Using Geostationary and Polar Orbiting Satellite Imagery. Hydrol. Earth Syst. Sci. 2011, 15, 223–239. [Google Scholar] [CrossRef] [Green Version]
  185. Gao, F.; Hilker, T.; Zhu, X.; Anderson, M.; Masek, J.; Wang, P.; Yang, Y. Fusing Landsat and MODIS Data for Vegetation Monitoring. IEEE Geosci. Remote Sens. Mag. 2015, 3, 47–60. [Google Scholar] [CrossRef]
  186. Bindhu, V.M.; Narasimhan, B.; Sudheer, K.P. Development and Verification of a Non-Linear Disaggregation Method (NL-DisTrad) to Downscale MODIS Land Surface Temperature to the Spatial Scale of Landsat Thermal Data to Estimate Evapotranspiration. Remote Sens. Environ. 2013, 135, 118–129. [Google Scholar] [CrossRef]
  187. Ke, Y.; Im, J.; Park, S.; Gong, H. Downscaling of MODIS One Kilometer Evapotranspiration Using Landsat-8 Data and Machine Learning Approaches. Remote Sens. 2016, 8, 215. [Google Scholar] [CrossRef] [Green Version]
  188. Bisquert, M.; Sánchez, J.M.; López-Urrea, R.; Caselles, V. Estimating High Resolution Evapotranspiration from Disaggregated Thermal Images. Remote Sens. Environ. 2016, 187, 423–433. [Google Scholar] [CrossRef]
  189. Bisquert, M.; Sánchez, J.M.; Caselles, V. Evaluation of Disaggregation Methods for Downscaling MODIS Land Surface Temperature to Landsat Spatial Resolution in Barrax Test Site. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2016, 9, 1430–1438. [Google Scholar] [CrossRef] [Green Version]
  190. Mokhtari, A.; Noory, H.; Pourshakouri, F.; Haghighatmehr, P.; Afrasiabian, Y.; Razavi, M.; Fereydooni, F.; Sadeghi Naeni, A. Calculating Potential Evapotranspiration and Single Crop Coefficient Based on Energy Balance Equation Using Landsat 8 and Sentinel-2. ISPRS J. Photogramm. Remote Sens. 2019, 154, 231–245. [Google Scholar] [CrossRef]
  191. Guzinski, R.; Nieto, H.; Sandholt, I.; Karamitilios, G. Modelling High-Resolution Actual Evapotranspiration through Sentinel-2 and Sentinel-3 Data Fusion. Remote Sens. 2020, 12, 1433. [Google Scholar] [CrossRef]
  192. Bellvert, J.; Jofre-Ĉekalović, C.; Pelechá, A.; Mata, M.; Nieto, H. Feasibility of Using the Two-Source Energy Balance Model (TSEB) with Sentinel-2 and Sentinel-3 Images to Analyze the Spatio-Temporal Variability of Vine Water Status in a Vineyard. Remote Sens. 2020, 12, 2299. [Google Scholar] [CrossRef]
  193. Allam, M.; Mhawej, M.; Meng, Q.; Faour, G.; Abunnasr, Y.; Fadel, A.; Xinli, H. Monthly 10-m Evapotranspiration Rates Retrieved by SEBALI with Sentinel-2 and MODIS LST Data. Agric. Water Manag. 2021, 243, 106432. [Google Scholar] [CrossRef]
  194. Guzinski, R.; Nieto, H.; Sanchez, J.M.; Lopez-Urrea, R.; Boujnah, D.M.; Boulet, G. Utility of Copernicus-Based Inputs for Actual Evapotranspiration Modeling in Support of Sustainable Water Use in Agriculture. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2021, 14, 11466–11484. [Google Scholar] [CrossRef]
  195. Campbell, G.S.; Norman, J.M. An Introduction to Environmental Biophysics; Springer: Berlin/Heidelberg, Germany, 1998. [Google Scholar] [CrossRef]
  196. Gowda, P.H.; Chavez, J.L.; Colaizzi, P.D.; Evett, S.R.; Howell, T.A.; Tolk, J.A. ET Mapping for Agricultural Water Management: Present Status and Challenges. Irrig. Sci. 2007, 26, 223–237. [Google Scholar] [CrossRef] [Green Version]
  197. Ward, H.C. Scintillometry in Urban and Complex Environments: A Review. Meas. Sci. Technol. 2017, 28, 064005. [Google Scholar] [CrossRef]
  198. Vilà-Guerau De Arellano, J.; Ney, P.; Hartogensis, O.; de Boer, H.; van Diepen, K.; Emin, D.; de Groot, G.; Klosterhalfen, A.; Langensiepen, M.; Matveeva, M.; et al. CloudRoots: Integration of Advanced Instrumental Techniques and Process Modelling of Sub-Hourly and Sub-Kilometre Land-Atmosphere Interactions. Biogeosciences 2020, 17, 4375–4404. [Google Scholar] [CrossRef]
  199. Aboutalebi, M.; Torres-Rua, A.F.; McKee, M.; Kustas, W.P.; Nieto, H.; Alsina, M.M.; White, A.; Prueger, J.H.; McKee, L.; Alfieri, J.; et al. Incorporation of Unmanned Aerial Vehicle (UAV) Point Cloud Products into Remote Sensing Evapotranspiration Models. Remote Sens. 2019, 12, 50. [Google Scholar] [CrossRef] [Green Version]
  200. Brenner, C.; Thiem, C.E.; Wizemann, H.-D.; Bernhardt, M.; Schulz, K. Estimating Spatially Distributed Turbulent Heat Fluxes from High-Resolution Thermal Imagery Acquired with a UAV System. Int. J. Remote Sens. 2017, 38, 3003–3026. [Google Scholar] [CrossRef] [Green Version]
  201. Thorp, K.R.; Thompson, A.L.; Harders, S.J.; French, A.N.; Ward, R.W. High-Throughput Phenotyping of Crop Water Use Efficiency via Multispectral Drone Imagery and a Daily Soil Water Balance Model. Remote Sens. 2018, 10, 1682. [Google Scholar] [CrossRef] [Green Version]
  202. Tang, J.; Han, W.; Zhang, L. UAV Multispectral Imagery Combined with the FAO-56 Dual Approach for Maize Evapotranspiration Mapping in the North China Plain. Remote Sens. 2019, 11, 2519. [Google Scholar] [CrossRef] [Green Version]
  203. Torres-Rua, A.; Ticlavilca, A.M.; Aboutalebi, M.; Nieto, H.; Alsina, M.M.; White, A.; Prueger, J.H.; Alfieri, J.; Hipps, L.; McKee, L.; et al. Estimation of Evapotranspiration and Energy Fluxes Using a Deep-Learning-Based High-Resolution Emissivity Model and the Two-Source Energy Balance Model with SUAS Information. In Proceedings of the Autonomous Air and Ground Sensing Systems for Agricultural Optimization and Phenotyping V, Online, 27 April–9 May 2020; Volume 11414, p. 114140B. [Google Scholar] [CrossRef]
  204. Park, S.; Ryu, D.; Fuentes, S.; Chung, H.; O’Connell, M.; Kim, J. Mapping Very-High-Resolution Evapotranspiration from Unmanned Aerial Vehicle (UAV) Imagery. ISPRS Int. J. Geo-Inf. 2021, 10, 211. [Google Scholar] [CrossRef]
  205. Gautam, D.; Ostendorf, B.; Pagay, V. Estimation of Grapevine Crop Coefficient Using a Multispectral Camera on an Unmanned Aerial Vehicle. Remote Sens. 2021, 13, 2639. [Google Scholar] [CrossRef]
  206. Mokhtari, A.; Ahmadi, A.; Daccache, A.; Drechsler, K. Actual Evapotranspiration from UAV Images: A Multi-Sensor Data Fusion Approach. Remote Sens. 2021, 13, 2315. [Google Scholar] [CrossRef]
  207. Mu, Q.; Heinsch, F.A.; Zhao, M.; Running, S.W. Development of a Global Evapotranspiration Algorithm Based on MODIS and Global Meteorology Data. Remote Sens. Environ. 2007, 111, 519–536. [Google Scholar] [CrossRef]
  208. Trigo, I.F.; de Bruin, H.; Beyrich, F.; Bosveld, F.C.; Gavilán, P.; Groh, J.; López-Urrea, R. Validation of Reference Evapotranspiration from Meteosat Second Generation (MSG) Observations. Agric. For. Meteorol. 2018, 259, 271–285. [Google Scholar] [CrossRef]
  209. Jung, M.; Koirala, S.; Weber, U.; Ichii, K.; Gans, F.; Camps-Valls, G.; Papale, D.; Schwalm, C.; Tramontana, G.; Reichstein, M. The FLUXCOM Ensemble of Global Land-Atmosphere Energy Fluxes. Sci. Data 2019, 6, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. FAO. WaPOR Database Methodol; FAO: Rome, Italy, 2020. [Google Scholar] [CrossRef]
  211. Ghilain, N.; Arboleda, A.; Gellens-Meulenberghs, F. Evapotranspiration Modelling at Large Scale Using Near-Real Time MSG SEVIRI Derived Data. Hydrol. Earth Syst. Sci. 2011, 15, 771–786. [Google Scholar] [CrossRef] [Green Version]
  212. Anderson, M.C.; Yang, Y.; Xue, J.; Knipper, K.R.; Yang, Y.; Gao, F.; Hain, C.R.; Kustas, W.P.; Cawse-Nicholson, K.; Hulley, G.; et al. Interoperability of ECOSTRESS and Landsat for Mapping Evapotranspiration Time Series at Sub-Field Scales. Remote Sens. Environ. 2021, 252, 112189. [Google Scholar] [CrossRef]
  213. Fisher, J.B.; Lee, B.; Purdy, A.J.; Halverson, G.H.; Dohlen, M.B.; Cawse-Nicholson, K.; Wang, A.; Anderson, R.G.; Aragon, B.; Arain, M.A.; et al. ECOSTRESS: NASA’s Next Generation Mission to Measure Evapotranspiration from the International Space Station. Water Resour. Res. 2020, 56, e2019WR026058. [Google Scholar] [CrossRef]
  214. Hu, G.; Jia, L.; Menenti, M. Comparison of MOD16 and LSA-SAF MSG Evapotranspiration Products over Europe for 2011. Remote Sens. Environ. 2015, 156, 510–526. [Google Scholar] [CrossRef]
  215. Aguilar, A.L.; Flores, H.; Crespo, G.; Marín, M.I.; Campos, I.; Calera, A. Performance Assessment of MOD16 in Evapotranspiration Evaluation in Northwestern Mexico. Water 2018, 10, 901. [Google Scholar] [CrossRef] [Green Version]
  216. Chang, Y.; Qin, D.; Ding, Y.; Zhao, Q.; Zhang, S. A Modified MOD16 Algorithm to Estimate Evapotranspiration over Alpine Meadow on the Tibetan Plateau, China. J. Hydrol. 2018, 561, 16–30. [Google Scholar] [CrossRef]
  217. Degano, M.F.; Rivas, R.E.; Carmona, F.; Niclòs, R.; Sánchez, J.M. Evaluation of the MOD16A2 Evapotranspiration Product in an Agricultural Area of Argentina, the Pampas Region. Egypt. J. Remote Sens. Space Sci. 2021, 24, 319–328. [Google Scholar] [CrossRef]
  218. Miralles, D.G.; Holmes, T.R.H.; de Jeu, R.A.M.; Gash, J.H.; Meesters, A.G.C.A.; Dolman, A.J. Global Land-Surface Evaporation Estimated from Satellite-Based Observations. Hydrol. Earth Syst. Sci. 2011, 15, 453–469. [Google Scholar] [CrossRef] [Green Version]
  219. Martens, B.; Miralles, D.G.; Lievens, H.; van der Schalie, R.; de Jeu, R.A.M.; Fernández-Prieto, D.; Beck, H.E.; Dorigo, W.A.; Verhoest, N.E.C. GLEAM v3: Satellite-Based Land Evaporation and Root-Zone Soil Moisture. Geosci. Model Dev. 2017, 10, 1903–1925. [Google Scholar] [CrossRef] [Green Version]
  220. Irmak, A.; Allen, R.G.; Kjaersgaard, J.; Huntington, J.; Kamble, B.; Trezza, R.; Ratcliffe, I. Evapotranspiration—Remote Sensing and Modeling. Chapter Book 21: Operational Remote Sensing of ET and Challenges; InTech: Rijeka, Croatia, 2011; ISBN 9789533078083. [Google Scholar]
  221. Melton, F.S.; Huntington, J.; Grimm, R.; Herring, J.; Hall, M.; Rollison, D.; Erickson, T.; Allen, R.; Anderson, M.; Fisher, J.B.; et al. OpenET: Filling a Critical Data Gap in Water Management for the Western United States. JAWRA J. Am. Water Resour. Assoc. 2021, 1–24. [Google Scholar] [CrossRef]
  222. Kustas, W.P.; Agam, N.; Ortega-Farias, S. Forward to the GRAPEX Special Issue. Irrig. Sci. 2019, 37, 221–226. [Google Scholar] [CrossRef] [Green Version]
  223. TEAM, HyspIRI Mission Concept. HyspIRI Final Report; Jet Propulsion Laboratory, California Institute of Technology: Pasadena, CA, USA, 2018. [Google Scholar]
  224. Lagouarde, J.P.; Bach, M.; Sobrino, J.A.; Boulet, G.; Briottet, X.; Cherchali, S.; Coudert, B.; Dadou, I.; Dedieu, G.; Gamet, P.; et al. The Mistigri Thermal Infrared Project: Scientific Objectives And Mission Specifications. Int. J. Remote Sens. 2013, 34, 3437–3466. [Google Scholar] [CrossRef]
  225. Masek, J.G.; Wulder, M.A.; Markham, B.; McCorkel, J.; Crawford, C.J.; Storey, J.; Jenstrom, D.T. Landsat 9: Empowering Open Science and Applications through Continuity. Remote Sens. Environ. 2020, 248, 111968. [Google Scholar] [CrossRef]
  226. Hulley, G.C.; Göttsche, F.M.; Rivera, G.; Hook, S.J.; Freepartner, R.J.; Martin, M.A.; Cawse-Nicholson, K.; Johnson, W.R. Validation and Quality Assessment of the ECOSTRESS Level-2 Land Surface Temperature and Emissivity Product. IEEE Trans. Geosci. Remote Sens. 2022, 60, 1–23. [Google Scholar] [CrossRef]
  227. Koetz, B.; Bastiaanssen, W.; Berger, M.; Defourney, P.; del Bello, U.; Drusch, M.; Drinkwater, M.; Duca, R.; Fernandez, V.; Ghent, D.; et al. High Spatio-Temporal Resolution Land Surface Temperature Mission—A Copernicus Candidate Mission in Support of Agricultural Monitoring. In Proceedings of the IGARSS 2018—2018 IEEE International Geoscience and Remote Sensing Symposium, Valencia, Spain, 22–27 July 2018. [Google Scholar]
  228. Lagouarde, J.P.; Bhattacharya, B.K.; Crébassol, P.; Gamet, P.; Babu, S.S.; Boulet, G.; Briottet, X.; Buddhiraju, K.M.; Cherchali, S.; Dadou, I.; et al. The Indian-French Trishna Mission: Earth Observation in the Thermal Infrared with High Spatio-Temporal Resolution. In Proceedings of the IGARSS 2018-2018 IEEE International Geoscience and Remote Sensing Symposium, Valencia, Spain, 22–27 July 2018; pp. 4078–4081. [Google Scholar] [CrossRef]
  229. Jiang, C.; Ryu, Y. Multi-Scale Evaluation of Global Gross Primary Productivity and Evapotranspiration Products Derived from Breathing Earth System Simulator (BESS). Remote Sens. Environ. 2016, 186, 528–547. [Google Scholar] [CrossRef]
  230. Ke, Y.; Im, J.; Park, S.; Gong, H. Spatiotemporal Downscaling Approaches for Monitoring 8-Day 30 m Actual Evapotranspiration. ISPRS J. Photogramm. Remote Sens. 2017, 126, 79–93. [Google Scholar] [CrossRef]
  231. Gonzalo-Martin, C.; Lillo-Saavedra, M.; Garcia-Pedrero, A.; Lagos, O.; Menasalvas, E. Daily Evapotranspiration Mapping Using Regression Random Forest Models. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2017, 10, 5359–5368. [Google Scholar] [CrossRef]
  232. Poon, P.K.; Kinoshita, A.M. Estimating Evapotranspiration in a Post-Fire Environment Using Remote Sensing and Machine Learning. Remote Sens. 2018, 10, 1728. [Google Scholar] [CrossRef] [Green Version]
  233. Carter, C.; Liang, S. Evaluation of Ten Machine Learning Methods for Estimating Terrestrial Evapotranspiration from Remote Sensing. Int. J. Appl. Earth Obs. Geoinf. 2019, 78, 86–92. [Google Scholar] [CrossRef]
  234. Quintano, C.; Fernández-Manso, A.; Roberts, D.A. Enhanced Burn Severity Estimation Using Fine Resolution ET and MESMA Fraction Images with Machine Learning Algorithm. Remote Sens. Environ. 2020, 244, 111815. [Google Scholar] [CrossRef]
  235. Kim, N.; Kim, K.; Lee, S.; Cho, J.; Lee, Y. Retrieval of Daily Reference Evapotranspiration for Croplands in South Korea Using Machine Learning with Satellite Images and Numerical Weather Prediction Data. Remote Sens. 2020, 12, 3642. [Google Scholar] [CrossRef]
  236. Käfer, P.S.; da Rocha, N.S.; Diaz, L.R.; Kaiser, E.A.; Santos, D.C.; Veeck, G.P.; Robérti, D.R.; Rolim, S.B.A.; de Oliveira, G.G. Artificial Neural Networks Model Based on Remote Sensing to Retrieve Evapotranspiration over the Brazilian Pampa. J. Appl. Remote Sens. 2020, 14, 038504. [Google Scholar] [CrossRef]
  237. Cui, Y.; Song, L.; Fan, W. Generation of Spatio-Temporally Continuous Evapotranspiration and Its Components by Coupling a Two-Source Energy Balance Model and a Deep Neural Network over the Heihe River Basin. J. Hydrol. 2021, 597, 126176. [Google Scholar] [CrossRef]
  238. Ahansal, Y.; Bouziani, M.; Yaagoubi, R.; Sebari, I.; Sebari, K.; Kenny, L. Towards Smart Irrigation: A Literature Review on the Use of Geospatial Technologies and Machine Learning in the Management of Water Resources in Arboriculture. Agronomy 2022, 12, 297. [Google Scholar] [CrossRef]
  239. Burchard-Levine, V.; Nieto, H.; Kustas, W.P.; Gao, F.; Alfieri, J.G.; Prueger, J.H.; Hipps, L.E.; Bambach-Ortiz, N.; McElrone, A.J.; Castro, S.J.; et al. Application of a Remote-Sensing Three-Source Energy Balance Model to Improve Evapotranspiration Partitioning in Vineyards. Irrig. Sci. 2022, 1, 1–16. [Google Scholar] [CrossRef]
  240. Yao, Y.; Liang, S.; Li, X.; Chen, J.; Liu, S.; Jia, K.; Zhang, X.; Xiao, Z.; Fisher, J.B.; Mu, Q.; et al. Improving Global Terrestrial Evapotranspiration Estimation Using Support Vector Machine by Integrating Three Process-Based Algorithms. Agric. For. Meteorol. 2017, 242, 55–74. [Google Scholar] [CrossRef]
  241. Yang, Y.; Sun, H.; Xue, J.; Liu, Y.; Liu, L.; Yan, D.; Gui, D. Estimating Evapotranspiration by Coupling Bayesian Model Averaging Methods with Machine Learning Algorithms. Environ. Monit. Assess. 2021, 193, 1–15. [Google Scholar] [CrossRef]
Figure 1. Schematic relationship between Ts and albedo in the SEBI/S-SEBI algorithm (extracted from [17]).
Figure 1. Schematic relationship between Ts and albedo in the SEBI/S-SEBI algorithm (extracted from [17]).
Remotesensing 14 03440 g001
Figure 2. (a) Simplified Ts−VI triangular diagram. (b) (Ts-Ta)-FVC trapezoidal space. (Both diagrams extracted from [17]).
Figure 2. (a) Simplified Ts−VI triangular diagram. (b) (Ts-Ta)-FVC trapezoidal space. (Both diagrams extracted from [17]).
Remotesensing 14 03440 g002
Figure 3. Schematic diagram illustrating resistance network and key input/output parameters and variables for the: (a) ‘Parallel’ and (b) Series configuration (adapted from [43]).
Figure 3. Schematic diagram illustrating resistance network and key input/output parameters and variables for the: (a) ‘Parallel’ and (b) Series configuration (adapted from [43]).
Remotesensing 14 03440 g003
Table 1. Representative chronological compendium of validation studies of the S-SEBI method. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties on ETd (mm/d) comprise Root Mean Square Error (RMSE) and bias.
Table 1. Representative chronological compendium of validation studies of the S-SEBI method. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties on ETd (mm/d) comprise Root Mean Square Error (RMSE) and bias.
Year, ReferenceSRNSensorUncertainty (mm/d)Type of Surface
2005, [44]20 m19AirborneRMSE/bias = 1/0.2corn, alfalfa, sunflower and wheat
2005, [96]1100 m520AVHRRRMSE = 1.2forest
2011, [97]90 m11ASTERRMSE = 0.8vineyard
2014, [45]90 m7ASTERRMSE/bias = 4/1wheat, brocoli, frijoles chili, potatoes, chickpea, sunflower, orange and corn
2016, [98]30 m149Landsat 5-TM
Landsat 7-ETM
RMSE = 0.9Citric Orchad, grazing, swamps, lakes
2017, [99]30 m19Landsat 7-ETM
Landsat 8-TIRS
RMSE = 0.9sorghum
2017, [100]1000 m248MODISRMSE/bias = 2/0.2Pine trees, Green manure–weed– mustard (irrigated), Rice–rice (irrigated), Soybean–wheat
(irrigated), Mixed crops (sugarcane, vegetables, turmeric, maize)
2019, [101]30 m52Landsat 8-TIRSRMSE = 1.12Cropland (corn and muskmelon)
Table 2. Representative chronological compendium of validation studies of the SEBS algorithm. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties are Root Mean Square Error (RMSE), bias, and Mean Absolute Difference (MAD), are shown for ETd (mm/d). Different uncertainty rows inside the cell means uncertainty associated to the satellite sensor.
Table 2. Representative chronological compendium of validation studies of the SEBS algorithm. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties are Root Mean Square Error (RMSE), bias, and Mean Absolute Difference (MAD), are shown for ETd (mm/d). Different uncertainty rows inside the cell means uncertainty associated to the satellite sensor.
Year, ReferenceSRNSensorUncertainty (mm/d)Type of Surface
2002, [46]Tower footprint scale620Field Campaign
Data
RMSE/MAD = 1 to 3/0.9 to 1.9cotton
shrub, brush
grass
2005, [110]30 m1Lansat 7-ETMRMSE/bias = 1 to 3/0.3 to 0.6corn
soybean
2006, [14]60 m1Landsat 7-ETMbias = 0.2corn, soybean
90 mASTERbias = 0.6
1000 mMODISbias = 2
2007, [111]7–10 km935GSM 5-VISSRRMSE = 5Tibetan plateau
2007, [112]1000 m164MODISRMSE/bias = 0.7 to 4/−1 to −0.17Meadow, forest and corn
2010, [113]1000 m-MODISRMSE/bias/MAD = 2/0.2 to 1.1/1.6wheat, corn
2011, [114]1000 m33MODISRMSE/bias = 3/1.7wheat, corn
2014, [45]90 m7ASTERRMSE = 5wheat, brocoli, frijoles chili, potatoes, chickpea sunflower, orange and corn
2015, [115]1000 m-AATSRRMSE = 1.2varied non-especified surfaces (crop, forest, etc.)
2016, [98]30 m149Landsat 5-TM
Landsat 7-ETM
RMSE = 0.74Citric Orchad, grazing, swamps, lakes
2017, [99]30 m19Landsat 7-ETM
Landsat 8-TIRS
RMSE = 1.1sorghum
2018, [116]30 m11Landsat 8-TIRSRMSE/MAD = 0.22/0.21double-cropped rice, peanut/sweet potato rotation, and orange groves
2019, [117]30 m22Landsat 5-TM
Landsat 7-ETM
RMSE = 1.8wheat (dominant crop), barley or cotton in winter
2019, [101]30 m52Landsat 8-TIRSRMSE = 1.3Cropland (corn and muskmelon)
2020, [118]30 m27Landsat 8-TIRSRMSE = 0.8processing tomatoes and maize
2021, [119]30 m42Landsat 7-ETM
Landsat 8-TIRS
RMSE/bias/MAD = 1.3/−0.2/0.9winter wheat
Table 3. Representative chronological compendium of validation studies of the SEBAL algorithm. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties Root Mean Square Error (RMSE), bias, and Mean Absolute Difference (MAD), are shown for ETd (mm/d). Different uncertainty rows inside the cell means uncertainty associated to the satellite sensor.
Table 3. Representative chronological compendium of validation studies of the SEBAL algorithm. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties Root Mean Square Error (RMSE), bias, and Mean Absolute Difference (MAD), are shown for ETd (mm/d). Different uncertainty rows inside the cell means uncertainty associated to the satellite sensor.
Year, ReferenceSRNSensorUncertainty (mm/d)Type of Surface
2005, [132]90 m1ASTERRMSE/bias = 1.3/0.04corn, soybean
2007, [133]6–12 m5airborneRMSE/MAD = 2.4/2.0shrubs, meadow wheat
2009, [134]15 m1airbornebias = 0.02olive, vineyard and citric orchads
2009, [135]30 m10Landsat 5-TMRMSE = 1.2Mango, vineyard, vegetation
Landsat 7-ETMRMSE = 0.4
2012, [122]30 m3Landsat 5-TMRMSE/bias = 1.9/−0.5corn, soybean
Landsat 7-ETMRMSE/bias * = 1.4/−0.14
2012, [136]1000 m302MODISRMSE = 0.5wheat, corn, sunflower
2015, [137]1000 m7MODISRMSE = 1.5several non-especified crops
2016, [98]30 m149Landsat 5-TM
Landsat 7-ETM
RMSE = 0.8Citric Orchad, grazing, swamps, lakes
2017, [99]30 m19Landsat 7-ETM
Landsat 8-TIRS
RMSE = 0.97sorghum
2018, [116]30 m11Landsat 8-TIRSRMSE/MAD = 0.4/0.5double-cropped rice, peanut/sweet potato rotation, and orange groves
2019, [101]30 m52Landsat 8-TIRSRMSE = 1.3Cropland (corn and muskmelon)
2020, [118]30 m27Landsat 8-TIRSRMSE = 1.3Almond, processing tomatoes and maize
2021, [119]30 m42Landsat 7-ETM
Landsat 8-TIRS
RMSE/bias/MAD = 1/−0.4/0.8winter wheat
* Statistical uncertainties after applying the M-SEBAL method.
Table 4. Representative chronological compendium of validation studies of the METRIC algorithm. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties Root Mean Square Error (RMSE), and bias, are shown for ETd (mm/d). Different uncertainty rows inside the cell means uncertainty associated to the satellite sensor.
Table 4. Representative chronological compendium of validation studies of the METRIC algorithm. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties Root Mean Square Error (RMSE), and bias, are shown for ETd (mm/d). Different uncertainty rows inside the cell means uncertainty associated to the satellite sensor.
Year, ReferenceSRNSensorUncertainty (mm/d)Type of Surface
2009, [141]60 m2Landsat 5-TM
Landsat 7-ETM
RMSE/bias = 2.4/0.2corn, soybean
2009, [142]120 m
60 m
2
1
Landsat 5-TM
Landsat 7-ETM
RMSE/bias = 0.6/−0.3 corn, soybean
2014, [143]120 m
60 m
16Landsat 5-TM
Landsat 7-ETM
RMSE/bias = 0.7/0.5 olive orchads
2015, [144]30 m12Landsat 5-TMRMSE/bias = 0.8/−0.1cocoa, cotton, wheat, soybean
MODISRMSE/bias * = 0.5/−0.3
2015, [145]30 m34Landsat 5-TM
Landsat 7-ETM
RMSE/bias = 1.2/−0.3 cotton
2016, [98]30 m149Landsat 5-TM
Landsat 7-ETM
RMSE = 0.95 Citric Orchad, grazing, swamps, lakes
2017, [99]30 m19Landsat 7-ETM
Landsat 8-TIRS
RMSE = 1.5 sorghum
2019, [117]30 m22Landsat 5-TM
Landsat 7-ETM
RMSE = 1.6 wheat (dominant crop), barley or cotton in winter
2020, [118]30 m27Landsat 8-TIRSRMSE = 1.4 Almond, tomatoes and maize
2021, [119]30 m42Landsat 7-ETM
Landsat 8-TIRS
RMSE/bias/MAD = 1.2/0.4/0.9winter wheat
* Statistical uncertainties derived from METRIC model after downscalling MODIS resolution to Landsat 30 m scale.
Table 5. Representative chronological compendium of validation studies of the Ts−VI Triangular/Trapezoidal methods. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties Root Mean Square Error (RMSE) and bias are shown for ETd (mm/d).
Table 5. Representative chronological compendium of validation studies of the Ts−VI Triangular/Trapezoidal methods. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties Root Mean Square Error (RMSE) and bias are shown for ETd (mm/d).
Year, ReferenceSRNSensorUncertainty (mm/d)Type of Surface
2001, [154]1000 m6AVHRRRMSE/bias = 3/0.3crop, grass, forest
2003, [155]1000 m26AVHRRRMSE/bias = 1.6/025forest, shrub, wheat, corn, soybean
2006, [156]1000 m15AVHRR
MODIS
RMSE/bias = 1.9/−0.5wheat, cotton
2008, [149]5000 m123SEVIRIRMSE/bias = 1.4/−0.04grazing
2009, [157]1000 m730AVHRRRMSE/bias = 1.2/−0.5 swamp, sugar cane
water
2009, [141]60 m2Landsat 5-TM
Landsat 7-ETM
RMSE/bias = 4/0.8 soybean, corn
2011, [114]1000 m33MODISRMSE/bias = 3/−1.7 wheat, corn
2014, [158]1000 m52MODISRMSE/MAD = 0.9/0.7 grazing, shrub
2017, [100]1000 m248MODISRMSE/bias = 1.4/0.07Pine trees, Green manure–weed– mustard (irrigated), Rice–rice (irrigated), Soybean–wheat
(irrigated), Mixed crops (sugarcane, vegetables, turmeric, maize)
2019, [117]1000 m-MODISRMSE/bias = 3/0.9 to 1.3corn fields, sandy deserts, desert steppe, Gobi Desert, wetlands, and orchards
2021, [119])30 m42Landsat 7-ETM
Landsat 8-TIRS
RMSE/bias/MAD = 1.1/−0.4/0.8 winter wheat
Table 6. Representative chronological compendium of validation studies of the TSEB model. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties Root Mean Square Error (RMSE), bias and Mean Absolute Difference (MAD), are shown for ETd (mm/d). Different uncertainty rows inside the cell means uncertainty associated to the satellite sensor.
Table 6. Representative chronological compendium of validation studies of the TSEB model. Reference, spatial resolution (SR), number (N) of satellite scenes, Sensor Name, Uncertainty and Type of Surface are detailed (uncertainty statistics are averaged values of all type of surface studied). Statistical uncertainties Root Mean Square Error (RMSE), bias and Mean Absolute Difference (MAD), are shown for ETd (mm/d). Different uncertainty rows inside the cell means uncertainty associated to the satellite sensor.
Year, ReferenceSRNSensorUncertainty (mm/d)Type of Surface
2005, [171]60–120 m3Landsat 5-TM
Landsat 7-ETM
RMSE/bias = 1.7/1.4 corn, soybean
2005, [132]90 m1ASTERRMSE/bias = 3 corn, soybean
2007, [133]6–12 m5airborneRMSE/MAD = 2/1.9 shrub, grazing, wheat
2008, [172]30–120 m3Landsat 5-TMRMSE = 1.2 grass, shrub
2009, [141]60 m2Landsat 5-TM
Landsat 7-ETM
RMSE/bias = 2.4/1.3 corn, soybean
2009, [142]60–120 m3Landsat 5-TM
Landsat 7-ETM
RMSE/bias = 0.6/−0.06 corn, soybean
2009, [134]15 m1airbornebias = −0.04 olive, vineyard and citric orchads
2011, [114]1000 m33MODISRMSE/bias = 1.7/0.4 wheat, corn
2012, [173]30 m2Landsat 5-TMRMSE/bias/MAD =
1.9/0.5/1.4
cotton
2012, [174]30 m3Landsat series 5 and 7RMSE/bias = 1.6/0.4corn, soybean
90 m1ASTERRMSE/bias = 2.2/−0.17
Table 7. Same as Table VI but for studies dated from 2013 to 2021.
Table 7. Same as Table VI but for studies dated from 2013 to 2021.
Year, ReferenceSRNSensorUncertainty (mm/d)Type of Surface
2013, [175]30 m5Landsat 5-TM
Landsat 7-ETM
RMSE/bias/MAD =
1.4/−0.2/1.1
corn, soybean
2014, [176]30 m11Landsat 5-TM
Landsat 7-ETM
RMSE/bias/MAD =
1.5/−0.7/1.4
corn, cotton, soybean
2014, [45]90 m7ASTERRMSE = 4wheat, brocoli, frijoles chili, potatoes, chickpea, sunflower, orange and corn
2015, [145]30 m34Landsat 5-TM
Landsat 7-ETM
RMSE/bias = 1/−0.9 cotton
2015, [177]90 m6ASTERRMSE/bias = 1.8/0.17 crops surrounded by desert
2016, [178]0.06 m10DroneRMSE/MAD = 1.8/1.5 Olives orchads
2016, [179]30 m22Landsat 8-TIRSRMSE/bias/MAD =
0.7/0.2/0.5
vineyard
2016, [180]90 m9ASTERRMSE/bias = 2.4/0.5 cron, vegetables and trees
2017, [181]30 m8Landsat 7-ETM
Landsat 8-TIRS
RMSE/bias/MAD =
0.9/−0.1/0.7
forest
2018 [160]30 m5Landsat 5-TMRMSE/bias = 0.2/−0.6 Olives orchads
2020, [182]90 m9ASTERRMSE/bias = 4/3 vegetables, maize and orchard
2021, [119]30 m42Landsat 7-ETM
Landsat 8-TIRS
RMSE/bias/MAD = 1.5/0.3/1.1winter wheat
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

García-Santos, V.; Sánchez, J.M.; Cuxart, J. Evapotranspiration Acquired with Remote Sensing Thermal-Based Algorithms: A State-of-the-Art Review. Remote Sens. 2022, 14, 3440. https://0-doi-org.brum.beds.ac.uk/10.3390/rs14143440

AMA Style

García-Santos V, Sánchez JM, Cuxart J. Evapotranspiration Acquired with Remote Sensing Thermal-Based Algorithms: A State-of-the-Art Review. Remote Sensing. 2022; 14(14):3440. https://0-doi-org.brum.beds.ac.uk/10.3390/rs14143440

Chicago/Turabian Style

García-Santos, Vicente, Juan Manuel Sánchez, and Joan Cuxart. 2022. "Evapotranspiration Acquired with Remote Sensing Thermal-Based Algorithms: A State-of-the-Art Review" Remote Sensing 14, no. 14: 3440. https://0-doi-org.brum.beds.ac.uk/10.3390/rs14143440

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