Next Article in Journal
Application of Machine Learning Models for Fast and Accurate Predictions of Building Energy Need
Previous Article in Journal
Effectiveness of Electricity-Saving Communication Campaigns: Neurophysiological Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Energy-Based Approaches in Estimating Actual Evapotranspiration Focusing on Land Surface Temperature: A Review of Methods, Concepts, and Challenges

1
School of Civil Engineering, College of Engineering, University of Tehran, Tehran 14155-6619, Iran
2
Department of Civil Engineering, University of Ottawa, Ottawa, ON K1N 6N5, Canada
3
Department of Civil Engineering, Iowa State University, Ames, IA 50011, USA
*
Author to whom correspondence should be addressed.
Submission received: 26 December 2021 / Revised: 4 February 2022 / Accepted: 7 February 2022 / Published: 9 February 2022
(This article belongs to the Topic Advances in Clean Energies)

Abstract

:
The surface energy balance (SEB) model is a physically based approach in which aerodynamic principles and bulk transfer theory are used to estimate actual evapotranspiration. A wide range of different methods have been developed to parameterize the SEB equation; however, few studies addressed solutions to the SEB considering the land surface temperature (LST). Therefore, in the current review, a clear and comprehensive classification is provided for energy-based approaches considering the key role of LST in solving the energy budget. In this regard, three general approaches are presented using LSTs derived by climate and land surface models (LSMs), satellite-based data, and energy balance closure. In addition, this review surveys the concepts, required inputs, and assumptions of energy-based LSMs and SEB algorithms in detail. The limitations and challenges of aforementioned approaches including land surface temperature, surface energy imbalance, and calculation of surface and aerodynamic resistance network are also assessed. According to the results, since the accuracy of resulting LSTs are affected by weather conditions, surface energy closure, and use of vegetation/meteorological information, all approaches are faced with uncertainties in determining ET. In addition, for further study, an interactive evaluation of water and energy conservation laws is recommended to improve the ET estimation accuracy.

1. Introduction

Evapotranspiration (ET), which consists of evaporation (physical component) and transpiration (biological component) simultaneously, is defined as the process of water loss from the earth’s surface to the atmosphere. ET is the largest outgoing water flux, consuming more than half of the absorbed solar radiation and converting 60% of precipitation into water vapor [1]. Therefore, an accurate estimate of temporal and spatial changes in ET is considered an essential prerequisite for a broad range of applications such as agricultural management, drought monitoring, water use efficiency, atmospheric circulation and weather forecasting, energy and food security, carbon balance and plant growth, and hydrological modeling.
A common classification for the determination of ET includes direct and indirect methods. Techniques such as the lysimeter, sap flow (SF), eddy covariance (EC), Bowen ratio (BR), and scintillometer measure ET directly while indirect methods employ specified equations to estimate the latent heat flux [2]. Direct methods cannot be suitable spatial representatives for large-scale ET mapping due to costly setup and maintenance, flow distortion around the sensors, and point measurements [3,4,5]. In contrast, indirect methods, those that benefit from physical concepts for estimating ET, are capable of providing the spatial distributions of ET with an acceptable accuracy over large areas.
Indirect methods of ET estimation include mass transfer (MT) models, aerodynamic models, resistance-based models, radiation or temperature-based models, and energy balance models [6,7]. Mass transfer models, which estimate ET based on the difference in vapor pressure between the surface and reference height, are unsuitable for canopies with long-term water stress. Aerodynamic methods parameterize ET using the vertical gradient of the vapor pressure and eddy diffusivity. These methods are able to provide accurate estimates of ET using conventional data sources; however, considerable errors occurring in heterogeneous conditions restrict their wide employment. By relating the water vapor transport to the vapor pressure gradient and resistance network, resistance models were formed to estimate ET for both leaf and canopy. At the canopy scale, the method may be inappropriate due to assuming an identical location for sources and sinks of heat and vapor transfer. The models based on radiation or temperature data are popular empirical methods that are mainly used to estimate potential ET. Finally, Surface Energy Balance (SEB) models, which are the most popular physically based concept for simulating land surface processes including evapotranspiration, estimate ET based on the energy conservation law. Many complex algorithms have been developed for forecasting different parts of the energy balance [8]. SEB models rest on the assumption that the available surface energy is partitioned between heating the air and soil, transferring water vapor from the evaporating surface to the atmosphere, as well as heat storage for non-advective conditions. SEB algorithms and LSMs are governed by the land surface temperature (LST) as a significant boundary condition and driving force, and different solutions of the SEB equation can be presented using different retrieval methods of LST for estimating latent heat flux. Therefore, the current review was designed to provide a comprehensive assessment of energy-based models with a view to highlighting the key role of LST in the classification of these models.
In general, three types of approaches have been classified to estimate actual ET using LST and SEB. As the first approach, Penman–Monteith (PM) is a single-layer model combining energy balance and mass transfer concepts [9,10]. The PM equation follows the big leaf assumption, where the plant canopy is modeled as a big leaf to parameterize the transpiration through canopy stomatal conductance [11,12,13,14]. Given the changes in leaf distribution due to the variations in leaf size and density, plant height, and leaf angle, this assumption leads to unrealistic modeling of the canopy structure [11]. On the other hand, the variable of land surface temperature in the PM model is eliminated, and instead, ET is estimated using the slope of the saturation vapor pressure curve at air temperature. Despite the aforementioned assumptions, the PM model is widely used because of its simple nature and physically sound basis, and it was proven that the model is sufficient for simulating the ET of densely vegetated canopies [15]. TOPUP [16] and PROMET [17] are examples of one-layer PM-based models. For sparse canopies, the PM model tends to neglect the evaporation from the soil due to the big leaf theory [13], especially in models of the mosaic type. Therefore, two-layer models were introduced by Shuttleworth and Wallace [18], in which mass and heat interactions between soil and vegetation are used to estimate evaporation from bare soil and transpiration from plants and moist living surfaces. Being more complicated, multi-layer models, which have more than one canopy layer, compute the resistance network separately for discrete layers within the canopy [14,15]. These models are more flexible and capable of providing accurate ET than the two previous models under different weather and crop growing conditions. Although the first approach enjoys noticeable popularity, its accuracy can decrease because of the LST elimination. To address this issue, continuous LSMs can be used to solve the mass and energy balances simultaneously [19]. In these models, the energy balance equation in which all energy fluxes depend on LST is solved by a numerical method such as the Newton–Raphson method to find the LST that closes the equation [20]. ET can be subsequently estimated by the computed LST and MT equation, which uses the mean values of surface layer variables and turbulent transfer principles [21]. There are some examples for the application of LSMs in the literature, including the calibration of the Variable Infiltration Capacity (VIC) model for improving the accuracy of ET estimates [22], the solving of energy and water balances using the TOPLATS model [22,23,24,25], determination of the hydraulic properties of soil with the Noah land surface model [26], and validation of the Flash-Flood Event-Based Spatially Distributed Rainfall-Runoff Transformation Energy Water Balance (FEST-EWB) model using satellite LST [27,28].
The last approach calculates ET as a residual of the SEB equation (after the acquisition of all other energy fluxes) or a fraction of potential/reference ET. In this approach, LST information is derived from satellite thermal infrared (TIR) data, which makes it the most widely applied method to provide ET estimates with a relatively high degree of accuracy for an extended range of purposes (such as irrigation management [29], drought analysis [30], agricultural irrigation water requirement [31], water accounting [32,33,34], and water productivity [35]). However, the application of the approach is limited by cloud contamination, and TIR remote sensing-based LST retrievals under cloudy conditions cause discontinuities in ET calculations. Models using the TIR LST-based approach are divided into (1) one-source models, which assess vegetation and soil as a single component of the energy budget, and (2) two-source models, which assess the energy budget of vegetation and soil separately. One-source models include the Surface Energy Balance System (SEBS) [5], Mapping Evapotranspiration at High Resolution with Internalized Calibration (METRIC) [29], Surface Energy Balance Algorithm for Land (SEBAL) [36], Simplified Surface Energy Balance Index (S-SEBI) [37], and Operational Simplified Surface Energy Balance (SSEBop) [38,39]. Two-source models include the Two Source Energy Balance (TSEB) [40], Atmosphere–Land Exchange Inverse model (ALEXI) [41], Dual Temperature Difference (DTD) [42], and Enhanced Two-Source Evapotranspiration Model for Land (ETEML) [43].
ET simulation is faced with numerous challenges, both conceptually and practically. Firstly, the SEB closure and its effect on quantifying ET are discussed. During the past decades, it became evident that the SEB equation could not be closed completely [44,45]. In most cases, especially across EC sites, the available energy was found to be larger than the sum of turbulent fluxes (i.e., sensible and latent heat) [44,46]. The difference between available energy and turbulent fluxes is called the surface energy imbalance or the SEB closure problem, which can be 10% to 30% of the available energy (e.g., [47,48,49,50,51,52,53]). The underlying reasons for the surface energy imbalance can be classified into four groups: instrumental and measurement/computation errors such as systematic errors of soil heat and net radiation fluxes, errors related to data processing and analysis such as correction algorithms, sub-mesoscale transport processes, and neglected additional sources of energy (canopy heat storage, air heat storage, etc.) [54].
The second challenge is that of parameterization of the resistance network including the surface and aerodynamic resistances for vegetation and bare soil, and the evaluation of their impacts on the ET process. Canopy resistance (or surface resistance) is influenced by agronomical and climatological variables such as soil water content, canopy structure, etc. [55,56,57]. Since surface resistance is difficult to model mathematically and physically, empirical equations have received significant attention [55]. Although empirical methods need to be calibrated, many researchers have recommended them for practical use [55,57,58,59,60]; however, it was pointed out that empirical formulas lead to noticeable uncertainties in estimating resistance parameters [61]. The most common way of determining the canopy resistance has been presented in the Jarvis-type models, which consider environmental factors such as the air temperature, soil water condition, and vapor pressure of the atmosphere [62,63,64]. In addition to surface resistance, aerodynamic resistance to water and heat transfer also plays a crucial role in determining the latent heat flux [65]. Therefore, it must be computed considering different atmospheric and surface characteristics such as surface roughness, wind speed profile, atmospheric stability condition, air temperature, etc. Various methods ranging from the simplest (just the function of wind speed) to more sophisticated and accurate methods have been developed to calculate the aerodynamic resistance [66,67,68,69,70]. In this respect, the formulations by Thom [71], Brutsaert [72], and Monteith and Unsworth [73], which use a logarithmic form of the wind speed profile, are the most common methods.
This paper reviews the energy-based approaches used in the estimation of ET with a particular emphasis on the quantification of the land surface temperature. In this regard, the SEB algorithms, energy-based LSMs, and Penman–Monteith approach, along with related concepts and assumptions, are discussed in the following sections. In addition, this review highlights the main limitations and challenges by which the application of energy-based models is restricted. At the end, an initiative perspective on estimating ET is presented using water–energy systems. It should be noted that lists of acronyms and symbols are represented in Appendix A, and solutions of ET estimation approaches are provided in Appendix B. Figure 1 represents the flow diagram of the current review.

2. Materials and Methods

As mentioned, this paper reviews three energy-based approaches in terms of the estimation of ET with a particular emphasis on quantifying the land surface temperature. In this section, the Penman–Monteith approach, energy-based LSMs, and SEB algorithms are discussed.

2.1. Approach 1: Penman–Monteith (PM)-Based MODELS

The Penman equation, which is a semi-empirical formula developed by Penman [9], estimates the evaporation from open-water surfaces based on the mass transfer and energy balance concepts using monthly weather data, including solar radiation, air temperature, wind speed, and vapor content. The original Penman equation is defined as:
E = Δ × L E n + γ × L E a Δ + γ
where E is the evaporation rate (kg/m2s), L E n is the evaporation rate due to net radiation (kg/m2 s), Δ is the slope of the saturation vapor pressure curve at air temperature T (kPa/°C), γ is the psychrometric constant (kPa/°C), and L E a is the evaporation rate due to mass transfer (kg/m2s), which is obtained from an empirical mass transfer formula. The Bucket model [74] is the earliest attempt to model surface processes, employing the single-source Penman approach based on Budyko’s concept [75]. This model overestimates evaporation from vegetated and bare soil surfaces; however, some modifications can be applied to reduce the biases of the model [76,77]. By considering physical principles for estimating LEa over soil and vegetated surfaces, Monteith [10] improved the Penman model by introducing canopy and aerodynamic resistances, and the popular Penman–Monteith equation was formed. The combination of aerodynamic-based ET (representing the mass transfer concept) and SEB-based sensible heat flux (representing the energy budget concept) through the Bowen ratio ( β ) is the basis of the PM model, as shown below:
β = H L E = ( R n G L E ) ρ a C p ( e o e a ) γ ( r a + r s )
where H is the sensible heat flux (W·m−2), LE is the latent heat flux (W·m−2), R n is the net radiation flux (W·m−2), G is the soil heat flux (W·m−2), ρ a is the air density (Kg·m−3), C p is the specific heat capacity of air at constant pressure (J·kg−1·K−1), r s is the surface resistance (s·m−1), r a is the aerodynamic resistance to heat and/or vapor transport (s·m−1), e a is the actual vapor pressure of air (kPa), and e o represents the vapor pressure within the leaf (kPa), which is assumed to be equivalent to the saturation vapor pressure corresponding to the LST [6]. H is obtained from Equation (3), and by replacing it and arranging Equation (2), LE can be calculated using Equation (4):
H = ρ a C p ( L S T T a ) r a
L E = R n G 1 + β = R n G 1 + γ ( r a + r s ) ( L S T T a ) ( e o e a ) r a
where T a is the air temperature (K).
The main challenge for solving Equation (4) is related to land surface temperature information. Since the accurate measurement of the land surface temperature was not possible in the 1960s, the air surface temperature gradient could not be used to compute the sensible heat flux. Hence, a psychrometric approximation, which is represented by Equation (5), was implemented to eliminate LST from the PM equation, as shown in Equation (6) [78]:
Δ = e o e a L S T T a = ( e o e a ) ( e s e a ) L S T T a
L E = R n G 1 + β = R n G 1 + γ Δ [ 1 ( e s e a ) ( e o e a ) ] ( r a + r s ) r a
According to the aerodynamic expression for LE, the relationship given by Equation (7) can be replaced in Equation (6), and the final PM equation is written as Equation (8):
e o e a e s e a = L E L E a r a + r s r a
L E = R n G 1 + β = R n G 1 + γ Δ [ 1 ( e s e a ) ( e o e a ) ] ( r a + r s ) r a
where T a is the air temperature (K), and e s is the saturation vapor pressure at Ta (kPa).
The PM model is known as a robust approach in providing ET estimates due to a number of reasons: (1) it defines an energy limit on the ET rate; (2) it theoretically uses a combination of main factors affecting the ET process; (3) it provides both the simulation and prediction of ET successfully; and (4) the obtained ET is not sensitive to input data. Despite the theoretical appeal of the PM model, its application is restricted by the need for ground-based measurements such as humidity, wind speed, air temperature, surface roughness parameters, and vapor pressure deficit [79]. In addition, temporal upscaling from an instantaneous scale to a longer time scale presents PM-based models with some challenges under cloudy weather conditions [80]. The PM model follows a one-layer or big leaf concept in which the canopy is modeled as a big leaf and the parameterization of heat and vapor transport is implemented using single-surface/aerodynamic resistances. FAO-56 from the Food and Agricultural Organization is an important application of the big leaf approach [81]. In addition, the TOPUP model [82,83] benefits from a simplified big leaf approach with limited parameters in which ET can occur from three main pathways: the interception store, the root zone store, and the groundwater table. Although the TOPUP model is widely used, the accuracy of its results decrease due to the simplified structure, big leaf approximation, and ignoring of the impact of land surface temperature on the ET process [83]. Another model employing the big leaf approach is Process Oriented Models for Evapotranspiration (PROMET), developed by [17]. PROMET includes a kernel model for treating the ET process and a spatial data modeler for adjusting the spatial scale of data to that of the model. The spatial data modeler is the reason that the PROMET model is capable of estimating ET on different scales, from a field to a landscape scale. The Moderate Resolution Imaging Spectroradiometer (MODIS) ET algorithm (or MOD16 algorithm) improved the performance of the PM model by including the environmental stress factors into the canopy resistance [79]. In addition, Mu et al. [84] revised the MODIS ET model by accounting for nighttime ET and dividing the canopy into dry and wet surfaces (i.e., E = Ewet (canopy interception) + Es (soil evaporation) + Ec (plant transpiration)). Further, a Jarvis–Stewart-type canopy resistance was introduced by Zhang et al. [85] based on LAI and environmental factors. One of the significant revisions for the PM model is the Penman–Monteith–Leuning (PML) method that considers the impacts of the air humidity deficit and radiation through biophysical modeling [86]. PML estimates plant transpiration via the PM equation, soil evaporation via the Priestley–Taylor equation, and canopy conductance via the Leuning model [87,88]:
L E = Δ × R n , c + ( ρ C p / γ ) V P D g a Δ + 1 + g a / g c + f Δ ( R n , s G ) Δ + 1
where R n , c is the net radiation for canopy (W/m2), R n , s is the net radiation for soil (W/m2), g a is the aerodynamic conductance (m/s), g c is the canopy conductance (m/s), VPD is the vapor pressure deficit (kPa), and f is a factor related to the potential evaporation of the soil surface (dimensionless).
The big leaf method was reported to be suitable only for describing ET from fully vegetated surfaces [73]. For sparse canopies, where the surface can be separated into vegetated and bare soil parts, the one-source PM model tends to fail due to neglecting the soil evaporation. Therefore, a patch (or tile) approach, which was first presented by Avissar and Pielke [89], was used to distinguish fluxes of soil evaporation and plant transpiration with no interaction. One example of a patch model is the Interaction Soil–Biosphere–Atmosphere (ISBA) model, which uses a patch approach to partition the soil and canopy components, and estimates ET as a sum of the soil evaporation, plant transpiration, and interception [90]. Another example is the Surface Energy and Water Balance (SEWAB) model, in which the total latent heat flux is divided into bare soil evaporation and plant transpiration via a vegetation fraction and the component ET is estimated by coupling the water balance and surface energy budget [91]. For more heterogeneous surfaces, a multi-patch approach can be applied to divide the grid into multiple uncoupled patches [92]. However, some studies discovered the interaction between canopy and soil fluxes [93,94], which led to the formation of two-source models.
Two-source models include a canopy layer in addition to the soil layer, and the heat and vapor fluxes can interact between the components. Unlike the patch approach that considers the heterogeneity along the surface, two-layer models account for the vertical heterogeneity of components. Three types of resistance, besides the aerodynamic and surface resistances of single-layer models, are included in two-layer models, which regulate the heat and vapor transfers between the mean canopy airflow and canopy surface, the mean canopy airflow and soil surface, and the mean canopy airflow and a standard height. Determining the resistance network of two-layer models is complex due to the over-parameterization and the need for their calibration [83,95]. Therefore, multi-objective approaches were suggested to implement the models when prior information is limited [96]. The model introduced by Shuttleworth and Wallace [18] and the ETLook algorithm [97] are examples of the two-layer approach, which solve the PM equation for soil and canopy separately:
E = Δ   ( R n , s G ) + ρ c p ( e s e a ) r a , s 1 Δ + γ ( 1 + r s , s r a , s )
T r = Δ   ( R n , c ) + ρ c p ( e s e a ) r a , c 1 Δ + γ ( 1 + r s , c r a , c )
where E is evaporation (W·m−2), Tr is transpiration (W·m−2), and the s and c indices indicate the soil and canopy, respectively.
To consider the vertical canopy structure in detail, especially for a tall-vegetated ecosystem, multi-layer models were developed by extending the canopy structure to more than two layers (e.g., [98,99]). These models estimate the heat and water fluxes for each layer separately, and canopy-related fluxes are integrated over the canopy height. An example of a multi-layer structure is the Stanghellini model [100], which estimates the ET of greenhouse crops by including the LAI impacts within multiple layers of the canopy (Equation (12)).
L E = Δ R n e t + ( 2 L A I ρ c p V P D ) / r e γ ( 1 + ( Δ γ ) + ( r i r e )
where r i is the canopy’s internal resistance (s/m), and r e is the canopy’s external resistance (s/m).
The remote-sensing-based Ecosystem–Atmosphere Simulation Scheme (EASS) model is another multi-layer land surface model that adopts the PM equation for the estimation of components of ET by using a two-leaf strategy, which improves the accuracy of energy fluxes by almost 10% compared to the one-leaf strategy [101]. A modified version of EASS, the Dynamic Land Model (DLM), uses the PM equation to estimate ET for three components:   E c (evaporation from water intercepted by canopy), E s (evaporation and sublimation from the soil), and T r c (canopy transpiration) [101,102]:
To obtain E c and T r c , the DLM model adopts a leaf stratification strategy similar to EASS in which LAI is calculated for shaded and sunlit leaves separately [101]. It is worth noting that the turbulence within the canopy and also time-consuming computational processes with detailed data requirements (such as crop physiology and crop architecture information) limit the application of the multi-layer models at regional scales [15,98]. Figure 2 shows the schematic diagrams of the models discussed in the current section, i.e., the Penman model, the PM model (one-layer model), the patch model, the two-layer model, and the multi-layer model.

2.2. Approach 2: Land Surface Models (LSMs)

Since the late 1980s, advanced energy-based LSMs have been introduced to estimate water content, vegetation and soil temperature, and biogeochemical fluxes using sophisticated equations. Input data for LSMs include meteorological data (air temperature, air humidity, air pressure, shortwave and longwave radiation, wind speed and direction, and precipitation) and soil/vegetation information [103]. LSMs can calculate soil evaporation, plant transpiration, and evaporation from the water intercepted by the canopy using the PM and MT equations [104]. MT-based models relate ET to the vertical gradient of the vapor pressure and eddy diffusivity [105], and they are typically used under neutral atmospheric conditions. However, these models can be applied under unsteady conditions using atmospheric stability functions mainly quantified by Monin–Obukhov Similarity Theory (MOST) [106]. To solve the MT equation, the land surface temperature must be known, which is directly obtained from the energy balance equation at the surface. In other words, radiative and turbulent heat fluxes are written as functions of LST, and surface temperature values (known as the Equilibrium Energy Balance Temperature, EEBT) are calculated by closing the energy balance at the surface (Equation (13)):
R n ( E E B T ) G ( E E B T ) H ( E E B T ) λ E T ( E E B T ) = C H δ T c Δ t
where C s is the surface heat capacity, and δ T c is the change in the land surface temperature of the system over time Δ t . Some of the main LSMs based on the direct solution of the energy balance at the land surface are discussed below:
Biosphere–Atmosphere Transfer Scheme (BATS)
In 1981, a Soil–Vegetation–Atmosphere Transfer (SVAT) scheme was developed using the approaches of [107] in formulating the soil temperature and vegetation interception [108]. Further, Dickinson [109] improved the 1981 version by considering the conceptual resistance network and physical soil hydrology scheme, which made it appropriate for use in large-scale atmospheric models. In 1986, a notable improvement concerning the assignment of soil type and land cover was applied using a popular available global land cover dataset [110]. This modified version was called the Biosphere–Atmosphere Transfer Scheme (BATS), which could be implemented in both on-line and off-line ways [111].
The BATS scheme includes three soil layers and one vegetation layer, accounting for three surface temperatures of the canopy, surface soil, and subsurface soil layer. In the BATS model, a demand–supply approach is utilized to calculate evaporation from the soil surface ( E a ) as the minimum of the E p   ( potential   evaporation ) and E 0 (diffusion-limited maximum evaporation) [109].
On the other hand, the canopy evapotranspiration includes two parts of transpiration from dry foliage and evaporation from wet foliage. It should be noted that the BATS model considers vegetation as a uniform and porous surface with zero heat capacity, and neglects photosynthetic/respiratory energy transfer. The wet and dry canopy evaporation values are obtained from Equations (14) and (15), respectively:
E d e w = ρ a f c w [ q s ( T c ) q s ] r a
E t r = min ( E t r , d e m , E t r , s u p )
where ρ a is the air density, q s is the saturation specific humidity, T c is the canopy temperature, f c w is the fractional area of the canopy occupied by water, E t r , d e m is the transpiration by atmospheric demand, and E t r , s u p is the water supply by roots.
Simple Biosphere Model (SiB)
The Simple Biosphere Model (SiB) was developed by Sellers et al. [112] based on a land surface parameterization design originally used to compute exchanges of heat, water, and momentum. The land surface parameterization scheme utilizes two distinct vegetation layers so that the upper layer includes the perennial canopy of shrubs or trees and the lower layer includes the ground cover of grasses and other herbaceous plants. In addition, the physical and physiological characteristics of the vegetation layers, which determine the aerodynamic transfer of sensible heat, water vapor, and momentum from the vegetation and soil components to a reference level, govern the interception of radiation and the transfer of moisture through the atmosphere–soil–plant system [112]. SiB calculates the variables of evapotranspiration and surface temperature for layers as follows:
  • Two temperatures (i.e., the temperature of the canopy vegetation, T c   ,   and the temperature of both the ground cover and the soil surface, T g s ):
    C c T c   t = R n , c H c λ E c
    C g s T g s   t = R n , g s H g s λ E g s
  • The evapotranspiration from the canopy, E c , has two parts: (1) E w c , evaporation from the wetted fraction of the canopy, and (2) E d c , transpiration of the soil water extracted by the root zone and water lost from the dry fraction of the canopy.
    λ E w c = [ e ( T c ) e a ]   r ¯ b ρ c p   γ W c    
    λ E d c = [ e ( T c ) e a ]   r ¯ c + r ¯ b ρ c p   γ ( 1 W c )  
  • The evapotranspiration from the ground cover and soil surface, E g s , has three parts: (1) E w g and (2) E d g , which correspond to E w c and E d c for the ground cover, and (3) E s , direct evaporation from the soil surface.
    λ E w g = [ e s ( T g s ) e a ]   r d ρ c p   γ W g V g    
    λ E d g = [ e s ( T g s ) e a ]   r g + r d ρ c p   γ ( 1 W g ) V g
    λ E s = [ f h e s (   T g s ) e a ]   r s u r f + r d ρ c p   γ ( 1 V g )
    where the c and gs indices indicate the canopy and ground cover/soil surface, T is the surface temperature (K), C is the heat capacity (J/K·m2), λ is the latent heat of vaporization (J/kg), r ¯ b is the bulk boundary layer resistance (s/m), r d is the aerodynamic resistance between the ground and canopy airflow (s/m), r ¯ c is the bulk stomatal resistance of the upper story vegetation (s/m), r g   is the bulk stomatal resistance of the ground vegetation (s/m), r s u r f is the bare soil surface resistance (s/m), f h is the relative humidity within the pore space of the surface soil layer, V g   is the fractional cover of the ground cover, and W c and W g are the wetness fractions of the canopy and the ground cover.
Mosaic Land Surface Model
The Mosaic LSM uses the “mosaic” strategy in which a grid square can be separated into homogeneous sub-areas, i.e., tiles of the mosaic, which contain a single bare soil or vegetation type [113]. This characteristic allows the model to consider the sub-grid heterogeneity and coexisting of different vegetation types in a grid square area. The various types of surfaces that can be included in each tile include the following: (1) broadleaf deciduous trees; (2) broadleaf evergreen trees; (3) broadleaf shrubs; (4) grassland (groundcover); (5) dwarf trees (tundra); (6) needle-leaf trees; (7) desert soil; and (8) bare soil. If two tiles represent the same surface properties, they can be aggregated together into a single tile. Parameterizing the energy balance of each tile is analogous to that of a single vegetation version of the SiB model; however, it can be simplified to be written in PM form. The interaction between the tiles is only established through the atmosphere, and also each tile can be coupled to the atmosphere independently. It is worth noting that the concept of the coupling strategy for the Mosaic model is different from that for models that assume a homogeneous mixture of vegetation types within a grid square (such as SiB). Mosaic LSM eschews the simultaneous computation of ground cover transpiration and other fluxes via the removal of the interactive ground cover vegetation pathway for evaporation. Finally, the total evaporation rate, E, can be written as [113]:
E = ρ ϵ   p s [ e ( T s ) e a ]   r e f f + r a
where e s ( T s ) is the saturation vapor pressure at the surface temperature of the canopy and soil, which is derived by averaging the variables over the tiles for a grid cell, ϵ is the ratio of the molecular weight of the water vapor to that of dry air, p s is the surface pressure, r e f f is the single effective resistance composed of the r c , r d , and r s u r f resistances, and r a is the aerodynamic resistance between the canopy airflow and the reference height.
Variable Infiltration Capacity (VIC) Model
The Variable Infiltration Capacity (VIC) model is a macroscale land surface model that uses a mosaic scheme to represent the land cover classes within a grid cell. By solving the full energy and water balances, this model simulates land–atmospheric fluxes for each grid independently. Key features of the VIC model are its applicability for continental scale and its open-source availability. Early VIC models included two soil layers with a resolution of one or more degrees per grid cell. Later, Liang et al. [114] found that adding a thin top layer (5–15 cm) significantly improved the results of evapotranspiration in arid climates. Even though most applications use three layers, this version of the model is able to use an arbitrary number of soil layers [115].
In addition to plant transpiration and soil evaporation, which are parameterized by PM formulation, VIC simulates evaporation from the water intercepted by the vegetation. The evaporation from the canopy layer of the nth class, E c [ n ] , is [116]:
E c [ n ] = f [ n ] × E c [ n ]
where n is surface cover class index, E c [ n ] is maximum canopy evaporation, and f [ n ]   is the fraction of the time step required for the evaporation from the water intercepted by the canopy. The transpiration during the time step is then:
E t [ n ] = ( 1 f [ n ] ) × E p [ n ] r a [ n ] r a [ n ] + r 0 [ n ] + r c [ n ] + f [ n ] × [ 1 ( W i [ n ] W i m [ n ] ) 2 / 3 ] × E p [ n ] r a [ n ] r a [ n ] + r 0 [ n ] + r c [ n ]
where E p is the potential evaporation (which is calculated by the PM equation with zero canopy resistance [117]), r 0 [ n ]   is the architectural resistance, W i [ n ] is the intercepted water amount, and W i m [ n ] is the maximum amount of intercepted water. The first term indicates the fraction of the time step for which no evaporation occurs from the intercepted water, and the second term indicates the fraction of the time step for which both plant transpiration and evaporation from the interception occur.
Bare soil evaporation only occurs from the first layer ( E 1 ) , meaning that the evaporation from the second layer is assumed to be zero (i.e., E 2 = 0). When the first soil layer is saturated, the evaporation occurs at the potential rate E p [ N + 1 ] .
In addition, when the first layer is unsaturated, it evaporates at a rate E 1 which varies corresponding to the topography, infiltration, and soil characteristics. E 1 is also calculated via the formulation provided by Franchini and Pacciani [118], which follows the structure of the Xinanjian model [119]. It is noteworthy that for finding the land surface temperature at each time step, similar to other MT-based LSMs, the surface energy balance equation is iteratively solved using a trial-and-error strategy.
Noah Land Surface Model
The original OSU LSM included two soil layers and it used the thermal conduction equation and Richardson’s equation for soil temperature and soil moisture, respectively [120]. Later, it was enhanced to include more soil layers (from two to four layers), and the formulation of the canopy resistance was also improved [121]. As an outcome, Noah LSM was developed by the National Centers for Environmental Prediction (NCEP) and several agencies [122]. According to this model, the surface temperature is obtained from a single linearized SEB equation representing the combined vegetation/soil surface [120]. In addition, the soil heat flux is determined via the diffusion equation for the ith soil layer temperature, T i [121]:
Δ z i C i T i   t = ( K t T   z ) z i + 1 ( K t T   z ) z i
where z is the depth of the soil layer, C is the volumetric heat capacity, and K t is the thermal conductivity (C and K t are calculated by volumetric soil water content). The prediction of T i is performed using the fully implicit Crank–Nicholson scheme.
The total evaporation, E, is considered to be the sum of the direct evaporation from the top shallow soil layer, E dit , the evaporation from the interception storage of the canopy, E c , and the transpiration from the roots and the canopy, E t [121].   E dit is calculated by Equation (27):
E dit = ( 1 σ f )   M I N ( [ D   ( Θ z ) z 1 K z 1 ] , E p )
where Θ is the volumetric soil water content, D is the soil water diffusivity, E p is the potential evaporation, and σ f is the fraction of green vegetation. E dit can proceed at the potential rate E p when the soil surface is rather moist. Otherwise, direct evaporation occurs at the rate by which the top soil layer can transfer water upward from below.
Canopy evaporation from intercepted water and plant transpiration values are obtained from Equations (28) and (29), respectively:
E c = σ f E p ( W c S ) n
E t = σ f E p B c   [ 1 ( W c S ) n ]
where W c is the maximum wetness fraction of canopy S, and B c embodies the canopy resistance, including the soil moisture stress. The factor ( W c / S ) n   is considered to be a weighting coefficient to suppress E t in favor of E c due to the canopy surface becoming increasingly wet. Table 1 includes a summary of the characteristics of MT-based LSMs that follow approach 2.

2.3. Approach 3: Surface Energy Balance (SEB) Models

Approach 3 includes the residual methods and evaporative fraction (EF)-based models, which estimate ET using surface–air temperature difference or absolute LST. In residual methods, ET is obtained as a remainder term of the SEB equation, i.e., by subtracting sensible and soil heat fluxes from net radiation, a residual energy flux is achieved that is called ET. The surface energy balance for estimating ET is of the form:
λ E T = R n G H  
where λET (or LE) is the latent heat flux (W·m−2), and λ indicates the latent heat of vaporization (J·kg−1 °C−1). The SEBAL, METRIC, SEBS, TSEB, TSTIM, ALEXI, DTD, and HTEM models are indexed under this classification. On the other hand, EF-based models use an evaporative fraction to estimate the latent heat flux, including the SEBI, S-SEBI, SSEB, SSEBop, TTME, and ETEML models.
Similar to approach 2, this approach also requires LST data as a key prerequisite to quantify the energy fluxes. Satellite TIR sensors, such as AVHRR (Advanced Very High Resolution Radiometer), MODIS (Moderate Resolution Imaging Spectroradiometer), ASTER (Advanced Spaceborne Thermal Emission and Reflection Radiometer), and Landsat, provide temporal and spatial distributions of surface temperature data with an acceptable accuracy. Different methods have been proposed to derive LST from TIR data, which are explained in subsequent sections, as well as SEB models that employ these TIR-derived LSTs for estimating ET and other energy fluxes.

2.3.1. Retrieval of LST from Satellite TIR Observations

LST information for this approach is obtained from satellite TIR observations. Since the LST estimation from the radiation of the TIR spectral region depends on atmospheric effects and emissivity [123,124], the extraction of LST requires corrections for emissivity, as well as radiometric and atmospheric effects [124]. Various algorithms have been developed to extract the LST information from the sensors using the radiative transfer equation (RTE), including single-channel methods, multichannel methods, multitemporal methods, and multiangle methods.
The Single-channel (or mono-channel) method, which was proposed by Hook et al. [125], estimates LST using a corrected radiance measured in a single channel through the inverted RTE [125,126,127,128,129,130]. The accuracy of the method is primarily limited by the atmospheric profile, surface emissivity, and corrections for the topographic and atmospheric effects [131,132,133]. However, the accuracy of these methods is acceptable in retrieving LST if the emissivity and atmospheric profile are known.
The Multichannel method, unlike the single-channel method in which the land surface emissivity and atmospheric profiles must be known at the satellite overpass, estimates LST without the need for additional information. For the first time, McMillin [134] proposed a split-window algorithm to calculate the surface temperature of sea by the differential atmospheric absorption in the two channels. Since then, many studies attempted to use different split-window algorithms for retrieving LST (e.g., [135,136,137,138,139,140,141,142,143,144,145]), which include linear and nonlinear split-window algorithms, linear or nonlinear multichannel algorithms, and the temperature emissivity separation method. A linear algorithm utilizes a simple relationship between the land surface temperature and the two brightness temperatures corresponding to the two TIR channels [134]. Given the error of linear methods due to the linearization of RTE, nonlinear approaches were developed based on a nonlinear relationship between LST and the top-of-atmosphere brightness temperatures [141,144,146,147,148]. By considering more than two TIR channels, LST can be also obtained from a linear or nonlinear combination of brightness temperatures similar to split-window algorithms [149,150]. In comparison with two-channel algorithms, multichannel algorithms are more accurate in retrieving LST [151]; however, adding more channels is associated with an increase in measurement errors.
The Multiangle Method, which is based on differential atmospheric absorption, employs various viewing angles to reduce the effect of atmospheric water vapor on LST [142,152,153]. Sobrino and Jiménez-Muñoz [154] compared a dual-angle algorithm with a nonlinear split-window algorithm in terms of the accuracy of retrieved LSTs, and the results indicated that the dual-angle algorithm was more appropriate than the split-window algorithm. However, it should be pointed out that multiangle methods have angular dependence on the LST and emissivity [155]. In addition, different viewing angles may provide different specifications for an object, and therefore, this method is not a suitable approach for unstable atmospheric conditions and heterogonous areas.
The Multitemporal Method [156] employs different times to estimate LST and includes the two-temperature and the physics-based day/night operational methods. The two-temperature method utilizes multiple observations to reduce the unknowns so that the number of measurements is equal to that of the unknowns. The main assumption of this method is that changes in the land surface emissivity are considered constant over time. Although the method is straightforward with a simple formulation, solutions may be unstable due to high-correlated equations [156,157]. Similar to multiangle methods, this method can only be employed in homogenous areas [158]. The physics-based day/night operational (D/N) method was developed to retrieve LST from day and night TIR data [159]. In fact, this method is a modified version of the two-temperature method, which utilizes two time observations to retrieve LST and land surface emissivity simultaneously. To improve the accuracy of the LST estimation, two variables, air temperature and atmospheric water vapor, are considered to adjust initial atmospheric profiles. Even though the D/N method is more accurate than other methods in retrieving the LST quantities, it suffers from misregistration problems and variations of the viewing zenith angle. To overcome the misregistration issues, the pixels can be aggregated into coarser spatial scales [158]. In addition, the combined use of night and daytime observations with a larger weight on the daytime information and the combination of the viewing angle-based method, split-window, and the D/N algorithm can be implemented to modify the LST retrievals.

2.3.2. SEB Algorithms

The SEB models consist of two categories, one-source (or single-source) and two-source (or dual-source) models, which solve the SEB equation using TIR LST to estimate ET as the residual of the difference between the available energy (Rn-G) and the loss due to the sensible heat flux or as a fraction of potential/reference ET. One-source SEB models analyze the soil and canopy by using a combined energy budget while two-source models are capable of partitioning the radiometric surface temperature into soil and canopy components to estimate soil evaporation and plant transpiration. Accordingly, a large number of SEB models have been developed, which are discussed in detail in the following sub-sections.

One-Source (or One-Layer) Models

Surface Energy Balance Index (SEBI)
Choudhury and Menenti [160] developed the Surface Energy Balance Index (SEBI) based on an evaporative fraction to compute the relative ET using the crop water stress index [161,162] redefined by theoretical pixel-wise ranges for LST and LE. In this method, upper and lower limits are defined based on the difference between the air and surface temperature. Evaporation is assumed to be zero in dry conditions, and total available energy (with a maximum value of surface temperature) is then assigned to the sensible heat flux. On the other hand, the minimum LST under wet conditions is obtained from the potential ET (PET) using the Penman–Monteith equation with the assumption of zero internal resistance. Finally, the relative latent heat flux is calculated from the equation given below [163]:
L E = L E p   ( 1 ( L S T T p b l ) × r a 1 ( L S T m i n T p b l ) × r a , m i n 1 ( L S T m a x T p b l ) × r a , m a x 1 ( L S T m i n T p b l ) × r a , m i n 1 )
where L E is the latent heat flux (W/m2), L E p is the potential latent heat flux (W/m2), L S T is the land surface temperature (K), T p b l is the average temperature of the planetary boundary layer (K), L S T m i n and L S T m a x are the minimum and maximum land surface temperatures, respectively, ra represents the aerodynamic resistance (s/m), and ra,min and ra,max are the minimum and maximum aerodynamic resistances (s/m), respectively. It should be kept in mind that the pixel-wise conception of SEBI is employed in other models, i.e., SEBAL, SEBS, and S-SEBI, with different definitions of dry and wet limits.
Surface Energy Balance Algorithm for Land (SEBAL)
The SEBAL algorithm utilizes satellite imagery and meteorological measurements for estimating ET within a single-source surface energy budget [36]. This algorithm addresses the problems of early one-source SEB models concerning the local application and the failure in spatiotemporal scalability. In addition, SEBAL is capable of handling TIR images with different resolutions and provides a strong tool for estimating spatiotemporal changes in ET with minimal data requirements.
The complexity of the SEBAL model is a result of the computation process of sensible heat flux and aerodynamic resistance due to changes in the weather and surface characteristics. As a result, SEBAL considers an internal calibration using two reference pixels chosen by the user’s judgment for estimating the sensible heat flux and stability-corrected aerodynamic resistance during an iterative procedure [36]. It should be noted that the iterative calibration process can reduce the effect of errors associated with LST observations on the algorithm [164].
The reference points include cold pixels, where ET = PET (or reference ET) and H = 0, and hot pixels, where ET = 0 and H = Rn − G. The successful application of SEBAL depends on the accurate selection of these pixels in the scene, which is implemented through a subjective procedure. According to [36], the cold and hot pixels are defined as a water body and a bare agricultural field, respectively. However, the aforementioned definitions can be unrealistic in some situations such as areas with extreme weather conditions, semi-arid areas without any water bodies, mountainous basins, and generally, heterogeneous landscapes due to the invalid linear relationship between LST and dT (which is the near-surface temperature difference) [165,166]. In addition, SEBAL is restricted by different atmospheric stability regimes and bumpy land surfaces, and errors in the TIR land surface temperature or dT can affect the H estimates. Because of these drawbacks in SEBAL, the physically based model of SEBS was developed, which is described as follows.
Surface Energy Balance System (SEBS)
The Surface Energy Balance System (SEBS) is a well-known model first proposed by Su [5]. The SEBS, a derivative of SEBI, is implemented using spectral radiance and reflectance observations in combination with ground-based data for estimating the surface energy balance and the atmospheric turbulent fluxes. In this model, a new SEB-based formulation is presented for calculating the evaporative fraction at limiting cases (i.e., completely dry and wet areas) using a non-subjective procedure. At the dry boundary, ET is assumed to be zero, and the sensible heat flux reaches its maximum value (i.e., H d r y = R n     G ). In contrast, ET occurs at a potential rate limited by energy availability at the wet boundary, and sensible heat flux takes its minimum value (i.e., H w e t = R n     G L E w e t ). The evaporative fraction (EF) is finally expressed as:
E F = L E R n G = ( 1 H H w e t H d r y H w e t ) L E w e t R n G
where H is the sensible heat flux (W/m2), H d r y and H w e t are the sensible heat fluxes for the dry and wet boundaries (W/m2), respectively, and L E w e t is the latent heat flux for the wet condition (W/m2). This method is capable of estimating ET in various spatial scales from local to regional under all atmospheric stability conditions using the Bulk Atmospheric Similarity theory [167] for atmospheric boundary layer scaling [5] and the Monin–Obukhov Atmospheric Surface Layer theory [106] for surface layer scaling. In addition, the roughness length for heat transfer is computed using an extended model as well as surface parameters such as albedo, emissivity, canopy cover, etc. It is worth noting that the SEBS algorithm does not require any prior knowledge of turbulent heat fluxes, and the consideration of SEB at the calibration limits can mitigate the uncertainties originating from LST and atmospheric variables. However, enormous amounts of required data and the relatively complex structure of the model can cause inconveniences in data-sparse regions. In addition, SEBS is the most sensitive single-source model to the aerodynamic parameters and surface–air temperature gradient [168,169].
Simplified Surface Energy Balance Index (S-SEBI)
Similar to the SEBI model, the Simplified Surface Energy Balance Index (S-SEBI) calculates the latent heat flux by an evaporative fraction, which is dependent on wet and dry conditions [37]:
L E = T H L S T T H T L E ( R n G )
where T H   represents the temperature for dry conditions controlled by radiation (K), and T L E shows the temperature for wet conditions controlled by evaporation (K). T H corresponds to the minimum LE (i.e., LEdry = 0) and maximum H (i.e., Hdry = Rn − G), and T L E corresponds to the maximum LE (i.e., LEwet = Rn − G) and minimum H (i.e., Hwet = 0). In the S-SEBI model, T H and T L E are determined by the image itself, which has constant atmospheric conditions with dry and wet pixels, while SEBI estimates extreme temperatures through an external data source [170].
In addition, T H and T L E can be regressed to the surface albedo, and LE is calculated as shown below:
L E = ( a m a x + b m a x α ) L S T ( a m a x + b m a x α ) ( a m i n + b m i n α ) ( R n G )
where α is the surface albedo (dimensionless), and a m a x , b m a x , a m i n , and b m i n are regression coefficients.
Since the main concept of the model is derived from the contrast between the minimum and maximum land surface temperature for dry and wet limits, this method can only be applied to homogenous areas with both dry and wet parts and constant atmospheric conditions across the image [37,171]. S-SEBI does not calculate sensible heat flux, which makes the method different from other SEB models. Furthermore, there is no need for additional data except the surface temperature and surface albedo derived from remotely sensed imagery to estimate ET. Unlike other models, such as SEBAL, which consider a fixed temperature for extreme conditions, the S-SEBI model regulates the variations of LST in proportion to albedo values over dry and wet conditions. It should be noted that the determination of LST for extreme areas is location-specific under unstable atmospheric conditions over large areas. According to Rocha et al. [172], dependency on LST for the S-SEBI model is less than other models of ET estimation.
Mapping Evapotranspiration with high Resolution and Internalized Calibration (METRIC)
To reduce computational errors of remote sensing-based SEB models in mapping ET over heterogeneous surfaces, METRIC, developed by Allen et al. [29], focuses on an automatic internalized calibration of the SEB equation. The calibration process primarily depends on the air surface temperature difference retrieved radiometrically to simplify the calculation of H. In this way, two anchor pixels, similar to the SEBAL algorithm, are defined to describe the evaporative boundary conditions of the SEB. Cold (or wet) and hot (or dry) reference pixels are chosen in a well-irrigated field with full crop cover and a bare, dry field, respectively. Unlike the SEBAL algorithm, which has an assumption of zero evapotranspiration at the hot pixel, METRIC uses a daily soil water balance model to verify whether ET of the hot pixel is equal to zero. This modified assumption of non-zero evapotranspiration makes the estimates of H more reliable [173], and the selection process of candidate pixels is employed more accurately [174]. At the cold pixel, ET is equated to 1.05 times the reference evapotranspiration, which is not applicable during non-growing seasons as well as early growing seasons [175]. To reduce the uncertainties of selecting the anchor pixels, some studies have employed different search algorithms to automate the selection process (e.g., [176,177]).
METRIC uses the reference ET to scale instantaneous evapotranspiration instead of the evaporative fraction employed by SEBAL, which can better take into account the effects of wind speed and humidity variations. Both METRIC and SEBAL use the near-surface temperature gradient instead of the TIR land surface temperature, which eliminates the systemic bias of LST [178]. The SEBAL and METRIC models underestimate ET in non-agricultural and arid areas with extreme weather conditions [4]. In such situations, SEBS can be a suitable alternative because of its non-subjective approach in selecting the limiting cases [174].
Simplified Surface Energy Balance (SSEB)
To rapidly estimate ET over large areas, a simplified and cost-effective method called Simplified Surface Energy Balance (SSEB) was developed by Senay et al. [38]. A joint use of LST and reference ET is the basis of the SSEB model so that reference ET is adjusted by ET fraction calculated from the LST data. Finally, actual ET is acquired by the following Equations (35) and (36):
E T a = E F × β × E T r
E F = T h L S T T h T c
where E T a (m) is the actual ET, E F (dimensionless) is the evaporative fraction, β (dimensionless) is a coefficient for considering the specifications of the reference crop (such as corn, alfalfa, and wheat), E T r (m) is the reference ET calculated by the standardized Penman–Monteith equation [81],   T h (K) is the average LST for a set of three hot pixels, T c (K) is the average LST for a set of three cold pixels, and LST represents the value of the land surface temperature for any given pixel in the scene.
Similar to the SEBAL and METRIC models, SSEB scales ET between the minimum and maximum ET (corresponding to hot and cold pixels, respectively) in proportion to changes in LST. However, SSEB employs the LST scalar to directly derive ET without the estimation of sensible heat flux. On the other hand, SSEB ignores the soil heat flux and albedo, thus underestimating ET for surfaces with low albedo and overestimating ET for surfaces with high ground heat flux and high albedo. Therefore, some corrections such as the Normalized Difference Vegetation Index (NDVI) and Digital Elevation Model (DEM) corrections can be applied to improve SSEB for heterogeneous landscapes [179].
Operational Simplified Surface Energy Balance (SSEBop)
Another remote sensing one-source model is the Operational Simplified Surface Energy Balance (SSEBop) [39], which was established based on SSEB’s theory [38]. The SSEBop model, unlike SEBAL, METRIC, and SSEB, utilizes a non-subjective procedure in order to select the hot and cold reference pixels. Since the required data for this model are LST, E T r , and air temperature, SSEBop benefits from a much easier concept than other one-source models. In addition, SSEBop is more applicable in comparison with METRIC and SEBAL over large regions because SSEBop simulates the SEB model using fewer parameters. The actual ET is calculated using a procedure that is similar to that of SSEB, as shown below:
E T a = E T f × β × E T r
The land surface temperature for the cold pixel ( T c ) is calculated using the corrected air temperature with the assumption of clear-sky conditions. By adding the dT to T c , the land surface temperature for the hot pixel ( T h ) is computed. It is worth noting that dT is obtained by employing the partial solution of the SEB for the cold pixel (where ET = 0 and H is maximal) and the hot pixel (where ET is maximal and H is minimal). SSEBop may misestimate the ET values on surfaces with low vegetation cover and high albedo due to the underestimation of TIR LST. In this regard, a mask or correction factor is applied to these areas in order to increase LST.

Two-Source (or Two-Layer) Models

Two-source models employ the surface temperature of the soil and canopy to estimate the evaporation and transpiration. Given that the LST derived by TIR sensors is a single temperature, various approaches are adopted to decompose the surface temperature into the soil and vegetation temperatures, as explained below.
Two Source Energy Balance (TSEB)
The Two-Source Energy Balance (TSEB) model estimates evaporation from soil and transpiration from the canopy separately [40]. Accordingly, the SEB equation is solved for the soil and vegetation components by partitioning the energy fluxes based on the radiometric surface temperature (Trad) derived from remote sensing platforms. To calculate the ensemble radiometric surface temperature, the brightness temperatures at multiple angles are employed as follows:
T b r ( θ ) = [ ε ( θ ) ( T r a d ( θ ) ) N + ( 1 ε ( θ ) ) T s k y N ] 1 / N
where T b r ( θ ) represents the brightness temperature at a viewing angle θ (K), ε ( θ ) is the thermal emissivity at θ (dimensionless), and Tsky is the hemispherical brightness temperature of the sky. Trad is divided into the thermodynamic temperatures of vegetation and soil by a radiometer view fraction ( f ( θ ) ) via Equation (39):
T r a d ( θ ) = [ f ( θ ) T c N + ( 1 f ( θ ) ) T s N ] 1 / N
where T c and T s express the surface temperatures for the canopy and soil, respectively. T c and T s are used to compute the sensible heat flux, and LE for the canopy and soil are calculated using the residual approach of the energy budget, expressed as:
L E s = R n e t , s H s G
L E c = R n e t , c H c
where L E s and L E c are the latent heat fluxes for the soil and canopy (W/m2), R n e t , s and R n e t , c show the net radiation fluxes for the soil and canopy surfaces (W/m2), which are mainly estimated by Beer’s Law, G is the soil heat flux (W/m2), and H s and H c are the sensible heat fluxes for the soil and canopy components, calculated by decomposed Trad:
H s = ρ a c p ( T s T a r s + r a )
H c = ρ a c p ( T c T a r a )
where ρ a is the air density at constant pressure (Kg·m−3), C p is the specific heat capacity of the air at constant pressure (J·kg−1·K−1), T a is the air temperature (K), r s is the resistance of the boundary layer to heat flux just above the soil surface (s/m), and r a is the aerodynamic resistance (s/m).
If the brightness temperature is observed at a single view zenith angle, the relationships of Priestly–Taylor and Penman–Monteith are used to estimate T c and plant transpiration, called TSEB-PT [40] and TSEB-PM [180], respectively, and T s is then computed using Equation (39). One of the main difficulties in such methods is that the coefficients of Priestly–Taylor and Penman–Monteith vary over the heterogeneous areas [40,181,182].
Two-Source Time-Integrated Model (TSTIM)
The Two-Source Time-Integrated Model (TSTIM) is a modification of the TSEB model, which combines TSEB with a planetary boundary layer integrated by time for the measurement of turbulent fluxes at the regional to continental scales [183]. To estimate the sensible heat flux, TSTIM utilizes two instantaneous satellite-based surface temperatures during the morning with a 4-h interval. Accordingly, an instantaneous estimate for H is related to a radiometric surface temperature using the surface layer component of the model, and thus, a time-integrated value for sensible heat flux is obtained from these instantaneous estimates [184]. Another flux is estimated by the planetary boundary layer component of the model, and during an iterative procedure, TSTIM converges on estimates for H by means of a comparison between time-integrated fluxes. Due to the combination of the sensible heat flux with the temporal variations of the surface temperature and temperature of the boundary layer, the requirement for measuring air temperature is eliminated. Furthermore, the temporal variations of radiometric temperature can reduce the effects of biases of Trad on the computed fluxes. Therefore, the TSTIM model is less sensitive to systematic biases of remotely sensed surface temperature in comparison with TSEB. It is worth noting that the calculations required to estimate LE for this model are similar to the TSEB model when two surface temperatures are observed (see the previous section), except Equation (44) that is simplified by linearizing, as in:
T r a d ( θ ) f ( θ ) T c + ( 1 f ( θ ) ) T s
Atmosphere–Land Exchange Inverse Model (ALEXI)
The TSEB scheme can be employed to generate maps of daily energy fluxes at a continental scale using brightness temperature variations derived by the Meteosat Second Generation (MSG) and Geostationary Operational Environmental Satellites (GOES), the land cover map derived by AVHRR, and weather data within a time-integrated framework called the Atmosphere–Land Exchange Inverse Model (ALEXI) [41]. The real-time daily energy fluxes of LE, H, G, Rn, and soil moisture are obtained from this model with a spatial resolution of 5 to 10 km [174]. In order to improve the spatial resolution of the resulting fluxes, DisALEXI [185], which follows a spatially disaggregated technique, was developed using LST retrieved from polar orbiting satellites.
Dual Temperature Difference (DTD)
Another two-source time-integrated model is Dual Temperature Difference (DTD), which utilizes two LST observations similar to the TSTIM and ALEXI models [42]. However, DTD benefits from a simpler structure and fewer input data, thereby making it more applicable than the two previous models. As with TSEB, the partitioning of LST for the soil and canopy components in the DTD model is based on a fraction of the radiometer view. The flux H can be computed by Equation (45):
H = ρ c p ( T r a d T a   )   f ( θ ) × H c × r a ( 1 f ( θ ) ) ( r s + r a )
By applying Equation (45) two times and subtracting the flux H at t1 from that at t2, the main equation of the DTD model is obtained [42].
Finally, the total latent heat flux at time t2 (LEt2 in W/m2) is estimated as a remainder term of the SEB equation:
L E t 2 = R n , t 2 H t 2 G t 2
and the latent heat flux for soil (LEs in W/m2) is also calculated as a residual term:
L E s = L E L E C
where L E C is the latent heat flux for the canopy (W/m2), which is obtained from the relationship of the Priestly–Taylor coefficient.
Two-source Trapezoid Model for Evapotranspiration (TTME)
Two-source models can estimate the LST for soil and canopy components using either surface temperature observations at a single viewing angle or radiometric temperature observations at two-view zenith angles. To retrieve the LST data from a single observation, vegetation index (VI)/LST space-based models were developed using a VI/LST space, which is defined as a triangular or trapezoidal space formed by the scatter plot of VI versus LST [186]. In addition, VI/LST spaces include isolines of soil moisture, which are used to represent the soil surface temperature [186,187,188]. Trapezoidal spaces are more adequate in determining water scarcity and moisture stress, plant transpiration, and aerodynamic behavior of the surface in comparison with triangular spaces. Space-based models, unlike non-space-based models, do not require the parametrization of resistance network, and therefore, the ET modeling is implemented without the need for excessive data such as plant physiology and micrometeorology information [189,190,191,192,193,194]. The two-source Trapezoid Model for Evapotranspiration (TTME) is one of the space-based models that utilizes a trapezoidal fc-Trad framework and soil wetness isopleths for estimating SEB components [188]. Flux LE is estimated using a patch approach by which the latent heat flux for vegetation and soil is weighted through a canopy fraction ( f c ):
L E = f c L E c + ( 1 f c ) L E s
where L E c denotes the latent heat flux for the canopy ( L E c = R n e t , c × E F c ) and L E s denotes the latent heat flux for the soil ( L E s = ( R n e t , s G ) × E F s ) . To estimate evaporative fractions, the radiometric temperature is divided into surface temperatures for the canopy and soil by the upper and lower boundary conditions, i.e., (Trad = Ts,max, fc = 0) for a bare surface with the highest humidity, (Trad = Ts,min, fc = 0) for a bare soil with the lowest humidity, (Trad = Tc,min, fc = 1) for a fully vegetated area with the lowest temperature, and (Trad = Tc,max, fc = 1) for a fully vegetated area with the highest temperature.
Hybrid dual-source scheme and Trapezoid framework-based Evapotranspiration Model (HTEM)
HTEM is a modified version of the TTME model in which a hybrid dual-source framework of the patch and layer methods is employed for estimating surface energy fluxes [195]. The layer approach is used to assign Rn values for the canopy and soil components established based on Beer’s law, and the patch approach is used to decompose the net radiation flux into the sensible heat, ground heat, and latent heat fluxes weighted by a canopy fraction. In addition, a sub-model is defined to partition the bulk LST into soil and vegetation surface temperatures, which is the same as that of TTME. As with TSEB, HTEM determines the sensible heat fluxes of the soil and canopy based on Ohm’s law, and the latent heat fluxes for the soil and canopy are computed as residual terms of the SEB equation:
L E s = ( R n e t , s G 1 f c ) H s
L E c = ( R n e t , c f c ) H c
Enhanced Two-Source Evapotranspiration Model for Land (ETEML)
A key point of VI/LST space-based models is their relatively independent structure in terms of site-specific calibration of parameter sets [186,196]. However, the application of these models is restricted by several issues, including (1) the requirement for a large space to contain a wide range of canopy fractions and soil moisture levels, limiting the application of models over heterogeneous regions [80], (2) the correct selection of wet and dry edges [197,198], (3) satisfying the assumption of the isoline control by soil moisture availability [197], and (4) the difficulty of coupling the VI/LST space-based models with other models due to different selections of wet/dry edges [199].
To overcome some of the issues, the Enhanced Two-Source Evapotranspiration Model for Land (ETEML) was developed by describing a theoretical VI/LST space [43]. Similar to other models, the flux decomposition into soil and canopy is employed using a patch approach, and the component latent heat fluxes can be estimated as follows:
L E s = ( 1 S W D I ) × P E T s
L E c = ( 1 C W S I ) × P E T c
where S W D I denotes the soil water deficit index (dimensionless), C W S I is the crop water stress index (dimensionless), P E T s is the potential ET for the soil (W/m2), and P E T c is the potential ET for the canopy (W/m2).
To derive P E T s and P E T c , a theoretical VI/LST space is determined for each pixel in which the surface–air temperature difference is used instead of the absolute LST. Accordingly, four boundary points, i.e., dry bare soil, saturated bare soil, fully vegetated and well-watered surface, and fully vegetated surface with well-watered conditions, are defined for each pixel based on the Rn, Ta, G, and vapor pressure deficit quantities. In addition, the radiometric surface temperature for each pixel is divided into soil and canopy temperatures based on the described theoretical VI/LST trapezoid space, as follows:
T c T a = K s × ( 1 f c ) + ( L S T T a )
T s T a = ( L S T T a ) K s × f c
where K s is the isoline slope. Table 2 shows a summary of SEB algorithms, one-source and two-source models, and LST configurations.
It is worth noting that a wide range of remotely sensed ET products, such as MODIS (MOD16), AVHRR, GLEAM, and SSEBop, were developed based on the SEB algorithms to provide spatiotemporal changes in evapotranspiration at different scales. In addition to internal uncertainties caused by the estimation method (stated before in detail in Section 2), ET estimates include some external inaccuracies related to the input data. Satellite data with high spatial resolution (such as ASTER and Landsat) suffer from low return rates, and the platforms with coarse spatial resolution (such as SMAP and MODIS) have low temporal resolution. In addition, delays in data collection by sensors, which vary from an hour to several days, make practical applications challenging. Another major uncertainty is the spatial scale mismatch that occurs between coarse meteorological data and fine vegetation information [195]. In addition, biome-specific physiological parameters and the leaf area index, which are assumed to be constant over space and time, can lead to the errors in ET retrieval, as is the case for the MOD16 product.

3. Limitations and Challenges

3.1. Land Surface Temperature

As stated above, PM-based models use remotely sensed vegetation indices and meteorological data instead of LST information, and practically, the resistance terms ra and rs are determinative for implementing these models. This assumption, i.e., the big leaf basis and elimination of LST information, leads to a decrease in the accuracy of ET results [83]. On the other hand, the models following approach 3 require cloud-free LST images and atmospheric corrections for mapping ET values, although it is not straightforward to retrieve cloud-free LST from TIR data in large scales with monsoon climates due to large spatio-temporal gaps [97]. In other words, TIR radiation is sensitive to atmospheric effects such as water vapor, surface emissivity, and cloud cover, and this leads to discontinuous information for ET estimation [97,183,200]. In this regard, LST observations with high temporal and spatial resolutions are useful to provide cloud-free surface temperature information [174]. In addition, microwave remote sensing can be a suitable surrogate for providing LST in all-sky conditions because it is not limited by cloud cover; however, its coarse spatial resolution causes ET estimation to only be possible for large scales [201].
Another important challenge of TIR remote sensing in estimating energy fluxes is that the radiometric surface temperature (Trad) derived by satellite is assumed to be equivalent to the aerodynamic temperature of the surface (Taero), which is defined as a bulk effective, unmeasurable temperature for estimating heat fluxes in the atmospheric surface layer [202,203,204,205]. While the relationship between Trad and Taero is not clear, they cannot be considered equal because of the difference between the thermodynamic temperatures of vegetation and soil components. Therefore, one of the error factors in estimating ET, unlike the atmospheric correction of Trad, is related to the assumption that the difference between these two temperatures can be ignored [79].
Component temperatures contribute to the aerodynamic and brightness temperatures in proportion to the roughness length and view geometry of the sensor [40,203,204,205,206,207]. This issue is intensified for one-source models employing a single LST observation for estimating the sensible heat flux so that depending on the viewing angle of the sensor, differences between Taero and Trad for the canopy and soil can be more than 10 degrees [203,204,205,208]. A common solution for reducing the errors due to this assumption is that an excessive resistance, associated with heat transfer, is introduced to reflect the difference between the heat and momentum transfer [209,210]. However, this method is limited due to the need for calibrating the parameters concerning resistance changes in space and time [210,211]; that is why many models were developed without using the excessive resistance term [212,213,214].
Absolute measurements of LST lead to considerable errors when applied to partially vegetated surfaces [215], and the effects of the canopy architecture and viewing geometry on the LST retrievals are troublesome in semi-arid areas with partial canopy cover [40]. To overcome the issues of sensitivity of SEB algorithms to errors of absolute LST, models that employ a temperature gradient instead of an absolute surface temperature have been developed to reduce the errors associated with the air temperature [29,36]; however, this issue still causes uncertainties over heterogeneous areas. To address this challenge, two-source models, which decompose LST and energy balance regimes into soil and vegetation components, and include the resistances in boundary layers of components by accounting for the difference between Taero and Trad, have been proposed to estimate ET over complex surfaces [40,216].
In addition to TIR LST data, land surface temperature values can be obtained from the energy closing at the land surface (i.e., approach 2, which uses the MT equation). Unlike TIR-based LST, which is restricted under overcast weather conditions, EEBT derived by approach 2 is usable in all-sky condition; however, in such situations, the relationships used to calculate energy fluxes and as well as the proper energy closure have a significant effect on the quality of the achieved LST responses. If the energy balance equation is dictated to close, the quantities of LST can be less or more than its real values.

3.2. Energy Balance Closure

In addition to fluxes H, LE, G, and Rn, the SEB equation includes some neglected effects that make it unclosed, i.e., the Imb term:
R n = λ E T + G + H + I m b
Many studies have reported an underestimation of turbulent fluxes compared to available energy, showing a non-closed energy balance (e.g., [217,218,219,220,221,222]. A large number of underlying reasons can cause the non-closure of SEB, which can be classified into four categories: instrumental and measurement/computation errors, errors related to data processing and analysis, sub-mesoscale transport processes, and neglected sources of energy [54].
Sensors are responsible for instrumental errors when measuring wind speed [223,224], soil humidity [54], net radiation [225], and ground heat flux [226,227]. Several methods were proposed to satisfy the assumption of zero vertical wind over the averaging time; however, some of them failed to achieve this condition and led to errors in the processing chain [228,229,230]. In addition, some corrections need to be applied to the measured flux of the EC system to avoid the lack of SEB closure [231]. Over heterogeneous surfaces, EC systems are not capable of measuring the non-turbulently transported energy, and as a result, energy fluxes are underestimated due to sub- mesoscale transport processes [232,233]. On the other hand, ignoring the additional thermal terms, i.e., the soil heat storage, potential energy of air and water, and biomass and photosynthesis/respiration heat storage, can cause considerable errors, especially in areas with tall vegetation [234,235,236].
With these considerations in mind, SEB closure seems to occur for a horizontally uniform, non-vegetated surface with vertical fluxes [54]. However, such an assumption cannot be realistic in practical applications, and effort is required to achieve an acceptable SEB closure. If the SEB equation is not appropriately closed, ET estimation models that follow approach 1 overestimate the evapotranspiration quantities due to adding the Imb term to the ET [200]. Similarly, approach 2 misestimates ET values because LSTs are obtained from an energy balance dictated to be closed. However, utilizing the land surface temperature for calculating all the energy terms leads to an increase in the accuracy of estimated fluxes [237]. Accordingly, ET estimation models employing the MT equation and obtained LST from the SEB closing (i.e., approach 2) are more accurate than the other models. In contrast, PM-based models are less accurate in estimating energy fluxes due to the elimination of land surface temperature.

3.3. Resistance Network

Depending on the parameterization of the resistance network, resistance-based ET models, including PM-based models (approach 1), MT-based LSMs (approach 2), and residual- and EF-based models (approach 3), are grouped into one, two, and multi-layer structures; therefore, accurate quantification of the resistance network is important to better estimate turbulent fluxes [238,239,240,241]. The resistance network can include aerodynamic and surface resistances in the on-layer structure as well as boundary layer and canopy bulk stomatal resistances in the multi-layer scheme [40,92,241,242,243].
Measuring the aerodynamic resistance is difficult because of its unmeasurable parameters such as the roughness length and atmospheric stability coefficient [244]. Various methods have been proposed to estimate aerodynamic resistance, which can be classified into three groups: empirical, semi-empirical, and methods following MOST [65]. The degree of validity and simplification in solution are the most important characteristics of the aerodynamic resistance estimation models. Unlike MOST-based methods that differentiate eddy diffusivities for heat and momentum transfer (e.g., [71,245]), most empirical approaches assume that the roughness length for momentum and heat transfer are identical (e.g., [120,246,247,248]), which makes it a simple method but invalid for sparsely vegetated surfaces. Similar to MOST-based models, semi-empirical methods also consider a distinction between the eddy diffusivities of heat and the momentum transfer (e.g., [69,249]). In addition to the surface property of roughness length [65], the atmospheric stability condition is considered as a crucial parameter to account for the micrometeorological changes in estimating aerodynamic resistance, which can be assumed to be neutral when a stable wind speed profile and a homogeneous LST are dominant at the near-surface [250]. Under stable conditions, the profiles of air temperature and wind speed follow the linear equations, and there is an exact solution for estimating aerodynamic resistance. While the aforementioned profiles follow non-linear equations under unstable conditions, an iterative process is needed to obtain an exact solution for aerodynamic resistance. In this regard, two indicators, the Richardson number [251] and the Monin–Obukhov stability length [106], have been proposed to characterize non-neutral conditions, which are mainly adopted for non-iterative and iterative procedures, respectively [36,209,252,253,254].
In addition to the aerodynamic resistance, the application of ET estimation models is also limited by the accurate parameterization of canopy resistance [55], which can be considered as the diffusive resistance to water vapor transfer from stomatal cavities of leaves [255]. Estimating the surface resistance is challenging because of the combined physiological and climatological effects included in the concept of resistance [10]. Considering the big leaf approach, Szeicz and Long [256] proposed a widely used method based on the effective LAI for estimating canopy surface resistance. Afterwards, Jarvis [257] presented a novel method based on the summation of resistance values of individual leaves within a canopy. To progress this method toward a more complicated one, Lhomme [258] introduced a multilayer approach, which estimates canopy resistance as a physiological parameter in zero-evaporation conditions. The sap flux method, by considering a moderator between canopy- and leaf-level measurements, provides estimates of the canopy resistance that agree well with the canopy conductance measured using a photometer [259,260]. Even though this method avoids disturbance of the leaf boundary layer, its application is more widespread for woody canopies than agronomic vegetation [255]. Some methods were also proposed to estimate canopy resistance based on a linear [261] and nonlinear [262] relationship between aerodynamic and climatic resistance. The empirical basis of these methods leads to the need for the calibration of parameters [55,57,58,59,60], and in most cases, they are faced with great uncertainties in estimating ET [61]. To eliminate the need for calibration, Todorovic [263] introduced a mechanistic model, in which canopy resistance is estimated based on a relationship between aerodynamic resistance and climatic variables. However, this method failed to estimate the resistance in some study cases [264]. Without including the effect of aerodynamic resistance on the canopy resistance, Jarvis [62] proposed a method in which the canopy resistance depends on environmental factors such as the vapor pressure deficit and solar radiation. This method was completed by considering more environmental factors, and was named the Jarvis–Stewart equation [62,63]. Even though the Jarvis–Stewart equation is a widely used method for estimating canopy resistance, it is, especially in LSMs, mostly used in a specific growth stage of the crop [238].

4. Conclusions and Perspective

LST is the most important factor in determining the biophysical processes in land–atmosphere boundary areas including ET. Due to the absence of a sufficiently dense network of stations measuring LST, the surface temperature values are obtained from indirect methods, which leads to the development of various models of ET estimation. The current review provided a clear classification for ET models based on available retrieval methods of LST information. The first approach includes PM-based models, which follow a psychrometric assumption based on eliminating LST; instead, they use remotely sensed vegetation indices and meteorological data for estimating ET. Despite the wide popularity of this approach, the assumption of ignoring LST leads to decreased accuracy in the resulting ET. With the development of LSMs, which simulate the Soil–Vegetation–Atmosphere system by using sophisticated parameterizations, the land surface temperature data (known as the Equilibrium Energy Balance Temperature, EEBT) result from closing the energy balance equation at the land surface, which forms approach 2. The EEBT values are valid for the all-sky condition; however, a precise formulation for the SEB equation and an acceptable SEB closure are needed to ensure the accuracy of the ET results. By providing spatiotemporal distribution of LST via TIR-based earth observation data, approach 3 was made to map ET over large and heterogeneous regions. Different methods were proposed to derive LST from TIR data, which are grouped into four classes of single-channel methods, multichannel methods, multitemporal methods, and multiangle methods. In addition, a wide range of models, residual- and EF-based, were developed to employ these TIR-derived LSTs to estimate ET through energy conservation principles. Despite the strength of these models in determining energy fluxes, all of the above approaches face a number of limitations in estimating ET. One of the most important challenges is related to the accuracy of LST information. As stated above, approach 1 suffers from the assumption of ignoring LST and approach 2 may estimate unrealistic LST values if the surface energy balance is not properly closed. In addition, the models following approach 3 require cloud-free LST images, but it is not straightforward to retrieve cloud-free LST from TIR data, and it can lead to a decrease in the accuracy of the ET results. It should be noted that models using approach 3 assume that the radiometric surface temperature derived by satellite is equivalent to the aerodynamic temperature of the surface, which can be one source of error. Therefore, models such as SEBAL and METRIC have been developed to reduce the uncertainty due to the aforementioned assumption using the air surface temperature gradient instead of absolute LST. On the other hand, it was proven that the SEB equation cannot be closed completely. Neglected effects, including instrumental and measurement/computation errors, errors related to data processing and analyzing, sub-mesoscale transport processes, and ignored sources of energy, can affect the LST and ET results. As the last important challenge, the complicated parametrization of resistance network can limit the application of ET estimation models. Different configurations of the resistance network (surface and aerodynamic resistances) lead to the development of different resistance-based models, including one-layer, two-layer, and multi-layer models. Given the unmeasurable parameters of the resistance network, empirical, semi-empirical, and MOST-based methods were introduced to estimate the aerodynamic and surface resistances. Empirical and semi-empirical methods mainly require the calibration of parameters while the MOST-based approach can provide a stronger tool in estimating the resistance network.
ET evaluation involves energy and mass conservation laws [265]. However, available methods of ET estimation that belong to either approach 1 or 3 are determined largely by energy availability, without explicit water resource constraints. Similarly, Ets estimated using LSMs (approach 2), despite their use of the energy–water equation system, are mainly limited by energy resources due to the one-way solution of the system. In such situations, ET values are overestimated or underestimated under extremely dry and wet conditions, which lead to larger time periods for drought recovery and larger potential for floods, respectively. Therefore, for the future work, a simultaneous and interactive evaluation of water and surface energy balances is recommended to reciprocally connect the atmospheric demand, available surface radiation, and water availability. Accordingly, the SEB closure is conducted to derive LST values using a forward derivation, and ET can then be estimated based on the computed LST and MT equation. By applying ET to the water balance model, the surface soil moisture is updated and then used to compute the LST and ET values at the next time step. In this regard, a two-way interaction is formed between the surface and hydrological properties, i.e., LST, ET, and soil moisture, which is capable of implementing the accurate and correct conceptual relationships between components.

Author Contributions

Conceptualization, A.M., M.N. and M.T.; Data curation, M.T. and F.G.; Formal analysis, A.M., M.T., F.G. and M.B.; Funding acquisition, A.M.; Investigation, M.T., F.G. and M.B.; Methodology, A.M., M.N. and M.T.; Project administration, A.M.; Resources, M.T.; Software, M.T., F.G. and M.B.; Supervision, A.M. and M.N.; Visualization, M.T., F.G. and M.B.; Writing—original draft, M.T., F.G. and M.B.; Writing—review and editing, A.M. and M.N. All authors have read and agreed to the published version of the manuscript.

Funding

The research of A.M was supported by the Natural Sciences and Engineering Research Council of Canada (NSERC).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in Appendix A.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Acronyms
SEBSurface Energy Balance
LSMLand Surface Model
ETEvapotranspiration
SFSap Flow
ECEddy Covariance
BRBowen Ratio
PMPenman–Monteith
MTMass Transfer
VICVariable Infiltration Capacity
FEST-EWBFlash-Flood Event-Based Spatially Distributed Rainfall-Runoff Transformation Energy Water Balance
TIRThermal Infrared
SEBSSurface Energy Balance System
METRICMapping Evapotranspiration at High Resolution with Internalized Calibration
SEBALSurface Energy Balance Algorithm for Land
SEBISurface Energy Balance Index
S-SEBISimplified Surface Energy Balance Index
SSEBSimplified Surface Energy Balance
SSEBopOperational Simplified Surface Energy Balance
TSEBTwo Source Energy Balance
TTMETwo-source Trapezoid Model for Evapotranspiration
ALEXIAtmosphere–Land Exchange Inverse
DTDDual Temperature Difference
ETEMLEnhanced Two-Source Evapotranspiration Model for Land
TSTIMTwo-Source Time-Integrated Model
HTEMHybrid dual-source scheme and Trapezoid framework–based Evapotranspiration Model
FAOFood and Agricultural Organization
PROMETProcess Oriented Models for Evapotranspiration
MODISModerate Resolution Imaging Spectroradiometer
MSGMeteosat Second Generation
GOESGeostationary Operational Environmental Satellites
DisALEXIDisaggregated ALEXI
VIVegetation Index
PMLPenman–Monteith–Leuning
ISBAInteraction Soil–Biosphere–Atmosphere
SEWABSurface Energy and Water Balance
EASSEcosystem–Atmosphere Simulation Scheme
DLMDynamic Land Model
MOSTMonin–Obukhov Similarity Theory
EEBTEquilibrium Energy Balance Temperature
BATSBiosphere–Atmosphere Transfer Scheme
TOPLATSTOPMODEL-Based Land Surface–Atmosphere Transfer Scheme
SVATSoil–Vegetation–Atmosphere Transfer
SiBSimple Biosphere Model
NCEPNational Centers for Environmental Prediction
AVHRRAdvanced Very High Resolution Radiometer
ASTERAdvanced Spaceborne Thermal Emission and Reflection Radiometer
RTERadiative Transfer Equation
D/NDay/Night
VPDVapor Pressure Deficit
LSTLand Surface Temperature
EFEvaporative Fraction
SWDISoil Water Deficit Index
CWSICrop Water Stress Index
DoYDay of Year
NDVINormalized Difference Vegetation Index
LAILeaf Area Index
LULC Land Use Land Cover
RHRelative Humidity
Symbols
EEvaporation rate
E a Actual evaporation
E p Potential evaporation
L E n Evaporation rate due to net radiation
Δ Slope of the saturation vapor pressure curve at air temperature T
γ Psychrometric constant
L E a Evaporation rate due to mass transfer
L E w e t Latent heat flux for wet condition
β Bowen ratio
HSensible heat flux
H d r y Sensible heat fluxes for dry boundaries
H w e t Sensible heat fluxes for wet boundaries
LELatent heat flux
L E p Potential latent heat flux
R n Net radiation flux
GSoil heat flux
ρ a Air density
C p Specific heat capacity of air at constant pressure
r s Surface resistance
r a Aerodynamic resistance to heat and/or vapor transport
e a Actual vapor pressure of air
e o Presents vapor pressure within the leaf
T a Air temperature
TSurface temperature
e s Saturation vapor pressure at Ta
R n Net radiation flux
R n , c Net radiation for canopy
R n , s Net radiation for soil
g a Aerodynamic conductance
g c Canopy conductance
fFactor related to potential evaporation of soil surface
TrTranspiration
CHeat capacity
r i Canopy internal resistance
r e Canopy external resistance
E c Evaporation from intercepted water
E s Evaporation and sublimation from the soil
T r c Canopy transpiration
C s Surface heat capacity
δ T c Change in land surface temperature of the system over time Δ t
E 0 Diffusion-limited maximum evaporation
q s Saturation specific humidity
T c Canopy temperature
f c w Fractional area of canopy occupied by water
E t r , d e m Transpiration by atmospheric demand
E t r , s u p Water supply by roots
E w c Evaporation from the wetted fraction of the canopy
E w g Evaporation from the wetted fraction of the ground cover
E g s Evapotranspiration from the ground cover and soil surface
E s ,Direct evaporation from the soil surface
λLatent heat of vaporization
r ¯ b Bulk boundary layer resistance
r d Aerodynamic resistance between ground and canopy airflow
r ¯ c Bulk stomatal resistance of upper story vegetation
r g Bulk stomatal resistance of ground vegetation
r s u r f Bare soil surface resistance
f h Relative humidity within pore space of surface soil layer
V g Fractional cover of the ground cover
W c Wetness fraction of canopy
W g Wetness fraction of ground cover
e s ( T s ) Saturation vapor pressure at surface temperature
ϵRatio of the molecular weight of water vapor to that of dry air
p s Surface pressure
r e f f Single effective resistance
r a Aerodynamic resistance
nSurface cover class index
E c [ n ] Maximum canopy evaporation
r 0 [ n ] Architectural resistance
W i [ n ] Intercepted water amount
W i m [ n ] Maximum amount of intercepted water
zDepth of soil layer
ΘVolumetric soil water content
DSoil water diffusivity
σ f Fraction of green vegetation
E d i t Evaporation from the top shallow soil layer
B c Embodies canopy resistance
T p b l Average temperature of planetary boundary layer
T H Temperature for dry conditions controlled by radiation
T L E Temperature for wet conditions controlled by evaporation
α Surface albedo
E T r Reference ET
T b r ( θ ) Bleftness temperature at a viewing angle θ
ε ( θ ) Thermal emissivity at θ
TskyHemispherical bleftness temperature of sky
TradRadiometric surface temperature
R s , i n Incoming shortwave radiation
R s , o u t outgoing shortwave radiation
R l , i n Incoming longwave radiation
R l , o u t outgoing longwave radiation
P E T s Potential ET for soil
P E T c Potential ET for canopy
K s Isoline slope
PSurface atmospheric pressure
P 0 Standard atmospheric pressure at sea level
wAtmospheric precipitable water
w e A variable for calculating w
S 0 Solar constant
θ Solar zenith angle
dSun–earth distance
τ Atmospheric transmissivity
zElevation above mean sea level
R a Extraterrestrial radiation
K t Empirical turbidity coefficient
β 24 Daily average of sun angle above the horizon
KBoTransparency index for direct solar radiation
KDoTransparency index for diffuse sky radiation
KtEmpirical turbidity coefficient
φ Geographical latitude
η Turbidity coefficient
MOptical length of atmosphere
JParameter for considering the surface albedo effect
IVapor optical parameter
F w A variable for calculating I
T d Dew point temperature
bParameter for considering the air pressure effect on w
cCloudiness percentage
R s , i n , c l r   Incoming shortwave radiation for clear-sky conditions
c l o w Amount of low clouds
cothAmount of other clouds
Δ T Difference of minimum and maximum air temperatures
QFEAverage air pressure at desired location
QFFAverage air pressure at sea level
nActual sunshine hours
NMaximum possible sunshine hours
P r Precipitation
eActual vapor pressure
σ Stefan–Boltzmann constant
R s , i n , c l r   Incoming shortwave radiation for clear-sky conditions.
δ Solar declination
ωs Sunset hour angle
ε a Atmospheric emissivity
Γ C Ratio of G to R n for areas with fully vegetated cover
ΓsRatio of G to R n for bare soil areas
z m Standard heights for measuring the wind speed
z h Standard heights for measuring the humidity
kVon Karman constant
u Wind speed at the standard height
w z Moisture content at depth z
Rs,min,LLimit value of incident radiation
wcrField capacity
wwiltSoil moisture at the wilting point
r 1 Stomatal resistance for a well-conditioned leaf
L A I a c t i v e Fraction of LAI that contributes to the transfer of vapor and heat

Appendix B

1.
Net radiation flux ( R n )
R n is a key factor in determining the energy balance at the land surface and is computed by budgeting shortwave and longwave radiations as Equation (A1):
R n = R s , i n R s , o u t + R l , i n R l , o u t = ( 1 a l b ) R s , i n + R l , i n R l , o u t
where R n is the net radiation flux (W/m2), R s , i n and R s , o u t are the incoming and outgoing shortwave radiations (W/m2), R l , i n and R l , o u t are the incoming and outgoing longwave radiations (W/m2), respectively, and α represents the surface albedo (dimensionless). R s , i n and R l , i n are the most complicated components in determining the radiation budget due to atmospheric transmissivity and atmospheric emissivity, respectively. To estimate R s , i n and R l , i n , vertical profiles of water vapor, the temperature, ozone, and optical characteristics of tracer gases and aerosols are needed as well as the micro- and macro-physical properties of clouds such as height, drop size, and water content. A general classification is provided for calculating R s , i n   , which includes models based on artificial intelligence [266,267,268], remote sensing and satellite imagery [269,270], and empirical relationships derived from measured radiation fluxes [271]. Among the aforementioned models, empirical methods are the most widely used because of low calculation effort and available input data [272]. Given the types of input data, empirical models are also classified into four groups of models based on sunshine hours [273,274], air temperature [275,276,277], cloudiness [272,278,279], and the combined models [280,281]. Table A1 shows some important estimation models of R s , i n under clear- and all-sky conditions.
In addition to R s , i n , R l , i n has a significant effect on providing the energy for natural processes such as photosynthesis/respiration, evapotranspiration, snowmelt, etc. [272,282]. The Stefan–Boltzmann equation is mainly used to calculate R l , i n , which depends on the atmospheric emissivity ( ε a ). The direct measurement of ε a is restricted by the high cost of equipment [283,284,285]; therefore, empirical relationships associated with meteorological observations are mainly used to calculate the radiation component as well as models based on the radiation transfer theory [282,286,287,288]. Table A2 presents some important methods of estimating R l , i n under clear- and all-sky conditions.
Table A1. Estimation models of R s , i n under clear- and all-sky conditions.
Table A1. Estimation models of R s , i n under clear- and all-sky conditions.
RowMathematical StructureSky ConditionExplanationReferences
1 R s , i n   = 0.72 S 0 cos ( θ ) Clear-sky conditionIt is the simplest method, which uses the average atmospheric transmissivity. The model does not consider atmospheric transmissivity decrease in proportion to solar zenith angle increase; therefore, it is not recommended to use for data with an hourly basis.[289]
2 R s , i n   = 10 + 1411 cos ( θ ) 310 cos ( θ ) Clear-sky conditionThis method has been developed using average hourly measurements for flux R s , i n   in Australia.[290]
3 R s , i n   = S 0 cos ( θ ) ( 0.47 + 0.47 cos ( θ ) ) Clear-sky conditionThe method is based on the model provided by Lumb [291], which is usable for hourly, daily, and monthly scales. Since Lumb’s model is sensitive to local conditions, the coefficients of this relationship are obtained from a new fitness by measured fluxes in Canada.[292]
4 R s , i n   = S 0 ( cos ( θ ) ) 2 1.085 cos ( θ ) + e 0 ( 2.7 + cos ( θ ) ) 10 3 + 0.10 Clear-sky conditionThe model has been calibrated by radiation data in Indian Ocean Islands. Saturation vapor pressure is applied to the model as an input, which leads to the improvement of resulting fluxes’ accuracy. However, seasonal changes in the sun–earth distance are not considered in the model.[293]
5 R s , i n   = S 0 ( cos ( θ ) ) 2 1.085 cos ( θ ) + e 0 ( 2.7 + cos ( θ ) ) 10 3 + 0.1 d 2 Clear-sky conditionIt is a modified version of Zillman’s model due to considering seasonal changes in sun–earth distance.[294]
6 R s , i n   = S 0 ( cos ( θ ) ) 2 1.2 cos ( θ ) + e 0 ( 1.0 + cos ( θ ) ) 10 3 + 0.0455 Clear-sky conditionThis model has better performance than previous models because it uses surface albedo and cloud thickness data. Seasonal variability of sun–earth distance is also ignored in the model.[295]
7 R s , i n   = S 0 ( cos ( θ ) ) 2 1.2 cos ( θ ) + e 0 ( 1.0 + cos ( θ ) ) 10 3 + 0.0455 d 2 Clear-sky conditionBy adding seasonal variations of sun–earth distance, this model improved the performance of the previous method.[294]
8 R s , i n   = ( 0.75 + 2 10 5 z ) S 0 d 2 cos ( θ ) Clear-sky conditionThe model assumes that τ linearly increases in proportion to height.[296]
9 R s , i n   = ( 0.75 + 2 10 5 z ) R a                             f o r   l o w   a i r   t u r b i d i t y
R s , i n = R a   exp ( 0.0018 P K t sin ( β 24 ) )                                 f o r   h i g h   a i r   t u r b i d i t y
Clear-sky conditionThe basis of the model is linearizing the law of radiation attenuation of Beer with an assumption of a sun angle above the horizon of 50 degrees, and it is valid for the stations where have a height of fewer than 6000 m with low air turbidity. Therefore, the model should be revised for conditions with high air turbidity.[297]
10 R s , i n   = ( K B o + K D o ) S 0 d 2 cos ( θ )
K B o = 0.98 e 0.00146 P K t sin ( β 24 ) 0.075 ( w sin ( β 24 ) ) 0.4
K D o = { 0.35 0.36 K B o ,                 K B o 0.15 0.18 + 0.82 K B o ,                   0.065 < K B o 0.10 + 2.08 K B o ,                 K B o 0.065 < 0.15
sin ( β 24 ) = sin   [ 0.85 + 0.3 φ s i n ( 2 π 365 D O Y 1.39 ) 0.42 φ 2 ]
w = 0.14 e 0 P + 2.1
Clear-sky conditionDirect and diffuse radiations are differentiated in this model; therefore, detailed information is required to estimate fluxes. In addition, considering the effects of water vapor on shortwave radiation absorption leads to the improvement of R s , i n , c l r   estimates.[298]
11 R s , i n   = ( C + 0.70 e M F l n 10 ) ( 1 I ) ( 1 + J ) S 0 d 2 cos ( θ )
M = ( P P 0 ) / cos ( θ )
C = 0.21 0.2 η
F = 0.056 + 0.16 η 2
J = ( 0.066 + 0.34 η 2 ) ( α 0.15 )
I = 0.014 ( M + 7.0 + 2.0 F w ) F w
F w = 0.4343 ln ( w 10 )
w = 1.234 w e 0.21
w e = { e 0.0622 t d + 1.985 b ,                                                 T d < 5 e 0.0714 t d + 2.003 b ,                 5 < T d < 23 e 0.0345 t d + 2.851 b ,                                                 23 < T d
b = { 0 ,                                                                             99 < P < P 0 2.3026 ( 1 P P 0 ) 2 ,     P > P 0   o r     P < 99  
Clear-sky conditionThis model uses an empirical coefficient for turbidity (equal to 1 in clear-sky conditions), and it is complicated in comparison to other models.[299]
12 R s , i n   = [ ( 1 c ) + t c c ] R s , i n , c l r  
τ = 0.48           ( Bennett ,   1960 )
All-sky conditionIt is the simplest structure for calculating the radiation flux under all-sky conditions, which uses an average atmospheric transmissivity for estimating monthly fluxes. However, it is inappropriate for calculating radiation fluxes with an hourly temporal scale because changes of atmospheric transmissivity in proportion to solar zenith angle are not considered.[300]
13 R s , i n   = ( 1 0.6 c 3 ) R s , i n , c l r   All-sky conditionThe used data are related to oceans with mid-latitude, and the cloudiness coefficient is a cubic function of total cloudiness.[301]
14 R s , i n   = ( 1 c [ 4.7 2.24 c l o w c + 10 3 ] + 0.31 c l o w 2.46 + 0.73 c o t h 4.7 ) R s , i n , c l r   All-sky conditionIn comparison to other models, this model uses more information such as total cloudiness I, low clouds (clow), and other clouds (coth = c-clow).[271]
15 R s , i n   = R a a Δ T 0.5
a = 0.16
All-sky conditionIt is one of the simplest models for estimating R s , i n , a l l   which uses the difference of minimum and maximum air temperatures[302]
16 R s , i n   = R a b ( Q F E / Q F F ) Δ T 0.5
b = 0.17
All-sky conditionThis model is a modified form of the previous model (i.e., Hargreaves and Samani [302]) in which the average air pressure at the desired location (QFE) and the average air pressure at the sea level (QFF) are used.[303]
17 R s , i n   = R a [ c 1 ( Δ T 2.5 ) + c 2 ( Δ T 1.5 ) + c 3 ( Δ T 0.5 ) ]
c 1 = 0.00185 ,   c 2 = 0.0433 ,   c 3 = 0.4023
All-sky conditionSimilar to the model introduced by Hargreaves and Samani [302], this model utilizes the difference between the maximum and minimum air temperatures.[304]
18 R s , i n   = R a f ( 1 + 2.7 × 10 6 z ) Δ T 0.5 All-sky conditionIt is the modified form of Hargreaves and Samani’s model, which depends on z.[305]
19 R s , i n   = R a [ e 1 + e 2 l n ( Δ T ) ]
e 1 = 0.28 ,   e 2 = 0.15
All-sky conditionIt is another air temperature-based model that uses the difference between the maximum and minimum air temperatures.[306]
20 R s , i n   = R a [ f 1 + f 2 ( n / N ) ]
f 1 = 0.25 ,   f 2 = 0.5
All-sky conditionIt is the simplest sunshine hours-based model which utilizes the empirical coefficients f 1 and f 2 .[307]
21 R s , i n   = R a [ g 1 + g 2 ( n / N ) + g 3 ( n / N ) 2 + g 4 ( n / N ) 3 ]
g 1 = 0.16 ,   g 2 = 0.87 ,   g 3 = 0.16 ,   g 4 = 0.34
All-sky conditionThis model is a polynomial form of sunshine hours-based models.[308]
22 R s , i n   = R a [ h 1 + h 2 l n ( Δ T ) + h 3 ( n / N ) h 4 ]
h 1 = 0.11 ,   h 2 = 0.04 ,   h 3 = 0.48 ,   h 4 = 0.83
All-sky conditionIt is a combined model which estimates the radiation flux using air temperature and sunshine hours.[306]
23 R s , i n = R a [ h 1 + h 2 ( n N ) + h 3 ( P ) + h 4 ( R H ) + h 5 ( T a ) ]
h 1 = 0.3811 ,   h 2 = 0.0086 ,   h 3 = 0.0388 ,   h 4 = 0.0075 ,   h 5 = 0.0002
All-sky conditionIt is a combined model in which air temperature, relative humidity, and sunshine hours are used to estimate the radiation flux.[309]
24 R s , i n = R a [ 0.4677 + 0.3279 ( n N ) + 0.0001 ( P r ) 0.0019 ( R H ) 0.0024 ( T a ) ] All-sky conditionIt is a combined model in which air temperature, relative humidity, precipitation, and sunshine hours (actual and maximum) are used to estimate the radiation flux.[310]
P = surface atmospheric pressure; P 0 = standard atmospheric pressure at sea level; w = atmospheric precipitable water; S 0 = solar constant;   θ = solar zenith angle; e 0 = saturation vapor pressure; d = sun–earth distance; τ = atmospheric transmissivity; z = elevation above mean sea level; R a = extraterrestrial radiation; K t = empirical turbidity coefficient; β 24 = daily average of sun angle above the horizon; KBo = transparency index for direct solar radiation; KDo = transparency index for diffuse sky radiation; Kt = empirical turbidity coefficient;   φ   = geographical latitude; DoY = day of year; η = turbidity coefficient; α = surface albedo; C and F = aerosol optical parameters for turbidity coefficient; M = optical length of atmosphere; J = parameter for considering the surface albedo effect; I = vapor optical parameter; F w = a variable for calculating I; w e = a variable for calculating w; T d = dew point temperature; b = parameter for considering the air pressure effect on w; c = cloudiness percentage; R s , in , clr   = incoming shortwave radiation for clear-sky conditions; c low = amount of low clouds; coth = amount of other clouds; Δ T = difference of minimum and maximum air temperatures; QFE = average air pressure at desired location; QFF = average air pressure at sea level; f = model parameter; n = actual sunshine hours; N = maximum possible sunshine hours; RH = relative humidity; T a = air temperature; Pr = precipitation.
Table A2. Estimation models of R l , i n under clear- and all-sky conditions.
Table A2. Estimation models of R l , i n under clear- and all-sky conditions.
RowMathematical StructureSky ConditionExplanationReferences
1 R l , i n   = ( 0.83 0.18 10 0.067 e 0 ) σ T a 4 Clear-sky conditionThe model has an empirical structure and is sensitive to local conditions.[311]
2 R l , i n   = ( 0.52 0.065 e 0 ) σ T a 4 Clear-sky conditionSimilar to the previous method, this empirical model is sensitive to local conditions.[312]
3 R l , i n   = 1.24 ( e 0 T a ) 1 / 7 σ T a 4 Clear-sky conditionThe model is based on radiation transfer theory and is recommended for dry and humid weather conditions.[313]
4 R l , i n   = [ 1 0.261 exp ( 7.77 10 4 ( 273 T a ) 2 ] σ T a 4 Clear-sky conditionIt is an empirical method, which only depends on air temperature.[314]
5 R l , i n = [ 0.7 + 5.95 10 5 e 0 exp ( 1500 T a ) ] σ T a 4 Clear-sky conditionThis model follows an empirical structure, and it leads to overestimating R l , i n   under strictly dry weather conditions.[315]
6 R l , i n   = [ 1 ( 1 + w ) exp ( ( 1.2 + 3.0 w ) 0.5 ) ] σ T a 4
w = 46.5 ( e 0 T a )
Clear-sky conditionIt is a model based on radiation transfer theory, which has been calibrated by ground observations. According to Prata [316], this model has provided the best estimation for R l , i n   compared to other models.[316]
7 R l , i n   = [ 0.59 + 0.038 ln ( w e ) + 0.011 ( ln ( w e ) ) 2 ] σ T a 4 Clear-sky conditionAn empirical model that depends on w e and Ta.[299]
8 R l , i n = 1.08 ( ln ( τ ) ) 0.265 σ T a 4
τ = 0.75 + 2 × 10 5 × z
Clear-sky conditionIt is an empirical method that is used in SEBAL and METRIC models.[317]
9 R l , i n = 9.2 × 10 6 σ T a 6 Clear-sky conditionIt is an empirical mode that only depends on air temperature.[318]
10 R l , i n = 59.38 + 113.7 ( T a 273.16 ) 6 + 96.96 w / 25 Clear-sky conditionThis model underestimates R l , i n under inversion conditions.[319]
11 R l , i n   = 1.08 [ 1 exp { e 0 T a 2016 } ] σ T a 4 Clear-sky conditionAn empirical model that depends on e 0 and Ta.[320]
12 R l , i n = [ 1 ( 1 + ϑ ) exp { ( 1.2 + 3.0 ϑ ) 0.5 } ] σ T a 4
ϑ = 46.5( e 0 / T a )
Clear-sky conditionThe model has shown better results than the models presented in rows 3, 4, 5, 9, and 11.[316]
13 R l , i n = [ 0.83 + 0.21 exp ( 0.0027 e ) ]   σ T a 4 Clear-sky condition [321]
14 R l , i n = ( 1 + 0.26 c ) R l , i n , c l r All-sky conditionThe model has been calibrated by data collected in Canada, and it has shown the best performance in Alaska and the Northern areas.[322]
15 R l , i n = ( 1 + 0.22 c 2.75 )   R l , i n , c l r All-sky conditionSince the data for calibrating the model have been collected from Alaska, this method is appropriate for cold weather conditions.[323]
16 R l , i n = ( 1 + [ R l , o u t   R l , i n , c l r 1 ] 0.87 c 3.49 )   R l , i n , c l r All-sky conditionThe data for calibrating the model have been collected in the summer.[313]
e 0 = saturation vapor pressure; e = actual vapor pressure; z = elevation above mean sea level; R l , out   = outgoing longwave radiation; τ = atmospheric transmissivity; z = elevation above mean sea level; σ = Stefan–Boltzmann constant; T a = air temperature; α = surface albedo; w = atmospheric precipitable water; w e = a variable for calculating w; b = parameter for considering the air pressure effect on w; c = cloudiness percentage; R s , in , clr   = incoming shortwave radiation for clear-sky conditions.
Because of similar climatic conditions, the model provided by Jahani et al. [310] is used to estimate R s , i n   for approaches 2 and 3 (Equation (A2)), and the model proposed by Allen et al. [79] is also employed to calculate the shortwave radiation flux for approach 1 (Equation (A3)):
R s , i n = R a [ 0.4677 + 0.3279 ( n N ) + 0.0001 ( P r ) 0.0019 ( R H ) 0.0024 ( T a ) ]   ×   11.6
where R a is the extraterrestrial radiation (MJ·m−2·day−1), n   is the actual sunshine hours (hour), N   is the maximum possible sunshine hours (hour), P r is the precipitation (mm), R H   is the relative humidity (%), T a is the air temperature ( ) , and 11.6 is used to convert MJ·m−2·day−1 to W·m−2.
R s , i n = ( a s + b s n N ) R a × 11.6
where a s and b s are the regression parameters, which are assumed to be 0.25 and 0.5 when there are no data for the calibration process. To calculate R a , the set of equations represented by Allen et al. [81] is utilized as shown below:
R a = 24 × 60 π × S 0 × d × [ ω s × sin φ × sin δ + cos φ × cos δ × sin ω s ]
d = 1 + 0.033 cos ( 2 π D o Y 365 )
δ = 0.4093 sin ( 2 π D o Y 365 1.39 )
ω s = cos 1 ( tan ( δ ) × tan ( φ ) )
N = 24 π ω s
where S 0 is the solar constant (0.0820 MJ·m−2·day−1), d is the sun–earth distance (rad), D o Y is the day of year, δ is the solar declination (rad), ω s is the sunset hour angle (rad), and φ represents the latitude (rad).
As stated above, R l . i n is calculated by the Stefan–Boltzmann equation in W·m−2:
R l . i n = ε a σ T a 4
where σ is the Stefan–Boltzmann constant (5.67 × 10−8 W·m−2·K−4), Ta is the air temperature (K), and ε a is the atmospheric emissivity (dimensionless), which is estimated based on the equation introduced by Swinbank [318], i.e., ε a = 92 × 10 7   T a 2 . Similar to R l . i n , the outgoing longwave radiation ( R l . o u t ) is obtained from the Stefan–Boltzmann equation:
R l . o u t = ε s σ L S T 4
where LST is the land surface temperature (K), and ε s is the surface emissivity (dimensionless), which is calculated through the model provided by Waters et al. [324]:
ε s = { N D V I > 0 { L A I 3 ε s = 0.98                                           L A I < 3 ε s = 0.95 + 0.01 L A I N D V I < 0 ε s = 0.985                                                                                    
where NDVI is the Normalized Difference Vegetation Index, and LAI is the Leaf Area Index. For approach 1, the net flux of R l . i n and R l . o u t (i.e., R n l ) is calculated by Equation (A12) [79]:
R n l = σ [ T a , m a x 4 + T a , m i n 4 2 ] ( 0.34 0.14 e a ) ( 1.35 R s , i n R s o 0.35 )
where Ta,max and Ta,min are the maximum and minimum air temperatures (K), respectively, R s o is the solar radiation for the clear-sky condition in which n = N and it is computed by R s o = ( a s + b s ) R a , and e a is the actual vapor pressure (kPa) calculated by Equation (A13):
e a = R H 100 [ e 0 ( T a , m a x ) + e 0 ( T a , m i n ) 2 ]
where RH is the relative humidity (%), and e 0 ( T a , m a x ) and e 0 ( T a , m i n ) are the saturation vapor pressures at the maximum and minimum air temperature (kPa), which are calculated by e 0 ( T a , m a x ) = 0.611 e x p [ 17.27 T a , m a x T a , m a x + 237.3 ] and e 0 ( T a , m i n ) = 0.611 e x p [ 17.27 T a , m i n T a , m i n + 237.3 ] , respectively.
2.
Soil heat flux ( G )
Soil heat flux (G) is defined as the conducted heat between the land surface and the underlying soil layer, and it depends on thermal conductivity and the vertical gradient of the temperature. Flux G varies between 2% and 20% of Rn; however, it is mainly ignored on a daily scale [72]. At larger temporal scales, G is considered as a percentage of Rn depending on solar zenith angle, vegetation cover, and soil moisture status. For example, G can be calculated using the relationships provided by [5,269,325], i.e., Equations (A14), (A15) and (A16), respectively:
G R n = L S T a l b   ( 0.0038 a l b + 0.0074 a l b   2 ) ( 1 0.98 N D V I 4 )
G = R n   [ Γ C + ( 1 f c ) ( Γ s Γ C ) ]
{ G R n = 0.05 + 0.18 e 0.521 L A I                               i f             L A I 0.5           G R n = 1.80 ( L S T ) R n + 0.084                                               i f                     L A I < 0.5  
where G is the soil heat flux (W·m−2), R n is the net radiation flux (W·m−2), L S T is the land surface temperature (°C),   α is the surface albedo (dimensionless), Γ C is the ratio of G to R n for areas with fully vegetated cover (equal to 0.05), Γ s is the ratio of G to R n for bare soil areas (equal to 0.315), and f c is the canopy fraction (dimensionless). In the current study, Equation (A15) is used for estimating G.
3.
Sensible heat flux ( H )
Sensible heat flux is the amount of heat that is exchanged between the land surface and the air because of the temperature gradient. Flux H is conventionally estimated using a flux-gradient expression based on Ohm’s law [326]:
H = H c + H s = ( 1 f c ) ( ρ a C p r a , s ) ( L S T T a ) + f c ( ρ a C p r a , c ) ( L S T T a )
where H c and H s are sensible heat fluxes for the canopy and soil (W·m−2), respectively, ρ a is the air density at constant pressure (Kg.m−3), C p is the specific heat of the air at constant pressure (1004 J·kg−1·K−1), L S T is the land surface temperature (K), T a is the air temperature (K), r a , s is the aerodynamic resistance for the soil (s·m−1), and r a , c is the aerodynamic resistance for the canopy (s·m−1). ρ a [81], f c [327] and r a [71] are obtained from Equations (A18), (A19) and (A20), respectively:
ρ a = 1000 P 1.01 T a R
where R is the specific gas constant (287 J·kg−1·K−1), and P is the atmospheric pressure (kPa) calculated by 101.3 [ 293 0.0065 Z 293 ] 5.26 (Z is elevation above sea level (m)).
f c = ( N D V I N D V I m i n N D V I m a x N D V I m i n ) 2
where N D V I m a x and N D V I m i n are the maximum and minimum vegetation indices, respectively.
r a = ln ( z m d z o m ) ln ( z h d z o h ) k 2 u z
where z m and z h are standard heights for measuring the wind speed and humidity (m), respectively, d is the displacement height (m), k is the von Karman constant (equal to 0.41), z o m and z o h are the roughness lengths governing the transfer of momentum and heat/vapor, respectively, which are obtained based on an LULC map using the values given in the literature (s/m), and u is the wind speed at the standard height (m·s−1), i.e., 2 m, and it can be calculated by Equation (A21) (Allen et al., 1998):
u = u z   [ 4.87 ln ( 67.8 z 5.42 ) ]
where u z is the wind speed (m/s) at height z (m) in which anemometers are located. In addition, d is calculated based on Allen et al. [79] using the plant height, i.e., d = 0.666 h . Similar to z o m , h is also obtained from an LULC map, and the values given in the literature (m).
4.
Latent heat flux ( λ E T or LE)
The latent heat flux ( λ E T or LE) is defined as the required energy for converting the water phase from the liquid to the vapor. Similar to H, LE is also obtained from the flux-gradient form for approach 2 as shown below:
L E = L E c + L E s = f c ( ρ a C p γ ( r a , c + r c ) ) ( e s e a ) + ( 1 f c ) ( ρ a C p γ ( r a , s + r s , s ) ) ( e s e a )
where L E c and L E s are latent heat fluxes for the canopy and soil (W·m−2), respectively, e s is the saturation vapor pressure at LST (kPa), γ is the psychometric constant (kPa· ° C 1 ), which is calculated by 0.665 × 10 3 × P , r c is the surface resistance for the canopy (s·m−1), and r s , s represents the surface resistance for the soil (s·m−1). According to Allen et al. [81], e s is computed by Equation (A23):
e s = 0.611 exp [ 17.27 L S T L S T + 237.3 ]
where L S T is the surface temperature (°C). Equations (A24) and (A25) are used to estimate r s , s and r c , respectively:
r s . s = e x p ( 8.206 4.255 ( w z w s a t ) )
where w z is the moisture content at depth z (m3/m3), and w s a t is the saturation moisture content (m3/m3).
r c   o r   r s . c = r c , m i n L A I F 1 1 F 2 1 F 3 1 F 4 1
where r c , m i n is the minimum stomatal resistance (s·m−1), and parameters F1, F2, F3, and F4 are environmental stresses (dimensionless), representing the effects of the active photosynthesis radiation, water stress, deficiency of atmospheric vapor pressure, and air temperature on the surface resistance, respectively, and are calculated using Equations (A26)–(A29), which were proposed by Sellers et al. [112], Noilhan and Planton [90], Sellers et al. [112], and Dickinson [109]:
F 1 = f + r c , m i n / r c , m a x 1 + f     ,     f = 0.55 R s , i n R s . m i n , L 2 L A I
F 2 = { 1             i f                   w c r < w z     w z w w i l t w c r w w i l t   i f   w w i l t < w z < w c r 0               i f         w z < w w i l t          
F 3 = 1 g ( e s ( L S T ) e a )
F 4 = 1 0.0016 ( 298 T a ) 2
where rc,min and rc,max are the minimum and maximum values of canopy resistance (s/m), Rs,min,L is a limit value of incident radiation (W/m2), wcr is the field capacity (m3/m3) close to 0.75 wsat [328], wwilt is the soil moisture at the wilting point (m3/m3), and g is an experimental parameter that depends on the plant type, mostly considered as equal to 0.025 (h/pa).
Approach 1 uses the PM equation to estimate LE, as shown below:
L E = L E c + L E s = Δ ( R n G ) + ρ a C p ( e s e a ) / r a , c Δ + γ ( 1 + r s , c r a , c ) + Δ ( R n ) + ρ a C p ( e s e a ) / r a , s Δ + γ ( 1 + r s , s r a , s )
where Δ is the slope of the vapor pressure curve (kPa·°C−1), e s is the saturation vapor pressure (kPa), γ is the psychrometric constant (kPa·°C−1), r a , c and r a , s are the aerodynamic resistance for the crop and soil (s/m), and r s , c and r s , s are the surface resistance for the crop and soil (s/m). Δ , e s , and r s . c are obtained from Equations (A31)–(A33) [79]:
Δ = 4098 ( 0.6108 e x p ( 17.27 T a T a + 237.3 ) ( T a + 237.3 ) 2
e s = e 0 ( T a , m a x ) + e 0 ( T a , m i n ) 2
r s . c = r 1 L A I a c t i v e
where T a is the air temperature (°C), e 0 ( T a , m a x ) is the saturation vapor pressure at the maximum air temperature (kPa), e 0 ( T a , m i n ) is the saturation vapor pressure at the minimum air temperature (kPa), r 1 is the stomatal resistance for a well-conditioned leaf (s·m−1), and   L A I a c t i v e is the fraction of LAI that contributes to the transfer of vapor and heat (-). According to Allent et al. [79], r 1 and L A I a c t i v e are equal to 100 (s·m−1) and 0.5LAI (-), respectively.
Flux LE is calculated using approaches 1, 2, and 3 for a station with x = 49.82 m, y = 31.52 m, and z = 695 m as follows:
Approach 1:
Determining Rs,in
InputsLatitude = 31.529   decimal   degree   φ = 31.5291 π 180   = 0.55 (rad)
DoY = 166 (-)
S 0 = 0.0820 (MJ·m−2·day−1)
a s = 0.25 (-)
b s = 0.5 (-)
n = 11.38 (h)
N = 14.045 (h)
From Equation (A5) d = 1 + 0.033 cos ( 2 π × 166 365 ) =   0.968rad
From Equation (A6) δ = 0.4093 sin ( 2 π × 166 365 1.39 ) =   0.407rad
From Equation (A7) ω s = cos 1 ( tan ( 0.4071 ) × tan ( 0.55 ) ) = 1.838rad
From Equation (A8) N = 24 π × 1.838 = 14.045hour
From Equation (A4) R a = 24 × 60 π × 0.0820 × 0.968 × [ 1.838 × sin ( 0.55 ) × sin ( 0.407 ) + cos 0.55 × cos ( 0.407 ) × sin ( 1.838 ) ] = 41.327 MJ·m−2·day−1
From Equation (A3) R s , i n = ( 0.25 + 0.5 11.38   14.045 ) × 41.327 × 11.6 =   314.063W·m−2
Determining Rnl
Inputs R s , in =   314.063 (W·m−2)
a s = 0.25 (-)
b s = 0.5 (-)
n = 11.38 (h)
N = 14.045 (h)
R a = 41.327 (MJ·m−2·day−1)
Ta,min   =   24.47 ( )
Ta,max   =   43.09 ( )
σ = 5.67 × 10−8 (W·m−2·K−4)
RH = 13.23 (%)
e 0 ( T a , m a x ) = 0.611 exp [ 17.27 T a , m a x T a , m a x + 237.3 ] = 0.611 exp [ 17.27 × 43.09 43.09 + 237.3 ] = 8.683kPa
e 0 ( T a , m i n ) = 0.611 exp [ 17.27 T a , m i n T a , m i n + 237.3 ] = 0.611 exp [ 17.27 × 24.47 24.47 + 237.3 ] = 3.07kPa
From Equation (A13) e a = R H 100 [ e 0 ( T a , m a x ) + e 0 ( T a , m i n ) 2 ] = 13.23 100 × [ 8.683 + 3.07 2 ] = 0.777kPa
R s o = ( a s + b s ) × R a = ( 0.25 + 0.5 ) × 41.327 × 11.6 = 359.544W·m−2
From Equation (A12) R n l =   5.67×10−8   × [ ( 43.09 + 273.15 ) 4 + ( 24.47 + 273.15 ) 4 2 ] ( 0.34 0.14 3.07 ) ( 1.35 314.063 359.544 0.35 ) = 74.206W·m−2
Determining Rn
Inputs α = 0.24 (-)
Rs,in = 314.063 (W·m−2)
R nl = 74.206 (W·m−2)
From Equation (A1) R n = ( 1 0.24 ) × 314.063 74.206 =   164.481W·m−2
Determining G
InputsNDVI = 0.11 (-)
NDVImax = 0.1 (-)
NDVImin = 0.31 (-)
Γ C = 0.05 (-)
Γ s = 0.315 (-)
R n = 164.481 (W·m−2)
From Equation (A19) f c   =   ( 0.11   0.1 0.31   0.1 ) 2 =   0.002268-
From Equation (A15) G =   164.481 × [ 0.05 + ( 1 0.002268 ) ( 0.315   0.05 ) ] =   51.71W·m−2
Determining LE
Inputs R = 287 (J·kg−1·K−1)
R n = 164.481 (W·m−2)
G =   51.71 (W·m−2)
Z = 695 (m)
e 0 ( T a , max ) = 8.683 (kPa)
e 0 ( T a , min ) = 3.07 (kPa)
Ta = 33.78 ( )
C p = 1004 (J·kg−1·K−1)
e a = 0.777 (kPa)
γ = 0.062 (kPa· 1 )
r 1 = 100 (s·m−1)
LAI = 0.19 (-)
z m = z h = 2 (m)
z om = { 0.16       ( m )                 crop         0.001   ( m )               soil                   [329]
h = { 0.5       ( m )                 crop         0.0   ( m )               soil                   [329]
z oh = 0.1 × z om = { 0.016       ( m )                 crop         0.0001   ( m )               soil                   [81]
For crop: d = 0.666 × h = 0.666 × 0.5 = 0.333 (m)
For soil: d = 0.666 × h = 0.666 × 0 = 0.0 (m)
k = 0.41 (-)
u z = 1.138 (m·s−1)
w sat   = 0.5276 (m3/m3)
w z = 0.03 (m3/m3)
From Equation (A32) e s = 8.683 + 3.07 2 = 5.876kPa
From Equation (A18) ρ a = 1000 1.01 × ( 33.78 + 273.15 ) × 287 × ( 101.3 [ 293 0.0065 × 695 293 ] 5.26 ) = 1.04Kg·m−3
From Equation (A21) u = u z   [ 4.87 log ( 67.8 × 10 5.42 )   ] = 0.85m·s−1
From Equation (A20) r a , c = ln ( 2 0.333 0.16 ) ln ( 2 0.333 0.016   ) 0.41 2 × 0.85 = 76.06s·m−1
From Equation (A20) r a , s = ln ( 2 0 0.001 ) ln ( 2 0 0.0001   ) 0.41 2 × 0.85 = 525.83s·m−1
From Equation (A33) r s . c = 100 0.5 × 0.19 = 1052.632s·m−1
From Equation (A24) r s . s = e x p ( 8.206 4.255 ( 0.03 0.5276 ) ) = 2875.71s·m−1
From Equation (A31) Δ = 4098 ( 0.6108 e x p ( 17.27 × 33.78 33.78 + 237.3 ) ( 33.78 + 237.3 ) 2 = 0.293kPa·°C−1
γ = 0.665 × 10 3 × P = 0.665 × 10 3 × 101.3 [ 293 0.0065 × 695 293 ] 5.26 = 0.062kPa. 1
From Equation (A30) L E = L E c + L E s = 0.293 × ( 164.481 51.71   ) + 1.04 × 1004 ( 5.876   0.777 ) / 76.06 0.293 +   0.062 × ( 1 + 1052.632 76.06 ) + 0.293 × ( 164.481 ) + 1.04 × 1004 ( 5.876   0.777 ) / 525.83 0.293 +   0.062 × ( 1 + 2875.71 525.83 ) = 168.96 W·m−2
Approach 2:
Determining Rs,in
Inputsn = 11.38 (h)
N = 14.045 (h)
Pr = 0 (mm)
RH = 13.23 (%)
Ta = 33.78 ( )
R a = 41.327 (MJ·m−2·day−1)
From Equation (A2) R s , i n = 41.327 × [ 0.4677 + 0.3279 ( 11.38 14.045 ) + 0.0001 ( 0 ) 0.0019 ( 13.23 ) 0.0024 ( 33.78 ) ]   × 11.6 = 301.361W·m−2
Determining Rl,in
Inputs σ = 5.67 × 10−8 (W·m−2·K−4)
Ta = 33.78 ( )
ε a = 92 × 10 7   T a 2 = 92 × 10 7   ( 33.78 + 273.15 ) 2 = 0.866-
From Equation (A9) R l . i n = 0.866 × 5.67 × 10 8 ( 33.78 + 273.15 ) 4 =   436.1209W·m−2
Determining Rl,out
InputsNDVI = 0.11 (-)
LAI = 0.19 (-)
σ = 5.67 × 10−8 (W·m−2·K−4)
LST = unknown (K)
From Equation (A11) ε s = 0.95 + 0.01 × 0.19 = 0.9519-
From Equation (A10) R l . o u t = 0.9519 × 5.67 × 10 8 ( LST ) 4 =   5.397 × 10−8 LST 4 W·m−2
Determining Rn
Inputs α = 0.24 (-)
Rs,in = 301.361 (W·m−2)
Rl,in = 436.1209 (W·m−2)
Rl,out = 5.397 × 10−8 LST 4 (W·m−2)
From Equation (A1) R n = ( 1 0.24 ) × 301.361 + 436.1209 5.397 e−8 LST 4 = 665.155 5.397 × 10−8 LST 4 W·m−2
Determining G
Inputs Γ C = 0.05 (-)
Γ s = 0.315 (-)
f c   =   0.002268 (-)
R n = 665.155 5.397 × 10−8 LST 4 (W·m−2)
From Equation (A15)G = (665.155 5.397 × 10−8 LST 4 ) × [ 0.05 + ( 1 0.002268 ) ( 0.315   0.05 ) ] =   209.124–1.6968 × 10−8 LST4W·m−2
Determining H
Inputs f c   =   0.002268 (-)
ρ a   = 1.04 (Kg.m−3)
Ta = 33.78 ( )
C p = 1004 (J·kg−1·K−1)
LST = unknown   ( )
r a , c = 76.06 (s·m−1)
r a , s = 525.83 (s·m−1)
From Equation (A17) H = ( 1 0.002268 ) ( 1.04 × 1004 525.83 ) ( ( LST + 273.15 ) ( 33.78 + 273.15 ) ) + 0.002268 ( 1.04 × 1004 76.06 ) ( ( LST + 273.15 ) ( 33.78 + 273.15 ) ) = 1.948 × ( LST - 33.78 ) (LST)W·m−2
Determining LE
Inputs γ = 0.062 (kPa. 1 )
f c   =   0.002268 (-)
w sat   = 0.5276 (m3/m3)
w z = 0.03 (m3/m3)
ρ a   = 1.04 (Kg.m−3)
Ta = 33.78 ( )
C p = 1004 (J.kg−1.K−1)
LST = unknown ( )
r a , c = 76.06 (s·m−1)
r a , s = 525.83 (s·m−1)
rc,min = 100 (s·m−1)
rc,max = 5000 (s·m−1)
Rs,min,L = 100 (W/m2)
LAI = 0.19 (-)
R s , in = 301.361 (W/m2)
w wilt = 0.1624 (m3/m3)
w cr = 0.75 w sat   = 0.3957 (m3/m3)
g = 0.025 (h/pa)
r s . s = 2875.71 (s·m−1)
e a = 0.777 (kPa)
From Equation (A23) e s = 0.611 e x p [ 17.27 L S T L S T + 237.3 ] kPa
From Equation (A26) F 1 = ( 0.55 × 301.361 100 × 2 0.19 ) + ( 100 / 5000 ) 1 + ( 0.55 × 301.361 100 × 2 0.19 ) =   0.9468-
From Equation (A27) F 2 =   { 1                           i f                   0.3957 < 0.03     0.03 0.1624   0.3957 0.1624         i f   0.1624 < 0.03 < 0.3957 0                               i f         0.03 < 0.1624           = 0-
From Equation (A28) F 3 = 1 0.025 × ( 0.611 e x p [ 17.27 L S T L S T + 237.3 ] 0.777 ) -
From Equation (A29) F 4 = 1 0.0016 × ( 298 ( 33.78 + 273.15 ) ) 2 = 0.872-
From Equation (A25) r s . c = 100 0.19 × ( 0.9468 ) 1 × ( 0 ) 1 × ( 1 0.025 ( 0.611 e x p [ 17.27 L S T L S T + 237.3 ] 0.777 ) ) 1 × ( 0.872 ) 1   = ∞s·m−1
From Equation (A22) L E = 0.002268 ( 1.04 × 1004 0.062 × ( 76.06 + ) ) ( 0.611 e x p [ 17.27 L S T L S T + 237.3 ] 0.777 ) + ( 1 0.002268 ) ( 1.04 × 1004 0.062 × ( 525.83 + 2875.71 ) ) ( 0.611 e x p [ 17.27 L S T L S T + 237.3 ] 0.777 ) W/m2
Determining LE and LST using the Newton–Raphson method
Inputs Cs (volumetric heat capacity of soil) = 1268828.50 (J. m 3 K 1 )
LST t (LST at month t) = unknown+273.15 (K)
LST t 1 (LST at month t-1) = 39.45+273.15 (K)
Thickness (soil depth) = 0.1 m
R n = 665.1555.397 × 10−8 LST 4
G = 209.124–1.6968 × 10−8 LST4
H = 1.948 × (LST-33.78)
LE = 0.002268 ( 1.04 × 1004 0.062 × ( 76.06 + ) ) ( 0.611 e x p [ 17.27 L S T L S T + 237.3 ] 0.777 ) + ( 1 0.002268 ) ( 1.04 × 1004 0.062 × ( 525.83 + 2875.71 ) ) ( 0.611 e x p [ 17.27 L S T L S T + 237.3 ] 0.777 )
R n G H L E ( C s × T h i c k n e s s × ( L S T t   L S T t 1 ) × 11.6 / 1000000 ) = 0 → Newton Raphson → LST = 45.27 W/m2
From Equation (A22) L E = 0.002268 ( 1.04 × 1004 0.062 × ( 76.06 + ) ) ( 0.611 e x p [ 17.27 × 45.27 45.27 + 237.3 ] 0.777 ) + ( 1 0.002268 ) ( 1.04 × 1004 0.062 × ( 525.83 + 2875.71 ) ) ( 0.611 e x p [ 17.27 × 45.27 45.27 + 237.3 ] 0.777 ) = 44.63 W/m2
Approach 3:
Determining Rs,in
Inputsn = 11.38 (h)
N = 14.045 (h)
Pr = 0 (mm)
RH = 13.23 (%)
Ta = 33.78 ( )
R a = 41.327 (MJ·m−2·day−1)
From Equation (A2) R s , i n = 41.327 × [ 0.4677 + 0.3279 ( 11.38 14.045 ) + 0.0001 ( 0 ) 0.0019 ( 13.23 ) 0.0024 ( 33.78 ) ] × 11.6 = 301.361 W·m−2
Determining Rl,in
Inputs σ = 5.67 × 10−8 (W·m−2·K−4)
Ta = 33.78 ( )
ε a = 92 × 10 7   ( 33.78 + 273.15 ) 2 = 0.866-
From Equation (A9) R l . i n = 0.866 × 5.67 × 10 8 ( 33.78 + 273.15 ) 4 =   436.1209W·m−2
Determining Rl,out
InputsNDVI = 0.11 (-)
LAI = 0.19 (-)
σ = 5.67 × 10−8 (W·m−2·K−4) LST = 42.53 ( )
From Equation (A11) ε s = 0.95 + 0.01   × 0.19 =   0.9519-
From Equation (A10) R l . o u t = 0.9519 × 5.67 × 10 8 ( 42.53 + 273.15 ) 4 =   535.99W·m−2
Determining Rn
Inputs α = 0.24 (-)
Rs,in = 301.361 (W·m−2)
Rl,in = 436.1209 (W·m−2)
Rl,out = 490.013 (W·m−2)
From Equation (A1) R n = ( 1 0.24 ) × 301.361 + 436.1209 535.99 = 129.157W·m−2
Determining G
InputsNDVI = 0.11 (-)
NDVImax = 0.1 (-)
NDVImin = 0.31 (-)
Γ C = 0.05 (-)
Γ s = 0.315 (-)
f c =   0.002268 (-)
From Equation (A15) G =   129.157 × [ 0.05 + ( 1 0.002268 ) ( 0.315   0.05 ) ] =   40.6W·m−2
Determining H
InputsTa = 33.78 ( )
C p = 1004 (J. kg−1.K−1)
LST = 42.53 ( )
For crop → r a = 76.06 (s·m−1)
For soil → r a = 525.83 (s·m−1)
ρ a   = 1.04 (Kg.m−3)
f c =   0.002268 (-)
From Equation (A17) H = ( 1 0.002268 ) ( 1.04 × 1004 525.83 ) ( ( 35.53 + 273.15 ) ( 33.78 + 273.15 ) ) + 0.002268 ( 1.04 × 1004 76.06 ) ( ( 35.53 + 273.15 ) ( 33.78 + 273.15 ) ) = 13.58 W·m−2
Determining LE
Inputs R n = 129.157 (W·m−2)
H = 13.58 (W·m−2)
G =   40.6 (W·m−2)
Residual approach L E =   R n H G =   74.96W·m−2

References

  1. Jung, M.; Reichstein, M.; Ciais, P.; Seneviratne, S.I.; Sheffield, J.; Goulden, M.L.; Bonan, G.; Cescatti, A.; Chen, J.; De Jeu, R. Recent decline in the global land evapotranspiration trend due to limited moisture supply. Nature 2010, 467, 951–954. [Google Scholar] [CrossRef] [PubMed]
  2. Doorenbos, J. Guidelines for Predicting Crop Water Requirements; Food and Agriculture organization: Rome, Italy, 1975; Irrig. Drainage paper; Volume 24. [Google Scholar]
  3. Singh, R.K.; Senay, G.B. Comparison of four different energy balance models for estimating evapotranspiration in the Midwestern United States. Water 2016, 8, 9. [Google Scholar] [CrossRef] [Green Version]
  4. Madugundu, R.; Al-Gaadi, K.A.; Tola, E.; Hassaballa, A.A.; Patil, V.C. Performance of the METRIC model in estimating evapotranspiration fluxes over an irrigated field in Saudi Arabia using Landsat-8 images. Hydrol. Earth Syst. Sci. 2017, 21, 6135–6151. [Google Scholar] [CrossRef] [Green Version]
  5. Su, Z. The Surface Energy Balance System (SEBS) for estimation of turbulent heat fluxes. Hydrol. Earth Syst. Sci. 2002, 6, 85–100. [Google Scholar] [CrossRef]
  6. Kairu, E. A review of methods for estimating evapotranspiration. GeoJournal 1991, 25, 371–376. [Google Scholar] [CrossRef]
  7. Xu, C.-Y.; Singh, V. Cross comparison of empirical equations for calculating potential evapotranspiration with data from Switzerland. Water Resour. Manag. 2002, 16, 197–219. [Google Scholar] [CrossRef]
  8. Wróblewski, P.; Drożdż, W.; Lewicki, W.; Miązek, P. Methodology for assessing the impact of aperiodic phenomena on the energy balance of propulsion engines in vehicle electromobility systems for given areas. Energies 2021, 14, 2314. [Google Scholar] [CrossRef]
  9. Penman, H.L. Natural evaporation from open water, bare soil and grass. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1948, 193, 120–145. [Google Scholar]
  10. Monteith, J. Evaporation and environment. The state and movement of water in living organisms. In Symposium of the Society of Experimental Biology; Cambridge University Press (CUP): Cambridge, UK, 1965; Volume 19, pp. 205–234. [Google Scholar]
  11. Luo, X.; Chen, J.M.; Liu, J.; Black, T.A.; Croft, H.; Staebler, R.; He, L.; Arain, M.A.; Chen, B.; Mo, G. Comparison of big-leaf, two-big-leaf, and two-leaf upscaling schemes for evapotranspiration estimation using coupled carbon-water modeling. J. Geophys. Res. Biogeosci. 2018, 123, 207–225. [Google Scholar] [CrossRef]
  12. Alves, I.; Perrier, A.; Pereira, L. Aerodynamic and surface resistances of complete cover crops: How good is the “big leaf”? Trans. ASAE 1998, 41, 345. [Google Scholar] [CrossRef]
  13. Verstraeten, W.W.; Veroustraete, F.; Feyen, J. Assessment of evapotranspiration and soil moisture content across different scales of observation. Sensors 2008, 8, 70–117. [Google Scholar] [CrossRef] [Green Version]
  14. 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]
  15. Overgaard, J.; Rosbjerg, D.; Butts, M. Land-surface modelling in hydrological perspective–a review. Biogeosciences 2006, 3, 229–241. [Google Scholar] [CrossRef] [Green Version]
  16. Schultz, K.; Franks, S.; Beven, K. TOPUP–A TOPMODEL based SVAT Model to Calculate Evaporative Fluxes between the Land Surface and the Atmosphere, Version 1.1, Program Documentation; Department of Environmental Science, Lancaster University: Lancaster, UK, 1998. [Google Scholar]
  17. Mauser, W.; Schädlich, S. Modelling the spatial distribution of evapotranspiration on different scales using remote sensing data. J. Hydrol. 1998, 212, 250–267. [Google Scholar] [CrossRef]
  18. Shuttleworth, W.J.; Wallace, J. Evaporation from sparse crops-an energy combination theory. Q. J. R. Meteorol. Soc. 1985, 111, 839–855. [Google Scholar] [CrossRef]
  19. Corbari, C.; Mancini, M. Calibration and validation of a distributed energy–water balance model using satellite data of land surface temperature and ground discharge measurements. J. Hydrometeorol. 2014, 15, 376–392. [Google Scholar] [CrossRef]
  20. Corbari, C.; Ravazzani, G.; Mancini, M. A distributed thermodynamic model for energy and mass balance computation: FEST–EWB. Hydrol. Process. 2011, 25, 1443–1452. [Google Scholar] [CrossRef]
  21. Wu, B.; Zhu, W.; Yan, N.; Xing, Q.; Xu, J.; Ma, Z.; Wang, L. Regional actual evapotranspiration estimation with land and meteorological variables derived from multi-source satellite data. Remote Sens. 2020, 12, 332. [Google Scholar] [CrossRef] [Green Version]
  22. Crow, W.T.; Wood, E.F. The assimilation of remotely sensed soil brightness temperature imagery into a land surface model using ensemble Kalman filtering: A case study based on ESTAR measurements during SGP97. Adv. Water Resour. 2003, 26, 137–149. [Google Scholar] [CrossRef]
  23. Famiglietti, J.; Wood, E.F. Multiscale modeling of spatially variable water and energy balance processes. Water Resour. Res. 1994, 30, 3061–3078. [Google Scholar] [CrossRef] [Green Version]
  24. Lakshmi, V. A simple surface temperature assimilation scheme for use in land surface models. Water Resour. Res. 2000, 36, 3687–3700. [Google Scholar] [CrossRef] [Green Version]
  25. Kumar, P.; Kaleita, A.L. Assimilation of Surface Temperature in a Land-Surface Model; IAHS-AISH Publications: Wallingford, UK, 2000; pp. 197–201. [Google Scholar]
  26. Gutmann, E.D.; Small, E.E. A method for the determination of the hydraulic properties of soil from MODIS surface temperature for use in land-surface models. Water Resour. Res. 2010, 46, 46. [Google Scholar] [CrossRef] [Green Version]
  27. Corbari, C.; Sobrino, J.; Mancini, M.; Hidalgo, V. Land surface temperature representativeness in a heterogeneous area through a distributed energy-water balance model and remote sensing data. Hydrol. Earth Syst. Sci. 2010, 14, 2141–2151. [Google Scholar] [CrossRef] [Green Version]
  28. Corbari, C.; Sobrino, J.A.; Mancini, M.; Hidalgo, V. Mass and energy flux estimates at different spatial resolutions in a heterogeneous area through a distributed energy–water balance model and remote-sensing data. Int. J. Remote Sens. 2013, 34, 3208–3230. [Google Scholar] [CrossRef]
  29. 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]
  30. Calcagno, G.; Mendicino, G.; Monacelli, G.; Senatore, A.; Versace, P. Distributed estimation of actual evapotranspiration through remote sensing techniques. In Methods and Tools for Drought Analysis and Management; Springer: Berlin/Heidelberg, Germany, 2007; pp. 125–147. [Google Scholar]
  31. Taheri, M.; Emadzadeh, M.; Gholizadeh, M.; Tajrishi, M.; Ahmadi, M.; Moradi, M. Investigating the temporal and spatial variations of water consumption in Urmia Lake River Basin considering the climate and anthropogenic effects on the agriculture in the basin. Agric. Water Manag. 2019, 213, 782–791. [Google Scholar] [CrossRef]
  32. Shilpakar, R.L.; Bastiaanssen, W.G.; Molden, D.J. A remote sensing-based approach for water accounting in the East Rapti River Basin, Nepal. Himal. J. Sci. 2011, 7, 15–30. [Google Scholar] [CrossRef] [Green Version]
  33. Evans, R.; Hulbert, S.; Murrihy, E.; Bastiaanssen, W.M.R.; Molloy, R. Using satellite imagery to measure evaporation from storages–solving the great unknown in water accounting. In Proceedings of the Irrigation and Drainage Conference; Irrigation Australia Ltd.: Swan Hill, VIC, Australia, 2009. [Google Scholar]
  34. Karimi, P.; Bastiaanssen, W.G. Spatial evapotranspiration, rainfall and land use data in water accounting–Part 1: Review of the accuracy of the remote sensing data. Hydrol. Earth Syst. Sci. 2015, 19, 507–532. [Google Scholar] [CrossRef] [Green Version]
  35. Zwart, S.J.; Bastiaanssen, W.G.; de Fraiture, C.; Molden, D.J. WATPRO: A remote sensing based model for mapping water productivity of wheat. Agric. Water Manag. 2010, 97, 1628–1636. [Google Scholar] [CrossRef]
  36. Bastiaanssen, W.G.; Menenti, M.; Feddes, R.; Holtslag, A. A remote sensing surface energy balance algorithm for land (SEBAL). 1. Formulation. J. Hydrol. 1998, 212, 198–212. [Google Scholar] [CrossRef]
  37. Roerink, G.; Su, Z.; Menenti, M. S-SEBI: A simple remote sensing algorithm to estimate the surface energy balance. Phys. Chem. Earth Part B Hydrol. Ocean. Atmos. 2000, 25, 147–157. [Google Scholar] [CrossRef]
  38. Senay, G.B.; Budde, M.; Verdin, J.P.; Melesse, A.M. A coupled remote sensing and simplified surface energy balance approach to estimate actual evapotranspiration from irrigated fields. Sensors 2007, 7, 979–1000. [Google Scholar] [CrossRef] [Green Version]
  39. Senay, G.B.; Bohms, S.; Singh, R.K.; Gowda, P.H.; Velpuri, N.M.; Alemu, H.; Verdin, J.P. Operational evapotranspiration mapping using remote sensing and weather datasets: A new parameterization for the SSEB approach. JAWRA J. Am. Water Resour. Assoc. 2013, 49, 577–591. [Google Scholar] [CrossRef] [Green Version]
  40. 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]
  41. 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]
  42. Norman, J.; Kustas, W.; Prueger, J.; Diak, G. Surface flux estimation using radiometric temperature: A dual-temperature-difference method to minimize measurement errors. Water Resour. Res. 2000, 36, 2263–2274. [Google Scholar] [CrossRef] [Green Version]
  43. Yang, Y.; Su, H.; Zhang, R.; Tian, J.; Li, L. An enhanced two-source evapotranspiration model for land (ETEML): Algorithm and evaluation. Remote Sens. Environ. 2015, 168, 54–65. [Google Scholar] [CrossRef]
  44. Zhou, Y.; Li, X. Energy balance closures in diverse ecosystems of an endorheic river basin. Agric. For. Meteorol. 2019, 274, 118–131. [Google Scholar] [CrossRef]
  45. Grachev, A.A.; Fairall, C.W.; Blomquist, B.W.; Fernando, H.J.; Leo, L.S.; Otárola-Bustos, S.F.; Wilczak, J.M.; McCaffrey, K.L. On the surface energy balance closure at different temporal scales. Agric. For. Meteorol. 2020, 281, 107823. [Google Scholar] [CrossRef]
  46. Foken, T.; Oncley, S. Workshop on instrumental and methodical problems of land surface flux measurements. Bull. Am. Meteorol. Soc. 1995, 1191–1193. [Google Scholar] [CrossRef] [Green Version]
  47. Foken, T. The energy balance closure problem: An overview. Ecol. Appl. 2008, 18, 1351–1367. [Google Scholar] [CrossRef] [PubMed]
  48. Foken, T.; Mauder, M.; Liebethal, C.; Wimmer, F.; Beyrich, F.; Leps, J.-P.; Raasch, S.; DeBruin, H.A.; Meijninger, W.M.; Bange, J. Energy balance closure for the LITFASS-2003 experiment. Theor. Appl. Climatol. 2010, 101, 149–160. [Google Scholar] [CrossRef]
  49. Leuning, R.; Van Gorsel, E.; Massman, W.J.; Isaac, P.R. Reflections on the surface energy imbalance problem. Agric. For. Meteorol. 2012, 156, 65–74. [Google Scholar] [CrossRef]
  50. Stoy, P.C.; Mauder, M.; Foken, T.; Marcolla, B.; Boegh, E.; Ibrom, A.; Arain, M.A.; Arneth, A.; Aurela, M.; Bernhofer, C. A data-driven analysis of energy balance closure across FLUXNET research sites: The role of landscape scale heterogeneity. Agric. For. Meteorol. 2013, 171, 137–152. [Google Scholar] [CrossRef] [Green Version]
  51. McGloin, R.; Šigut, L.; Havránková, K.; Dušek, J.; Pavelka, M.; Sedlák, P. Energy balance closure at a variety of ecosystems in Central Europe with contrasting topographies. Agric. For. Meteorol. 2018, 248, 418–431. [Google Scholar] [CrossRef]
  52. Mauder, M.; Desjardins, R.L.; Pattey, E.; Worth, D. An attempt to close the daytime surface energy balance using spatially-averaged flux measurements. Bound. Layer Meteorol. 2010, 136, 175–191. [Google Scholar] [CrossRef]
  53. Xu, Z.; Ma, Y.; Liu, S.; Shi, W.; Wang, J. Assessment of the energy balance closure under advective conditions and its impact using remote sensing data. J. Appl. Meteorol. Climatol. 2017, 56, 127–140. [Google Scholar] [CrossRef]
  54. Mauder, M.; Foken, T.; Cuxart, J. Surface-energy-balance closure over land: A review. Bound. Layer Meteorol. 2020, 177, 395–426. [Google Scholar] [CrossRef]
  55. Katerji, N.; Rana, G.; Fahed, S. Parameterizing canopy resistance using mechanistic and semi-empirical estimates of hourly evapotranspiration: Critical evaluation for irrigated crops in the Mediterranean. Hydrol. Process. 2011, 25, 117–129. [Google Scholar] [CrossRef]
  56. Yan, H.; Zhang, C.; Oue, H.; Sugimoto, H. Comparison of different methods for estimating soil surface evaporation in a bare field. Meteorol. Atmos. Phys. 2012, 118, 143–149. [Google Scholar] [CrossRef]
  57. Yan, H.; Shi, H.; Hiroki, O.; Zhang, C.; Xue, Z.; Cai, B.; Wang, G. Modeling bulk canopy resistance from climatic variables for predicting hourly evapotranspiration of maize and buckwheat. Meteorol. Atmos. Phys. 2015, 127, 305–312. [Google Scholar] [CrossRef]
  58. Rana, G.; Katerji, N.; Mastrorilli, M.; El Moujabber, M.; Brisson, N. Validation of a model of actual evapotranspiration for water stressed soybeans. Agric. For. Meteorol. 1997, 86, 215–224. [Google Scholar] [CrossRef]
  59. Alves, I.; Pereira, L.S. Modelling surface resistance from climatic variables? Agric. Water Manag. 2000, 42, 371–385. [Google Scholar] [CrossRef]
  60. Steduto, P.; Todorovic, M.; Caliandro, A.; Rubino, P. Daily reference evapotranspiration estimates by the Penman-Monteith equation in Southern Italy. Constant vs. variable canopy resistance. Theor. Appl. Climatol. 2003, 74, 217–225. [Google Scholar] [CrossRef]
  61. Farahani, H.J.; Howell, T.A.; Shuttleworth, W.J.; Bausch, W.C. Evapotranspiration: Progress in measurement and modeling in agriculture. Trans. ASABE 2007, 50, 1627–1638. [Google Scholar] [CrossRef]
  62. Jarvis, P. The interpretation of the variations in leaf water potential and stomatal conductance found in canopies in the field. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1976, 273, 593–610. [Google Scholar]
  63. Stewart, J.B. Modelling surface conductance of pine forest. Agric. For. Meteorol. 1988, 43, 19–35. [Google Scholar] [CrossRef]
  64. Dolman, A. A multiple-source land surface energy balance model for use in general circulation models. Agric. For. Meteorol. 1993, 65, 21–45. [Google Scholar] [CrossRef]
  65. Liu, S.; Lu, L.; Mao, D.; Jia, L. Evaluating parameterizations of aerodynamic resistance to heat transfer using field measurements. Hydrol. Earth Syst. Sci. 2007, 11, 769–783. [Google Scholar] [CrossRef] [Green Version]
  66. Monteith, J. Principles of Environmental Physics; Edward Arnold: London, UK, 1973; 214p. [Google Scholar]
  67. Seguin, B.; Baelz, S.; Monget, J.-M.; Petit, V. Utilisation de la thermographie IR pour l’estimation de l’évaporation régionale II.-Résultats obtenus à partir des données de satellite. Agronomie 1982, 2, 113–115. [Google Scholar] [CrossRef]
  68. Hatfield, J. Evapotranspiration obtained from remote sensing methods. In Advances in Irrigation; Elsevier: Amsterdam, The Netherlands, 1983; Volume 2, pp. 395–416. [Google Scholar]
  69. Choudhury, B.; Reginato, R.; Idso, S. An analysis of infrared temperature observations over wheat and calculation of latent heat flux. Agric. For. Meteorol. 1986, 37, 75–88. [Google Scholar] [CrossRef]
  70. Moran, M.; Clarke, T.; 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. Thom, A.S. Momentum, mass, and heat exchange of plant communities. Veg. Atmos. 1975, 1, 57–109. [Google Scholar]
  72. Brutsaert, W.H. Evaporation into the Atmosphere; D. Reidel: London, UK, 1982. [Google Scholar]
  73. Monteith, J.; Unsworth, M. 1990: Principles of Environmental Physics; Edward Arnold: London, UK, 1990. [Google Scholar]
  74. Manabe, S. Climate and the ocean circulation: I. The atmospheric circulation and the hydrology of the earth’s surface. Mon. Weather Rev. 1969, 97, 739–774. [Google Scholar] [CrossRef]
  75. Budyko, M.I. The heat balance of the earth’s surface. Sov. Geogr. 1961, 2, 3–13. [Google Scholar] [CrossRef]
  76. Milly, P. Potential evaporation and soil moisture in general circulation models. J. Clim. 1992, 5, 209–226. [Google Scholar] [CrossRef] [Green Version]
  77. Delworth, T.; Manabe, S. The influence of soil wetness on near-surface atmospheric variability. J. Clim. 1989, 2, 1447–1462. [Google Scholar] [CrossRef]
  78. Webb, E. Evaluation of evapotranspiration and canopy resistance: An alternative combination approach. In Developments in Agricultural and Managed Forest Ecology; Elsevier: Amsterdam, The Netherlands, 1984; Volume 13, pp. 151–166. [Google Scholar]
  79. Cleugh, H.A.; Leuning, R.; Mu, Q.; Running, S.W. Regional evaporation estimates from flux tower and MODIS satellite data. Remote Sens. Environ. 2007, 106, 285–304. [Google Scholar] [CrossRef]
  80. 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] [Green Version]
  81. 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 Rome 1998, 300, D05109. [Google Scholar]
  82. Beven, K.J.; Quinn, P. Similarity and scale effects in the water balance of heterogeneous area. In Proceedings of the Conference on the Balance of Water--Present and Future, AGMET, Dublin, Ireland, 7–9 September 1994. [Google Scholar]
  83. Franks, S.; Beven, K.J.; Quinn, P.; Wright, I. On the sensitivity of soil-vegetation-atmosphere transfer (SVAT) schemes: Equifinality and the problem of robust calibration. Agric. For. Meteorol. 1997, 86, 63–75. [Google Scholar] [CrossRef]
  84. Mu, Q.; Zhao, M.; Running, S.W. Improvements to a MODIS global terrestrial evapotranspiration algorithm. Remote Sens. Environ. 2011, 115, 1781–1800. [Google Scholar] [CrossRef]
  85. Zhang, K.; Kimball, J.S.; Nemani, R.R.; Running, S.W. A continuous satellite-derived global record of land surface evapotranspiration from 1983 to 2006. Water Resour. Res. 2010, 46. [Google Scholar] [CrossRef] [Green Version]
  86. Leuning, R.; Zhang, Y.; Rajaud, A.; Cleugh, H.; Tu, K. A simple surface conductance model to estimate regional evaporation using MODIS leaf area index and the Penman-Monteith equation. Water Resour. Res. 2008, 44. [Google Scholar] [CrossRef]
  87. Gan, R.; Zhang, Y.; Shi, H.; Yang, Y.; Eamus, D.; Cheng, L.; Chiew, F.H.; Yu, Q. Use of satellite leaf area index estimating evapotranspiration and gross assimilation for Australian ecosystems. Ecohydrology 2018, 11, e1974. [Google Scholar] [CrossRef]
  88. Zhang, Y.; Peña-Arancibia, J.L.; McVicar, T.R.; Chiew, F.H.; Vaze, J.; Liu, C.; Lu, X.; Zheng, H.; Wang, Y.; Liu, Y.Y. Multi-decadal trends in global terrestrial evapotranspiration and its components. Sci. Rep. 2016, 6, 19124. [Google Scholar] [CrossRef] [Green Version]
  89. Avissar, R.; Pielke, R.A. A parameterization of heterogeneous land surfaces for atmospheric numerical models and its impact on regional meteorology. Mon. Weather Rev. 1989, 117, 2113–2136. [Google Scholar] [CrossRef] [Green Version]
  90. Noilhan, J.; Planton, S. A simple parameterization of land surface processes for meteorological models. Mon. Weather Rev. 1989, 117, 536–549. [Google Scholar] [CrossRef]
  91. Mengelkamp, H.-T.; Warrach, K.; Raschke, E. SEWAB–a parameterization of the surface energy and water balance for atmospheric and hydrologic models. Adv. Water Resour. 1999, 23, 165–175. [Google Scholar] [CrossRef]
  92. Zhang, K.; Kimball, J.S.; Running, S.W. A review of remote sensing based actual evapotranspiration estimation. WIREs Water 2016, 3, 834–853. [Google Scholar] [CrossRef]
  93. Heilman, J.; McInnes, K.; Savage, M.; Gesch, R.; Lascano, R. Soil and canopy energy balances in a west Texas vineyard. Agric. For. Meteorol. 1994, 71, 99–114. [Google Scholar] [CrossRef]
  94. Ham, J.; Heilman, J. Aerodynamic and surface resistances affecting energy transport in a sparse crop. Agric. For. Meteorol. 1991, 53, 267–284. [Google Scholar] [CrossRef]
  95. Beven, K. Changing ideas in hydrology—the case of physically-based models. J. Hydrol. 1989, 105, 157–172. [Google Scholar] [CrossRef]
  96. Demarty, J.; Ottle, C.; Braud, I.; Olioso, A.; Frangi, J.; Bastidas, L.; Gupta, H. Using a multiobjective approach to retrieve information on surface properties used in a SVAT model. J. Hydrol. 2004, 287, 214–236. [Google Scholar] [CrossRef]
  97. Bastiaanssen, W.; Cheema, M.; Immerzeel, W.; Miltenburg, I.; Pelgrum, H. Surface energy balance and actual evapotranspiration of the transboundary Indus Basin estimated from satellite measurements and the ETLook model. Water Resour. Res. 2012, 48. [Google Scholar] [CrossRef] [Green Version]
  98. Gu, L.; Shugart, H.H.; Fuentes, J.D.; Black, T.; Shewchuk, S.R. Micrometeorology, biophysical exchanges and NEE decomposition in a two-story boreal forest—Development and test of an integrated model. Agric. For. Meteorol. 1999, 94, 123–148. [Google Scholar] [CrossRef]
  99. Baldocchi, D.; Harley, P. Scaling carbon dioxide and water vapour exchange from leaf to canopy in a deciduous forest. II. Model testing and application. Plant Cell Environ. 1995, 18, 1157–1173. [Google Scholar] [CrossRef]
  100. Stanghellini, C. Transpiration of GREENHOUSE Crops: An Aid to Climate Management; Wageningen University and Research: Wageningen, The Netherlands, 1987. [Google Scholar]
  101. Chen, B.; Chen, J.M.; Ju, W. Remote sensing-based ecosystem–atmosphere simulation scheme (EASS)—Model formulation and test with multiple-year data. Ecol. Model. 2007, 209, 277–300. [Google Scholar] [CrossRef]
  102. Govind, A.; Chen, J.M.; Margolis, H.; Ju, W.; Sonnentag, O.; Giasson, M.-A. A spatially explicit hydro-ecological modeling framework (BEPS-TerrainLab V2. 0): Model description and test in a boreal ecosystem in Eastern North America. J. Hydrol. 2009, 367, 200–216. [Google Scholar] [CrossRef]
  103. Zhao, W.; Li, A. A review on land surface processes modelling over complex terrain. Adv. Meteorol. 2015, 2015, 1–17. [Google Scholar] [CrossRef] [Green Version]
  104. Chen, J.; Chen, B.; Black, T.A.; Innes, J.L.; Wang, G.; Kiely, G.; Hirano, T.; Wohlfahrt, G. Comparison of terrestrial evapotranspiration estimates using the mass transfer and Penman-Monteith equations in land surface models. J. Geophys. Res. Biogeosci. 2013, 118, 1715–1731. [Google Scholar] [CrossRef]
  105. Deardorff, J.W. Dependence of air-sea transfer coefficients on bulk stability. J. Geophys. Res. 1968, 73, 2549–2557. [Google Scholar] [CrossRef]
  106. Monin, A.S.; Obukhov, A.M. Basic laws of turbulent mixing in the surface layer of the atmosphere. Contrib. Geophys. Inst. Acad. Sci. USSR 1954, 151, e187. [Google Scholar]
  107. Deardorff, J.W. Efficient prediction of ground surface temperature and moisture, with inclusion of a layer of vegetation. J. Geophys. Res. Ocean. 1978, 83, 1889–1903. [Google Scholar] [CrossRef] [Green Version]
  108. Dickinson, R.E.; Jaeger, J.; Washington, W.M.; Wolski, R. Boundary Subroutine for the NCAR Global Climate Model; National Center for Atmospheric Research: Boulder, CO, USA, 1981. [Google Scholar]
  109. Dickinson, R.E. Modeling evapotranspiration for three-dimensional global climate models. Clim. Process. Clim. Sensit. 1984, 29, 58–72. [Google Scholar]
  110. Dickinson, R.E. Biosphere/Atmosphere Transfer Scheme (BATS) for the NCAR Community Climate Model; Technical; NCAR: Boulder, CO, USA, 1986. [Google Scholar]
  111. Yang, Z.-L.; Dickinson, R.E. Description of the Biosphere-Atmosphere Transfer Scheme (BATS) for the Soil Moisture Workshop and evaluation of its performance. Glob. Planet. Chang. 1996, 13, 117–134. [Google Scholar] [CrossRef]
  112. Sellers, P.; Mintz, Y.; Sud, Y.e.a.; Dalcher, A. A simple biosphere model (SiB) for use within general circulation models. J. Atmos. Sci. 1986, 43, 505–531. [Google Scholar] [CrossRef] [Green Version]
  113. Koster, R.D.; Suarez, M.J. Modeling the land surface boundary in climate models as a composite of independent vegetation stands. J. Geophys. Res. Atmos. 1992, 97, 2697–2715. [Google Scholar] [CrossRef]
  114. Liang, X.; Wood, E.F.; Lettenmaier, D.P. Surface soil moisture parameterization of the VIC-2L model: Evaluation and modification. Glob. Planet. Chang. 1996, 13, 195–206. [Google Scholar] [CrossRef]
  115. Cherkauer, K.A.; Bowling, L.C.; Lettenmaier, D.P. Variable infiltration capacity cold land process model updates. Glob. Planet. Chang. 2003, 38, 151–159. [Google Scholar] [CrossRef]
  116. Liang, X.; Lettenmaier, D.P.; Wood, E.F.; Burges, S.J. A simple hydrologically based model of land surface water and energy fluxes for general circulation models. J. Geophys. Res. Atmos. 1994, 99, 14415–14428. [Google Scholar] [CrossRef]
  117. Shuttleworth, W. Evaporation in: Maidment, DR Handbook of Hydrology; McGraw-Hill: New York, NY, USA, 1993. [Google Scholar]
  118. Franchini, M.; Pacciani, M. Comparative analysis of several conceptual rainfall-runoff models. J. Hydrol. 1991, 122, 161–219. [Google Scholar] [CrossRef]
  119. Zhao, R. The Xinganjiang Model. In Proceedings of the Hydrological Forecasting Proceedings Oxford Symposium, IASH 129; IAHS-AISH Publications: Wallingford, UK, 1980; pp. 351–356. [Google Scholar]
  120. Mahrt, L.; Ek, M. The influence of atmospheric stability on potential evaporation. J. Appl. Meteorol. Climatol. 1984, 23, 222–234. [Google Scholar] [CrossRef] [Green Version]
  121. Chen, F.; Mitchell, K.; Schaake, J.; Xue, Y.; Pan, H.L.; Koren, V.; Duan, Q.Y.; Ek, M.; Betts, A. Modeling of land surface evaporation by four schemes and comparison with FIFE observations. J. Geophys. Res. Atmos. 1996, 101, 7251–7268. [Google Scholar] [CrossRef] [Green Version]
  122. Ek, M.; Mitchell, K.; Lin, Y.; Rogers, E.; Grunmann, P.; Koren, V.; Gayno, G.; Tarpley, J. Implementation of Noah land surface model advances in the National Centers for Environmental Prediction operational mesoscale Eta model. J. Geophys. Res. Atmos. 2003, 108. [Google Scholar] [CrossRef]
  123. Prata, A.; Caselles, V.; Coll, C.; Sobrino, J.; Ottle, C. Thermal remote sensing of land surface temperature from satellites: Current status and future prospects. Remote Sens. Rev. 1995, 12, 175–224. [Google Scholar] [CrossRef]
  124. Li, Z.-L.; Becker, F. Feasibility of land surface temperature and emissivity determination from AVHRR data. Remote Sens. Environ. 1993, 43, 67–85. [Google Scholar] [CrossRef]
  125. Hook, S.J.; Gabell, A.R.; Green, A.A.; Kealy, P.S. A comparison of techniques for extracting emissivity information from thermal infrared data for geologic studies. Remote Sens. Environ. 1992, 42, 123–135. [Google Scholar] [CrossRef]
  126. Susskind, J.; Rosenfield, J.; Reuter, D.; Chahine, M. Remote sensing of weather and climate parameters from HIRS2/MSU on TIROS-N. J. Geophys. Res. Atmos. 1984, 89, 4677–4697. [Google Scholar] [CrossRef]
  127. Chedin, A.; Scott, N.; Wahiche, C.; Moulinier, P. The improved initialization inversion method: A high resolution physical method for temperature retrievals from satellites of the TIROS-N series. J. Appl. Meteorol. Climatol. 1985, 24, 128–143. [Google Scholar] [CrossRef] [Green Version]
  128. Price, J.C. Estimating surface temperatures from satellite thermal infrared data—A simple formulation for the atmospheric effect. Remote Sens. Environ. 1983, 13, 353–361. [Google Scholar] [CrossRef]
  129. Mushkin, A.; Balick, L.K.; Gillespie, A.R. Extending surface temperature and emissivity retrieval to the mid-infrared (3–5 μm) using the Multispectral Thermal Imager (MTI). Remote Sens. Environ. 2005, 98, 141–151. [Google Scholar] [CrossRef]
  130. Li, F.; Jackson, T.J.; Kustas, W.P.; Schmugge, T.J.; French, A.N.; Cosh, M.H.; Bindlish, R. Deriving land surface temperature from Landsat 5 and 7 during SMEX02/SMACEX. Remote Sens. Environ. 2004, 92, 521–534. [Google Scholar] [CrossRef]
  131. Dash, P.; Göttsche, F.-M.; Olesen, F.-S.; Fischer, H. Land surface temperature and emissivity estimation from passive sensor data: Theory and practice-current trends. Int. J. Remote Sens. 2002, 23, 2563–2594. [Google Scholar] [CrossRef]
  132. Gillespie, A.R.; Abbott, E.A.; Gilson, L.; Hulley, G.; Jiménez-Muñoz, J.-C.; Sobrino, J.A. Residual errors in ASTER temperature and emissivity standard products AST08 and AST05. Remote Sens. Environ. 2011, 115, 3681–3694. [Google Scholar] [CrossRef]
  133. Sobrino, J.; Sòria, G.; Prata, A. Surface temperature retrieval from Along Track Scanning Radiometer 2 data: Algorithms and validation. J. Geophys. Res. Atmos. 2004, 109. [Google Scholar] [CrossRef]
  134. McMillin, L.M. Estimation of sea surface temperatures from two infrared window measurements with different absorption. J. Geophys. Res. 1975, 80, 5113–5117. [Google Scholar] [CrossRef]
  135. Price, J.C. Land surface temperature measurements from the split window channels of the NOAA 7 Advanced Very High Resolution Radiometer. J. Geophys. Res. Atmos. 1984, 89, 7231–7237. [Google Scholar] [CrossRef]
  136. Becker, F.; Li, Z.-L. Towards a local split window method over land surfaces. Remote Sens. 1990, 11, 369–393. [Google Scholar] [CrossRef]
  137. Coll, C.; Caselles, V.; Sobrino, J.A.; Valor, E. On the atmospheric dependence of the split-window equation for land surface temperature. Remote Sens. 1994, 15, 105–122. [Google Scholar] [CrossRef]
  138. Ulivieri, C.; Castronuovo, M.; Francioni, R.; Cardillo, A. A split window algorithm for estimating land surface temperature from satellites. Adv. Space Res. 1994, 14, 59–65. [Google Scholar] [CrossRef]
  139. Wan, Z.; Dozier, J. A generalized split-window algorithm for retrieving land-surface temperature from space. IEEE Trans. Geosci. Remote Sens. 1996, 34, 892–905. [Google Scholar]
  140. Sobrino, J.; Coll, C.; Caselles, V. Atmospheric correction for land surface temperature using NOAA-11 AVHRR channels 4 and 5. Remote Sens. Environ. 1991, 38, 19–34. [Google Scholar] [CrossRef]
  141. Sobrino, J.A.; Li, Z.-L.; Stoll, M.P.; Becker, F. Improvements in the split-window technique for land surface temperature determination. IEEE Trans. Geosci. Remote Sens. 1994, 32, 243–253. [Google Scholar] [CrossRef]
  142. Sobrino, J.; Li, Z.; Stoll, M.; Becker, F. Multi-channel and multi-angle algorithms for estimating sea and land surface temperature with ATSR data. Int. J. Remote Sens. 1996, 17, 2089–2114. [Google Scholar] [CrossRef]
  143. Tang, B.; Bi, Y.; Li, Z.-L.; Xia, J. Generalized split-window algorithm for estimate of land surface temperature from Chinese geostationary FengYun meteorological satellite (FY-2C) data. Sensors 2008, 8, 933–951. [Google Scholar] [CrossRef] [Green Version]
  144. Atitar, M.; Sobrino, J.A. A split-window algorithm for estimating LST from Meteosat 9 data: Test and comparison with in situ data and MODIS LSTs. IEEE Geosci. Remote Sens. Lett. 2008, 6, 122–126. [Google Scholar] [CrossRef]
  145. Jiang, G.-M.; Zhou, W.; Liu, R. Development of split-window algorithm for land surface temperature estimation from the VIRR/FY-3A measurements. IEEE Geosci. Remote Sens. Lett. 2013, 10, 952–956. [Google Scholar] [CrossRef]
  146. Francois, C.; Ottlé, C. Atmospheric corrections in the thermal infrared: Global and water vapor dependent split-window algorithms-applications to ATSR and AVHRR data. IEEE Trans. Geosci. Remote Sens. 1996, 34, 457–470. [Google Scholar] [CrossRef]
  147. Coll, C.; Caselles, V. A split-window algorithm for land surface temperature from advanced very high resolution radiometer data: Validation and algorithm comparison. J. Geophys. Res. Atmos. 1997, 102, 16697–16713. [Google Scholar] [CrossRef]
  148. Sobrino, J.; Raissouni, N. Toward remote sensing methods for land cover dynamic monitoring: Application to Morocco. Int. J. Remote Sens. 2000, 21, 353–366. [Google Scholar] [CrossRef]
  149. Sun, D.; Pinker, R. Implementation of GOES-based land surface temperature diurnal cycle to AVHRR. Int. J. Remote Sens. 2005, 26, 3975–3984. [Google Scholar] [CrossRef]
  150. Sun, D.; Pinker, R. Retrieval of surface temperature from the MSG-SEVIRI observations: Part I. Methodology. Int. J. Remote Sens. 2007, 28, 5255–5272. [Google Scholar] [CrossRef]
  151. Sun, D.; Pinker, R.T. Estimation of land surface temperature from a Geostationary Operational Environmental Satellite (GOES-8). J. Geophys. Res. Atmos. 2003, 108. [Google Scholar] [CrossRef]
  152. Chedin, A.; Scott, N.; Berroir, A. A single-channel, double-viewing angle method for sea surface temperature determination from coincident Meteosat and TIROS-N radiometric measurements. J. Appl. Meteorol. Climatol. 1982, 21, 613–618. [Google Scholar] [CrossRef] [Green Version]
  153. Prata, A.J. Land surface temperatures derived from the advanced very high resolution radiometer and the along-track scanning radiometer: 1. Theory. J. Geophys. Res. Atmos. 1993, 98, 16689–16702. [Google Scholar] [CrossRef]
  154. Sobrino, J.A.; Jiménez-Muñoz, J.C. Land surface temperature retrieval from thermal infrared data: An assessment in the context of the Surface Processes and Ecosystem Changes Through Response Analysis (SPECTRA) mission. J. Geophys. Res. Atmos. 2005, 110. [Google Scholar] [CrossRef] [Green Version]
  155. Sobrino, J.A.; Cuenca, J. Angular variation of thermal infrared emissivity for some natural surfaces from experimental measurements. Appl. Opt. 1999, 38, 3931–3936. [Google Scholar] [CrossRef]
  156. Watson, K. Spectral ratio method for measuring emissivity. Remote Sens. Environ. 1992, 42, 113–116. [Google Scholar] [CrossRef]
  157. Caselles, V.; Coll, C.; Valor, E. Land surface emissivity and temperature determination in the whole HAPEX-Sahel area from AVHRR data. Int. J. Remote Sens. 1997, 18, 1009–1027. [Google Scholar] [CrossRef]
  158. Wan, Z. MODIS land-surface temperature algorithm theoretical basis document (LST ATBD). Inst. Comput. Earth Syst. Sci. St. Barbar. 1999, 75, 18. [Google Scholar]
  159. Wan, Z.; Li, Z.-L. A physics-based algorithm for retrieving land-surface emissivity and temperature from EOS/MODIS data. IEEE Trans. Geosci. Remote Sens. 1997, 35, 980–996. [Google Scholar]
  160. Choudhury, B.J.; Menenti, M. Parameteraization of Land Surface Evaporation by Means of Location Dependent Potential Evaporation and Surface Temperature Range; Department for Environment, Food and Rural Affairs (Defra): London, UK, 1993; Volume 212, pp. 561–568. [Google Scholar]
  161. Jackson, R.D.; Idso, S.; Reginato, R.; Pinter Jr, P. Canopy temperature as a crop water stress indicator. Water Resour. Res. 1981, 17, 1133–1138. [Google Scholar] [CrossRef]
  162. Idso, S.; Jackson, R.; Pinter Jr, P.; Reginato, R.; Hatfield, J. Normalizing the stress-degree-day parameter for environmental variability. Agric. Meteorol. 1981, 24, 45–55. [Google Scholar] [CrossRef]
  163. Van den Hurk, B. Energy balance based surface flux estimation from satellite data, and its application for surface moisture assimilation. Meteorol. Atmos. Phys. 2001, 76, 43–52. [Google Scholar] [CrossRef]
  164. Kilic, A.; Allen, R.; Trezza, R.; Ratcliffe, I.; Kamble, B.; Robison, C.; Ozturk, D. Sensitivity of evapotranspiration retrievals from the METRIC processing algorithm to improved radiometric resolution of Landsat 8 thermal data and to calibration bias in Landsat 7 and 8 surface temperature. Remote Sens. Environ. 2016, 185, 198–209. [Google Scholar] [CrossRef] [Green Version]
  165. Wang, X.-G.; Wang, W.; Huang, D.; Yong, B.; Chen, X. Modifying SEBAL model based on the trapezoidal relationship between land surface temperature and vegetation index for actual evapotranspiration estimation. Remote Sens. 2014, 6, 5909–5937. [Google Scholar] [CrossRef] [Green Version]
  166. Norman, J.; Anderson, M.; Kustas, W. Are Single-Source, Remote-Sensing Surface-Flux Models Too Simple? In Proceedings of the AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2006; pp. 170–177. [Google Scholar]
  167. Brutsaert, W. Aspects of bulk atmospheric boundary layer similarity under free-convective conditions. Rev. Geophys. 1999, 37, 439–451. [Google Scholar] [CrossRef]
  168. Byun, K.; Liaqat, U.W.; Choi, M. Dual-model approaches for evapotranspiration analyses over homo-and heterogeneous land surface conditions. Agric. For. Meteorol. 2014, 197, 169–187. [Google Scholar] [CrossRef]
  169. Webster, E.; Ramp, D.; Kingsford, R.T. Incorporating an iterative energy restraint for the Surface Energy Balance System (SEBS). Remote Sens. Environ. 2017, 198, 267–285. [Google Scholar] [CrossRef] [Green Version]
  170. McShane, R.R.; Driscoll, K.P.; Sando, R. A review of surface energy balance models for estimating actual evapotranspiration with remote sensing at high spatiotemporal resolution over large extents. In Scientific Investigations Report 2017–5087; US Geological Survey: Reston, VA, USA, 2017; p. 19. [Google Scholar]
  171. Sobrino, J.; Gómez, M.; Jiménez-Muñoz, J.; Olioso, A. Application of a simple algorithm to estimate daily evapotranspiration from NOAA–AVHRR images for the Iberian Peninsula. Remote Sens. Environ. 2007, 110, 139–148. [Google Scholar] [CrossRef]
  172. Rocha, N.S.d.; Käfer, P.S.; Skokovic, D.; Veeck, G.; Diaz, L.R.; Kaiser, E.A.; Carvalho, C.M.; Cruz, R.C.; Sobrino, J.A.; Roberti, D.R. The Influence of Land Surface Temperature in Evapotranspiration Estimated by the S-SEBI Model. Atmosphere 2020, 11, 1059. [Google Scholar] [CrossRef]
  173. Singh, R.K.; Irmak, A.; Irmak, S.; Martin, D.L. Application of SEBAL model for mapping evapotranspiration and estimating surface energy fluxes in south-central Nebraska. J. Irrig. Drain. Eng. 2008, 134, 273–285. [Google Scholar] [CrossRef]
  174. Prakash Mohan, M.; Rajitha, K.; Varma, M.R. Integration of soil moisture as an auxiliary parameter for the anchor pixel selection process in SEBAL using Landsat 8 and Sentinel-1A images. Int. J. Remote Sens. 2020, 41, 1214–1231. [Google Scholar] [CrossRef]
  175. Allen, R.; Irmak, A.; Trezza, R.; Hendrickx, J.M.; Bastiaanssen, W.; Kjaersgaard, J. Satellite-based ET estimation in agriculture using SEBAL and METRIC. Hydrol. Process. 2011, 25, 4011–4027. [Google Scholar] [CrossRef]
  176. 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]
  177. 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]
  178. Allen, R.; Tasumi, M.; Trezza, R.; Kjaersgaard, J. Mapping evapotranspiration at high resolution, applications manual for landsat satellite imagery. Version 2010, 2, 248. [Google Scholar]
  179. Senay, G.B.; Budde, M.E.; Verdin, J.P. Enhancing the Simplified Surface Energy Balance (SSEB) approach for estimating landscape ET: Validation with the METRIC model. Agric. Water Manag. 2011, 98, 606–618. [Google Scholar] [CrossRef]
  180. Colaizzi, P.D.; Kustas, W.P.; Anderson, M.C.; Agam, N.; Tolk, J.A.; Evett, S.R.; Howell, T.A.; Gowda, P.H.; O’Shaughnessy, S.A. Two-source energy balance model estimates of evapotranspiration using component and composite surface temperatures. Adv. Water Resour. 2012, 50, 134–151. [Google Scholar] [CrossRef] [Green Version]
  181. 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]
  182. Komatsu, T.S. Toward a robust phenomenological expression of evaporation efficiency for unsaturated soil surfaces. J. Appl. Meteorol. 2003, 42, 1330–1334. [Google Scholar] [CrossRef]
  183. Anderson, M.; Norman, J.; Diak, G.; Kustas, W.; Mecikalski, J. 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]
  184. McNaughton, K.; Spriggs, T. A mixed-layer model for regional evaporation. Bound.-Layer Meteorol. 1986, 34, 243–262. [Google Scholar] [CrossRef]
  185. Norman, J.; Anderson, M.; Kustas, W.; French, A.; Mecikalski, J.; Torn, R.; Diak, G.; Schmugge, T.; Tanner, B. Remote sensing of surface energy fluxes at 101-m pixel resolutions. Water Resour. Res. 2003, 39. [Google Scholar] [CrossRef] [Green Version]
  186. Carlson, T. An overview of the" triangle method" for estimating surface evapotranspiration and soil moisture from satellite imagery. Sensors 2007, 7, 1612–1629. [Google Scholar] [CrossRef] [Green Version]
  187. 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]
  188. 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]
  189. Owen, T.; Carlson, T.; Gillies, R. An assessment of satellite remotely-sensed land cover parameters in quantitatively describing the climatic effect of urbanization. Int. J. Remote Sens. 1998, 19, 1663–1681. [Google Scholar] [CrossRef]
  190. PRiCE, J.C. Using spatial context in satellite data to infer regional scale evapotranspiration. IEEE Trans. Geosci. Remote Sens. 1990, 28, 940–948. [Google Scholar] [CrossRef] [Green Version]
  191. Carlson, T.N.; Ripley, D.A. On the relation between NDVI, fractional vegetation cover, and leaf area index. Remote Sens. Environ. 1997, 62, 241–252. [Google Scholar] [CrossRef]
  192. Gillies, R.; Kustas, W.; Humes, K. A verification of the’triangle’method for obtaining surface soil water content and energy fluxes from remote measurements of the Normalized Difference Vegetation Index (NDVI) and surface e. Int. J. Remote Sens. 1997, 18, 3145–3166. [Google Scholar] [CrossRef]
  193. Jiang, L.; Islam, S.; Guo, W.; Jutla, A.S.; Senarath, S.U.; Ramsay, B.H.; Eltahir, E. A satellite-based daily actual evapotranspiration estimation algorithm over South Florida. Glob. Planet. Chang. 2009, 67, 62–77. [Google Scholar] [CrossRef]
  194. Nemani, R.; Pierce, L.; Running, S.; Goward, S. Developing satellite-derived estimates of surface moisture status. J. Appl. Meteorol. Climatol. 1993, 32, 548–557. [Google Scholar] [CrossRef] [Green Version]
  195. Yang, Y.; Shang, S. A hybrid dual-source scheme and trapezoid framework–based evapotranspiration model (HTEM) using satellite images: Algorithm and model test. J. Geophys. Res. Atmos. 2013, 118, 2284–2300. [Google Scholar] [CrossRef]
  196. Petropoulos, G.; Carlson, T.; Wooster, M.; Islam, S. A review of Ts/VI remote sensing based methods for the retrieval of land surface energy fluxes and soil surface moisture. Prog. Phys. Geogr. 2009, 33, 224–250. [Google Scholar] [CrossRef] [Green Version]
  197. Zhang, R.; Tian, J.; Su, H.; Sun, X.; Chen, S.; Xia, J. Two improvements of an operational two-layer model for terrestrial surface heat flux retrieval. Sensors 2008, 8, 6165–6187. [Google Scholar] [CrossRef]
  198. 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]
  199. Tian, J.; Su, H.; Sun, X.; Chen, S.; He, H.; Zhao, L. Impact of the spatial domain size on the performance of the Ts-VI triangle method in terrestrial evapotranspiration estimation. Remote Sens. 2013, 5, 1998–2013. [Google Scholar] [CrossRef] [Green Version]
  200. 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]
  201. Holmes, T.R.; Hain, C.R.; Crow, W.T.; Anderson, M.C.; Kustas, W.P. Microwave implementation of two-source energy balance approach for estimating evapotranspiration. Hydrol. Earth Syst. Sci. 2018, 22, 1351–1369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Brutsaert, W.; Sugita, M. Sensible heat transfer parameterization for surfaces with anisothermal dense vegetation. J. Atmos. Sci. 1996, 53, 209–216. [Google Scholar] [CrossRef] [Green Version]
  203. Hall, F.G.; Huemmrich, K.F.; Goetz, S.J.; Sellers, P.J.; Nickeson, J.E. Satellite remote sensing of surface energy balance: Success, failures, and unresolved issues in FIFE. J. Geophys. Res. Atmos. 1992, 97, 19061–19089. [Google Scholar] [CrossRef]
  204. Vining, R.; Blad, B. Estimation of sensible heat flux from remotely sensed canopy temperatures. J. Geophys. Res. Atmos. 1992, 97, 18951–18954. [Google Scholar] [CrossRef]
  205. Sun, J.; Mahrt, L. Relationship of surface heat flux to microscale temperature variations: Application to BOREAS. Bound.-Layer Meteorol. 1995, 76, 291–301. [Google Scholar] [CrossRef] [Green Version]
  206. Sun, J.; Massman, W.; Grantz, D.A. Aerodynamic variables in the bulk formulation of turbulent fluxes. Bound.-Layer Meteorol. 1999, 91, 109–125. [Google Scholar] [CrossRef]
  207. Qualls, R.J.; Brutsaert, W. Effect of vegetation density on the parameterization of scalar roughness to estimate spatially distributed sensible heat fluxes. Water Resour. Res. 1996, 32, 645–652. [Google Scholar] [CrossRef]
  208. Kustas, W.P. Estimates of evapotranspiration with a one-and two-layer model of heat transfer over partial canopy cover. J. Appl. Meteorol. Climatol. 1990, 29, 704–715. [Google Scholar] [CrossRef]
  209. Kustas, W.P.; Choudhury, B.J.; Moran, M.S.; Reginato, R.J.; Jackson, R.D.; Gay, L.W.; Weaver, H.L. Determination of sensible heat flux over sparse canopy using thermal infrared data. Agric. For. Meteorol. 1989, 44, 197–216. [Google Scholar] [CrossRef]
  210. Stewart, J.B.; Kustas, W.P.; Humes, K.S.; Nichols, W.D.; Moran, M.S.; de Bruin, H.A. Sensible heat flux-radiometric surface temperature relationship for eight semiarid areas. J. Appl. Meteorol. Climatol. 1994, 33, 1110–1117. [Google Scholar] [CrossRef] [Green Version]
  211. Lhomme, J.-P.; Troufleau, D.; Monteny, B.; Chehbouni, A.; Bauduin, S. Sensible heat flux and radiometric surface temperature over sparse Sahelian vegetation II. A model for the kB− 1 parameter. J. Hydrol. 1997, 188, 839–854. [Google Scholar] [CrossRef]
  212. Chehbouni, A.; Seen, D.L.; Njoku, E.; Monteny, B. Examination of the difference between radiative and aerodynamic surface temperatures over sparsely vegetated surfaces. Remote Sens. Environ. 1996, 58, 177–186. [Google Scholar] [CrossRef]
  213. Troufleau, D.; Lhomme, J.-P.; Monteny, B.; Vidal, A. Sensible heat flux and radiometric surface temperature over sparse Sahelian vegetation. I. An experimental analysis of the kB− 1 parameter. J. Hydrol. 1997, 188, 815–838. [Google Scholar] [CrossRef]
  214. Zhuang, Q.; Wu, B.; Yan, N.; Zhu, W.; Xing, Q. A method for sensible heat flux model parameterization based on radiometric surface temperature and environmental factors without involving the parameter KB− 1. Int. J. Appl. Earth Obs. Geoinf. 2016, 47, 50–59. [Google Scholar] [CrossRef]
  215. 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]
  216. Guzinski, R.; Anderson, M.; Kustas, W.; Nieto, H.; Sandholt, I. Using a thermal-based two source energy balance model with time-differencing to estimate surface energy fluxes with day–night MODIS observations. Hydrol. Earth Syst. Sci. 2013, 17, 2809–2825. [Google Scholar] [CrossRef] [Green Version]
  217. Leuning, R.; Denmead, O.; Lang, A.; Ohtaki, E. Effects of heat and water vapor transport on eddy covariance measurement of CO 2 fluxes. Bound.-Layer Meteorol. 1982, 23, 209–222. [Google Scholar] [CrossRef]
  218. Koitzsch, R.; Dzingel, M.; Foken, T.; Mücket, G. Probleme der experimentellen Erfassung des Energieaustausches über Winterweizen. Z. Für Meteorol. 1988, 38, 150–155. [Google Scholar]
  219. Desjardins, R. Carbon dioxide budget of maize. Agric. For. Meteorol. 1985, 36, 29–41. [Google Scholar] [CrossRef]
  220. Tsvang, L.; Fedorov, M.; Kader, B.; Zubkovskii, S.; Foken, T.; Richter, S.; Zeleny, Y. Turbulent exchange over a surface with chessboard-type inhomogeneities. Bound.-Layer Meteorol. 1991, 55, 141–160. [Google Scholar] [CrossRef]
  221. Panin, G.; Tetzlaff, G.; Raabe, A. Inhomogeneity of the Land Surface and Problems in theParameterization of Surface Fluxes in Natural Conditions. Theor. Appl. Climatol. 1998, 60, 163–178. [Google Scholar] [CrossRef]
  222. Kanemasu, E.; Verma, S.; Smith, E.; Fritschen, L.; Wesely, M.; Field, R.; Kustas, W.; Weaver, H.; Stewart, J.; Gurney, R. Surface flux measurements in FIFE: An overview. J. Geophys. Res. Atmos. 1992, 97, 18547–18555. [Google Scholar] [CrossRef]
  223. Van der Molen, M.; Gash, J.; Elbers, J. Sonic anemometer (co) sine response and flux measurement: II. The effect of introducing an angle of attack dependent calibration. Agric. For. Meteorol. 2004, 122, 95–109. [Google Scholar] [CrossRef]
  224. Nakai, T.; Van Der Molen, M.; Gash, J.; Kodama, Y. Correction of sonic anemometer angle of attack errors. Agric. For. Meteorol. 2006, 136, 19–30. [Google Scholar] [CrossRef] [Green Version]
  225. Halldin, S.; Lindroth, A. Errors in net radiometry: Comparison and evaluation of six radiometer designs. J. Atmos. Ocean. Technol. 1992, 9, 762–783. [Google Scholar] [CrossRef] [Green Version]
  226. Peters-Lidard, C.; Blackburn, E.; Liang, X.; Wood, E.F. The effect of soil thermal conductivity parameterization on surface energy fluxes and temperatures. J. Atmos. Sci. 1998, 55, 1209–1224. [Google Scholar] [CrossRef]
  227. Culf, A.D.; Foken, T.; Gash, J.H. The energy balance closure problem. In Vegetation, Water, Humans and the Climate; Springer: Berlin/Heidelberg, Germany, 2004; pp. 159–166. [Google Scholar]
  228. Rannik, Ü.; Vesala, T. Autoregressive filtering versus linear detrending in estimation of fluxes by the eddy covariance method. Bound.-Layer Meteorol. 1999, 91, 259–280. [Google Scholar] [CrossRef]
  229. Finnigan, J.J.; Clement, R.; Malhi, Y.; Leuning, R.; Cleugh, H. A re-evaluation of long-term flux measurement techniques part I: Averaging and coordinate rotation. Bound. Layer Meteorol. 2003, 107, 1–48. [Google Scholar] [CrossRef]
  230. Wang, W. The influence of topography on single-tower-based carbon flux measurements under unstable conditions: A modeling perspective. Theor. Appl. Climatol. 2010, 99, 125–138. [Google Scholar] [CrossRef]
  231. Foken, T.; Leuning, R.; Oncley, S.R.; Mauder, M.; Aubinet, M. Corrections and data quality control. In Eddy Covariance; Springer: Berlin/Heidelberg, Germany, 2012; pp. 85–131. [Google Scholar]
  232. Segal, M.; Arritt, R. Nonclassical mesoscale circulations caused by surface sensible heat-flux gradients. Bull. Am. Meteorol. Soc. 1992, 73, 1593–1604. [Google Scholar] [CrossRef] [Green Version]
  233. Etling, D.; Brown, R.A. Roll vortices in the planetary boundary layer: A review. Bound.-Layer Meteorol. 1993, 65, 215–248. [Google Scholar] [CrossRef]
  234. Haverd, V.; Cuntz, M.; Leuning, R.; Keith, H. Air and biomass heat storage fluxes in a forest canopy: Calculation within a soil vegetation atmosphere transfer model. Agric. For. Meteorol. 2007, 147, 125–139. [Google Scholar] [CrossRef]
  235. Moderow, U.; Aubinet, M.; Feigenwinter, C.; Kolle, O.; Lindroth, A.; Mölder, M.; Montagnani, L.; Rebmann, C.; Bernhofer, C. Available energy and energy balance closure at four coniferous forest sites across Europe. Theor. Appl. Climatol. 2009, 98, 397–412. [Google Scholar] [CrossRef] [Green Version]
  236. Lindroth, A.; Mölder, M.; Lagergren, F. Heat storage in forest biomass improves energy balance closure. Biogeosciences 2010, 7, 301–313. [Google Scholar] [CrossRef] [Green Version]
  237. Kracher, D.; Mengelkamp, H.-T.; Foken, T. The residual of the energy balance closure and its influence on the results of three SVAT models. Meteorol. Z. 2009, 18, 647. [Google Scholar] [CrossRef]
  238. Yan, H.; Zhang, C.; Hiroki, O. Parameterization of canopy resistance for modeling the energy partitioning of a paddy rice field. Paddy Water Environ. 2018, 16, 109–123. [Google Scholar] [CrossRef] [Green Version]
  239. Ershadi, A.; McCabe, M.; Evans, J.P.; Chaney, N.W.; Wood, E.F. Multi-site evaluation of terrestrial evaporation models using FLUXNET data. Agric. For. Meteorol. 2014, 187, 46–61. [Google Scholar] [CrossRef]
  240. Swenson, S.; Lawrence, D. Assessing a dry surface layer-based soil resistance parameterization for the Community Land Model using GRACE and FLUXNET-MTE data. J. Geophys. Res. Atmos. 2014, 119, 10–299. [Google Scholar] [CrossRef]
  241. Wang, K.; Dickinson, R.E. A review of global terrestrial evapotranspiration: Observation, modeling, climatology, and climatic variability. Rev. Geophys. 2012, 50. [Google Scholar] [CrossRef]
  242. Shuttleworth, W.J.; Gurney, R.J. The theoretical relationship between foliage temperature and canopy resistance in sparse crops. Q. J. R. Meteorol. Soc. 1990, 116, 497–519. [Google Scholar] [CrossRef]
  243. Deng, Z.; Guan, H.; Hutson, J.; Forster, M.A.; Wang, Y.; Simmons, C.T. A vegetation-focused soil-plant-atmospheric continuum model to study hydrodynamic soil-plant water relations. Water Resour. Res. 2017, 53, 4965–4983. [Google Scholar] [CrossRef]
  244. Brutsaert, W.; Stricker, H. An advection-aridity approach to estimate actual regional evapotranspiration. Water Resour. Res. 1979, 15, 443–450. [Google Scholar] [CrossRef]
  245. Yang, K.; Tamai, N.; Koike, T. Analytical solution of surface layer similarity equations. J. Appl. Meteorol. 2001, 40, 1647–1653. [Google Scholar] [CrossRef]
  246. Verma, S.; Rosenberg, N.; Blad, B.; Baradas, M. Resistance-energy Balance Method for Predicting Evapotranspiration: Determination of Boundary Layer Resistance and Evaluation of Error Effects 1. Agron. J. 1976, 68, 776–782. [Google Scholar] [CrossRef]
  247. Xie, X. An improved energy balance-aerodynamic resistance model used estimation of evapotranspiration on the wheat field. Acta Meteorol. Sin. 1988, 1, 102–106. [Google Scholar]
  248. Hatfield, J.; Perrier, A.; Jackson, R. Estimation of evapotranspiration at one time-of-day using remotely sensed surface temperatures. In Developments in Agricultural and Managed Forest Ecology; Elsevier: Amsterdam, The Netherlands, 1983; Volume 12, pp. 341–350. [Google Scholar]
  249. Viney, N.R. An empirical expression for aerodynamic resistance in the unstable boundary layer. Bound.-Layer Meteorol. 1991, 56, 381–393. [Google Scholar] [CrossRef]
  250. Jackson, R.D.; Reginato, R.; Idso, S. Wheat canopy temperature: A practical tool for evaluating water requirements. Water Resour. Res. 1977, 13, 651–656. [Google Scholar] [CrossRef]
  251. Richardson, L.F. The supply of energy from and to atmospheric eddies. Proc. R. Soc. Lond. Ser. A Contain. Pap. A Math. Phys. Character 1920, 97, 354–373. [Google Scholar]
  252. Su, Z.; Schmugge, T.; Kustas, W.; Massman, W. An evaluation of two models for estimation of the roughness height for heat transfer between the land surface and the atmosphere. J. Appl. Meteorol. 2001, 40, 1933–1951. [Google Scholar] [CrossRef] [Green Version]
  253. 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]
  254. Moran, M.S.; Jackson, R.D.; Raymond, L.H.; Gay, L.W.; Slater, P.N. Mapping surface energy balance components by combining Landsat Thematic Mapper and ground-based meteorological data. Remote Sens. Environ. 1989, 30, 77–87. [Google Scholar] [CrossRef]
  255. Irmak, S.; Mutiibwa, D. On the dynamics of canopy resistance: Generalized linear estimation and relationships with primary micrometeorological variables. Water Resour. Res. 2010, 46. [Google Scholar] [CrossRef] [Green Version]
  256. Szeicz, G.; Long, I. Surface resistance of crop canopies. Water Resour. Res. 1969, 5, 622–633. [Google Scholar] [CrossRef]
  257. Jarvis, P. Stomatal conductance, gaseous exchange and transpiration. In Proceedings of the Symposium-British Ecological Society; British Ecological Society: Oxford, UK, 1981. [Google Scholar]
  258. Lhomme, J.-P. The concept of canopy resistance: Historical survey and comparison of different approaches. Agric. For. Meteorol. 1991, 54, 227–240. [Google Scholar] [CrossRef]
  259. Sperry, J.S. Hydraulic constraints on plant gas exchange. Agric. For. Meteorol. 2000, 104, 13–23. [Google Scholar] [CrossRef]
  260. Ewers, B.; Oren, R.; Kim, H.S.; Bohrer, G.; Lai, C.T. Effects of hydraulic architecture and spatial variation in light on mean stomatal conductance of tree branches and crowns. Plant Cell Environ. 2007, 30, 483–496. [Google Scholar] [CrossRef]
  261. Katerji, N.; Perrier, A. A Model of Actual Evapotranspiration (ETR) for a Field of Lucerne: The Role of a Crop Coefficient. 1983. Available online: https://agris.fao.org/agris-search/search.do?recordID=FR2014005468 (accessed on 6 February 2022).
  262. He, B.; Oue, H.; Oki, T. Estimation of hourly evapotranspiration in arid regions by a simple parameterization of canopy resistance. In Proceedings of the Annual Meeting of the Society of Agricultural Meteorology of Japan Abstracts of International Symposium on Agricultural Meteorology, Hokkaido, Japan, 6 February 2008; p. 94. [Google Scholar]
  263. Todorovic, M. Single-layer evapotranspiration model with variable canopy resistance. J. Irrig. Drain. Eng. 1999, 125, 235–245. [Google Scholar] [CrossRef]
  264. Pauwels, V.R.; Samson, R. Comparison of different methods to measure and model actual evapotranspiration rates for a wet sloping grassland. Agric. Water Manag. 2006, 82, 1–24. [Google Scholar] [CrossRef]
  265. Taheri, M.; Gholizadeh, M.; Nasseri, M.; Zahraie, B.; Poorsepahy-Samian, H.; Espanmanesh, V. Performance evaluation of various evapotranspiration modeling scenarios based on METRIC method and climatic indexes. Environ. Monit. Assess. 2021, 193, 1–18. [Google Scholar] [CrossRef]
  266. Kumar, R.; Aggarwal, R.; Sharma, J. Comparison of regression and artificial neural network models for estimation of global solar radiations. Renew. Sustain. Energy Rev. 2015, 52, 1294–1299. [Google Scholar] [CrossRef]
  267. Shamshirband, S.; Mohammadi, K.; Khorasanizadeh, H.; Yee, L.; Lee, M.; Petković, D.; Zalnezhad, E. Estimating the diffuse solar radiation using a coupled support vector machine–wavelet transform model. Renew. Sustain. Energy Rev. 2016, 56, 428–435. [Google Scholar] [CrossRef]
  268. Mohammadi, K.; Shamshirband, S.; Danesh, A.S.; Abdullah, M.S.; Zamani, M. RETRACTED ARTICLE: Temperature-based estimation of global solar radiation using soft computing methodologies. Theor. Appl. Climatol. 2016, 125, 101–112. [Google Scholar] [CrossRef]
  269. Bastiaanssen, W.G. SEBAL-based sensible and latent heat fluxes in the irrigated Gediz Basin, Turkey. J. Hydrol. 2000, 229, 87–100. [Google Scholar] [CrossRef]
  270. Alonso-Montesinos, J.; Batlles, F.; Bosch, J. Beam, diffuse and global solar irradiance estimation with satellite imagery. Energy Convers. Manag. 2015, 105, 1205–1212. [Google Scholar] [CrossRef]
  271. Niemelä, S.; Räisänen, P.; Savijärvi, H. Comparison of surface radiative flux parameterizations: Part I: Longwave radiation. Atmos. Res. 2001, 58, 1–18. [Google Scholar] [CrossRef]
  272. Yang, K.; Koike, T.; Ye, B. Improving estimation of hourly, daily, and monthly solar radiation by importing global data sets. Agric. For. Meteorol. 2006, 137, 43–55. [Google Scholar] [CrossRef]
  273. Besharat, F.; Dehghan, A.A.; Faghih, A.R. Empirical models for estimating global solar radiation: A review and case study. Renew. Sustain. Energy Rev. 2013, 21, 798–821. [Google Scholar] [CrossRef]
  274. Al-Mostafa, Z.; Maghrabi, A.; Al-Shehri, S. Sunshine-based global radiation models: A review and case study. Energy Convers. Manag. 2014, 84, 209–216. [Google Scholar] [CrossRef]
  275. Meza, F.; Varas, E. Estimation of mean monthly solar global radiation as a function of temperature. Agric. For. Meteorol. 2000, 100, 231–241. [Google Scholar] [CrossRef]
  276. Dos Santos, C.M.; De Souza, J.L.; Junior, R.A.F.; Tiba, C.; de Melo, R.O.; Lyra, G.B.; Teodoro, I.; Lyra, G.B.; Lemes, M.A.M. On modeling global solar irradiation using air temperature for Alagoas State, Northeastern Brazil. Energy 2014, 71, 388–398. [Google Scholar] [CrossRef]
  277. Hassan, G.E.; Youssef, M.E.; Mohamed, Z.E.; Ali, M.A.; Hanafy, A.A. New temperature-based models for predicting global solar radiation. Appl. Energy 2016, 179, 437–450. [Google Scholar] [CrossRef]
  278. Nielsen, L.; Prahm, L.; Berkowicz, R.; Conradsen, K. Net incoming radiation estimated from hourly global radiation and/or cloud observations. J. Climatol. 1981, 1, 255–272. [Google Scholar] [CrossRef]
  279. Supit, I.; Van Kappel, R. A simple method to estimate global radiation. Sol. Energy 1998, 63, 147–160. [Google Scholar] [CrossRef]
  280. Wu, G.; Liu, Y.; Wang, T. Methods and strategy for modeling daily global solar radiation with measured meteorological data–A case study in Nanchang station, China. Energy Convers. Manag. 2007, 48, 2447–2452. [Google Scholar] [CrossRef]
  281. Zhao, N.; Zeng, X.; Han, S. Solar radiation estimation using sunshine hour and air pollution index in China. Energy Convers. Manag. 2013, 76, 846–851. [Google Scholar] [CrossRef]
  282. Choi, M.; Jacobs, J.M.; Kustas, W.P. Assessment of clear and cloudy sky parameterizations for daily downwelling longwave radiation over different land surfaces in Florida, USA. Geophys. Res. Lett. 2008, 35. [Google Scholar] [CrossRef] [Green Version]
  283. Wang, K.; Liang, S. Global atmospheric downward longwave radiation over land surface under all-sky conditions from 1973 to 2008. J. Geophys. Res. Atmos. 2009, 114. [Google Scholar] [CrossRef]
  284. Ruhoff, A.L.; Paz, A.R.; Collischonn, W.; Aragao, L.E.; Rocha, H.R.; Malhi, Y.S. A MODIS-based energy balance to estimate evapotranspiration for clear-sky days in Brazilian tropical savannas. Remote Sens. 2012, 4, 703–725. [Google Scholar] [CrossRef] [Green Version]
  285. Sun, Z.; Wang, Q.; Matsushita, B.; Fukushima, T.; Ouyang, Z.; Watanabe, M. Development of a simple remote sensing evapotranspiration model (Sim-ReSET): Algorithm and model test. J. Hydrol. 2009, 376, 476–485. [Google Scholar] [CrossRef] [Green Version]
  286. Sridhar, V.; Elliott, R.L. On the development of a simple downwelling longwave radiation scheme. Agric. For. Meteorol. 2002, 112, 237–243. [Google Scholar] [CrossRef]
  287. Kjaersgaard, J.; Plauborg, F.; Hansen, S. Comparison of models for calculating daytime long-wave irradiance using long term data set. Agric. For. Meteorol. 2007, 143, 49–63. [Google Scholar] [CrossRef]
  288. Key, J.R.; Schweiger, A.J. Tools for atmospheric radiative transfer: Streamer and FluxNet. Comput. Geosci. 1998, 24, 443–451. [Google Scholar] [CrossRef] [Green Version]
  289. Bennett, T.J., Jr. A coupled atmosphere-sea ice model study of the role of sea ice in climatic predictability. J. Atmos. Sci. 1982, 39, 1456–1465. [Google Scholar] [CrossRef]
  290. Paltridge, G.W.; CMR, P. Radiative Processes in Meteorology and Climatology. 1976. Available online: https://agris.fao.org/agris-search/search.do?recordID=US201300532740 (accessed on 6 February 2022).
  291. Lumb, F. The influence of cloud on hourly amounts of total solar radiation at the sea surface. Q. J. R. Meteorol. Soc. 1964, 90, 43–56. [Google Scholar] [CrossRef]
  292. Moritz, R.; Barry, R.; Jacobs, J. A model for estimating global solar radiation. Energy Budg. Stud. Relat. Fast-Ice Break. Process. Davis Strait Climatol. Overv. INSTAAR Occas. Pap. 1978, 26, 121–142. [Google Scholar]
  293. Zillmann, J. A study of some aspects of the radiation and the heat budgets of the Southern Hemisphere oceans. Meteor. Stud. 1972, 26, 562. [Google Scholar]
  294. Sun, Z.; Gebremichael, M.; Wang, Q.; Wang, J.; Sammis, T.W.; Nickless, A. Evaluation of clear-sky incoming radiation estimating equations typically used in remote sensing evapotranspiration algorithms. Remote Sens. 2013, 5, 4735–4752. [Google Scholar] [CrossRef] [Green Version]
  295. Shine, K. Parametrization of the shortwave flux over high albedo surfaces as a function of cloud thickness and surface albedo. Q. J. R. Meteorol. Soc. 1984, 110, 747–764. [Google Scholar] [CrossRef]
  296. Tasumi, M.; Allen, R.; Bastiaanssen, W.; Morse, A.; Tasumi, M.; Allen, R.; Kramber, W. The Theoretical Basis of SEBAL; Raytheon Systems Company, Earth Observation System Data and Information System Project, Idaho Department of Water Resources and University of Idaho: Arlington, VA, USA, 2000; Volume 1177. [Google Scholar]
  297. Allen, R.G. Assessing integrity of weather data for reference evapotranspiration estimation. J. Irrig. Drain. Eng. 1996, 122, 97–106. [Google Scholar] [CrossRef]
  298. Allen, R.G.; Trezza, R.; Tasumi, M. Analytical integrated functions for daily solar radiation on slopes. Agric. For. Meteorol. 2006, 139, 55–73. [Google Scholar] [CrossRef]
  299. Kondo, J. Atmospheric Science near the Ground Surface; University of Tokyo: Tokyo, Japan, 2000. [Google Scholar]
  300. Berliand, T. Method of climatological estimation of global radiation. Meteorol. Gidrol 1960, 6, 9–12. [Google Scholar]
  301. Laevastu, T. Factors affecting the temperature of the surface layer of the sea. Comment. Phys. Math. 1960, 25, 1–136. [Google Scholar]
  302. Hargreaves, G.H.; Samani, Z.A. Estimating potential evapotranspiration. J. Irrig. Drain. Div. 1982, 108, 225–230. [Google Scholar] [CrossRef]
  303. Allen, R. Evaluation of Procedures for Estimating Mean Monthly Solar Radiation from Air Temperature. 1995. Available online: https://agris.fao.org/agris-search/search.do?recordID=XF2016062112 (accessed on 6 February 2022).
  304. Samani, Z. Estimating solar radiation and evapotranspiration using minimum climatological data. J. Irrig. Drain. Eng. 2000, 126, 265–267. [Google Scholar] [CrossRef]
  305. Annandale, J.; Jovanovic, N.; Benade, N.; Allen, R. Software for missing data error analysis of Penman-Monteith reference evapotranspiration. Irrig. Sci. 2002, 21, 57–67. [Google Scholar]
  306. Chen, R.; Ersi, K.; Yang, J.; Lu, S.; Zhao, W. Validation of five global radiation models with measured daily data in China. Energy Convers. Manag. 2004, 45, 1759–1769. [Google Scholar] [CrossRef]
  307. Prescott, J. Evaporation from a water surface in relation to solar radiation. Trans. Roy. Soc. S. Aust. 1940, 46, 114–118. [Google Scholar]
  308. Bahel, V.; Bakhsh, H.; Srinivasan, R. A correlation for estimation of global solar radiation. Energy 1987, 12, 131–135. [Google Scholar] [CrossRef]
  309. Kirmani, S.; Jamil, M.; Rizwan, M. Empirical correlation of estimating global solar radiation using meteorological parameters. Int. J. Sustain. Energy 2015, 34, 327–339. [Google Scholar] [CrossRef]
  310. Jahani, B.; Dinpashoh, Y.; Nafchi, A.R. Evaluation and development of empirical models for estimating daily solar radiation. Renew. Sustain. Energy Rev. 2017, 73, 878–891. [Google Scholar] [CrossRef]
  311. Ångström, A. A Study of the Radiation of the Atmosphere: Based upon Observations of the Nocturnal Radiation during Expeditions to Algeria and to California; Smithsonian Institution: Washington, DC, USA, 1915; Volume 65. [Google Scholar]
  312. Brunt, D. Notes on radiation in the atmosphere. I. Q. J. R. Meteorol. Soc. 1932, 58, 389–420. [Google Scholar] [CrossRef]
  313. Brutsaert, W. On a derivable formula for long-wave radiation from clear skies. Water Resour. Res. 1975, 11, 742–744. [Google Scholar] [CrossRef]
  314. Idso, S.B.; Jackson, R.D. Thermal radiation from the atmosphere. J. Geophys. Res. 1969, 74, 5397–5403. [Google Scholar] [CrossRef]
  315. Idso, S.B. A set of equations for full spectrum and 8-to 14-μm and 10.5-to 12.5-μm thermal radiation from cloudless skies. Water Resour. Res. 1981, 17, 295–304. [Google Scholar] [CrossRef]
  316. Prata, A. A new long-wave formula for estimating downward clear-sky radiation at the surface. Q. J. R. Meteorol. Soc. 1996, 122, 1127–1151. [Google Scholar] [CrossRef]
  317. Bastiaanssen, W.G.M. Regionalization of Surface Flux Densities and Moisture Indicators in Composite Terrain: A Remote Sensing Approach under Clear Skies in Mediterranean Climates; Wageningen University and Research: Wageningen, The Netherlands, 1995. [Google Scholar]
  318. Swinbank, W.C. Long-wave radiation from clear skies. Q. J. R. Meteorol. Soc. 1963, 89, 339–348. [Google Scholar] [CrossRef]
  319. Dilley, A.; O’brien, D. Estimating downward clear sky long-wave irradiance at the surface from screen temperature and precipitable water. Q. J. R. Meteorol. Soc. 1998, 124, 1391–1401. [Google Scholar] [CrossRef]
  320. Satterlund, D.R. An improved equation for estimating long-wave radiation from the atmosphere. Water Resour. Res. 1979, 15, 1649–1650. [Google Scholar] [CrossRef]
  321. Garratt, J. Extreme maximum land surface temperatures. J. Appl. Meteorol. Climatol. 1992, 31, 1096–1105. [Google Scholar] [CrossRef] [Green Version]
  322. Jacobs, J. Radiation climate of broughton island. Energy Budg. Stud. Relat. Fast-Ice Break. Processes Davis Strait 1978, 105–120. [Google Scholar]
  323. Maykut, G.A.; Church, P.E. Radiation climate of barrow Alaska, 1962–1966. J. Appl. Meteorol. Climatol. 1973, 12, 620–628. [Google Scholar] [CrossRef] [Green Version]
  324. Waters, R.; Allen, R.; Bastiaanssen, W.; Tasumi, M.; Trezza, R. Sebal. Surface Energy Balance Algorithms for Land. Idaho Implementation. Advanced Training and Users Manual, Idaho, USA. 2002. Available online: https://www.posmet.ufv.br/wp-content/uploads/2016/09/MET-479-Waters-et-al-SEBAL.pdf (accessed on 6 February 2022).
  325. Tasumi, M. Progress in Operational Estimation of Regional Evapotranspiration Using Satellite Imagery; University of Idaho: Arlington, VA, USA, 2003. [Google Scholar]
  326. Friedl, M.A. Forward and inverse modeling of land surface energy balance using surface temperature measurements. Remote Sens. Environ. 2002, 79, 344–354. [Google Scholar] [CrossRef]
  327. Ge, N.; Zhong, L.; Ma, Y.; Cheng, M.; Wang, X.; Zou, M.; Huang, Z. Estimation of land surface heat fluxes based on Landsat 7 ETM+ data and field measurements over the northern Tibetan Plateau. Remote Sens. 2019, 11, 2899. [Google Scholar] [CrossRef] [Green Version]
  328. Thompson, N.; Barrie, I.; Ayles, M. The Meteorological Office Rainfall and Evaporation Calculation System: Morecs. 1981. Available online: https://catalogue.ceh.ac.uk/documents/b9155463-ac86-4e19-a24f-57cef6A24505 (accessed on 6 February 2022).
  329. Hansen, F.V. Surface Roughness Lengths; Army Research Lab White Sands Missile Range Nm: Adelphi, MD, USA, 1993. [Google Scholar]
Figure 1. The flow diagram of the current review.
Figure 1. The flow diagram of the current review.
Energies 15 01264 g001
Figure 2. Schematic diagrams of different PM-based models (ra = aerodynamic resistance (s/m); rs = surface resistance (s/m), ra,c = aerodynamic resistance for the canopy component (s/m); ra,s = aerodynamic resistance for the soil component (s/m); rs,c = surface resistance for the canopy component (s/m); rs,s = surface resistance for the soil component (s/m)).
Figure 2. Schematic diagrams of different PM-based models (ra = aerodynamic resistance (s/m); rs = surface resistance (s/m), ra,c = aerodynamic resistance for the canopy component (s/m); ra,s = aerodynamic resistance for the soil component (s/m); rs,c = surface resistance for the canopy component (s/m); rs,s = surface resistance for the soil component (s/m)).
Energies 15 01264 g002
Table 1. A summary of the MT-based LSMs that follow approach 2.
Table 1. A summary of the MT-based LSMs that follow approach 2.
ModelTypeMain AssumptionsAdvantagesDisadvantagesMosaic Scheme
BATSTwo-layerUsing the demand–supply approach for calculating evaporation
Heat capacity of foliage is assumed to be zero; therefore, the energy of the photosynthesis/respiration process is neglected
Offline and Online version
High vertical resolution
When the soil is dry, it will underestimate the latent heat flux
Low horizontal resolution
SiBTwo-layerThe soil heat flux is modeled by using a Slab model
Using an explicit backward-differencing scheme to calculate the T c   and T g s
The variations of leaf area density are assumed to be constant over height
High vertical resolution
Providing better estimates for the time-varying exchanges of water, energy, and carbon for croplands
Containing a crop phenology module to specifically simulate corn, soybeans, and wheat (SiBcrop)
Complicated calculation
Difficulty capturing spatial heterogeneity
Low horizontal resolution
MosaicTwo-layerUsing the ‘‘mosaic” strategy to consider sub-grid heterogeneity
The energy balance equation for each sub-grid is simplified enough to be written in Penman–Monteith form
Computationally efficient
Flexibility (various vegetation types)
Low horizontal resolution
Modeling a Savanna as a mosaic can lead to significant errors
VICTwo-layerHydrologically based
Using the mosaic scheme
Evaporation and transpiration are parameterized by the Penman–Monteith formulation
Fine resolution (vertical and horizontal)
Suitable performance over larger regions
Open-source availability
Inefficiency in snowpack-related parameterizations
Due to the non-closure of the energy budget, some conflicting results may arise
NoahTwo-layerLinearization of the surface energy balance for the Tskin calculation
The prediction of the ith soil layer temperature is performed using the fully implicit Crank–Nicholson scheme
Advanced snowpack-related physics
Offline and Online version
Insufficient consideration of vertical soil profile heterogeneity
Under the conditions of high radiative forcing, it will underestimate the LST
Table 2. A summary of the remote sensing ET models that follow approach 3.
Table 2. A summary of the remote sensing ET models that follow approach 3.
TypeModelData RequirementsLST ConfigurationMain AssumptionStrengthWeaknessTemporal ResolutionSpatial Scale
One-SourceSEBILST, W, Tpbl, hpbl,Absolute LSTBased on the CWSI and pixel-wise conceptPractical even for surfaces without wet and dry pixelsCharacteristics of the planetary boundary layer are needed
Relatively poor accuracy
DayRegional
SEBALLST, NDVI,
α , W, Ta, RH
Surface–air temperature gradientInternal calibration by anchor pixels through a subjective procedure
Uses the air surface temperature gradient instead of absolute LST
Defining a linear relationship between LST and dT
Minimum ground data requirement
Able to use slope and aspect in heterogeneous lands
Physical concept
Internal calibration
Solves all the energy fluxes
Time-consuming
Cost-intensive
Limited by different stability atmospheric conditions and heterogeneous surfaces
Uncertainties due to user’s judgment in selecting anchor pixels
Day, Month, Season, YearField to regional
SEBS LST ,   NDVI ,   α , hpbl, ea, es,
W, Ta, RH,
Absolute LSTEF is used by calibrating the limit points through a non-subjective procedureApplicable for all atmospheric stability conditions
Physical concept
No need for prior knowledge on turbulent heat fluxes
Calculates roughness length by a developed submodule
Selects the reference pixels automatically
Complex structure
High data requirements
High sensitivity to the air surface temperature gradient and aerodynamic parameters
Day, MonthLocal to regional
S-SEBILST, NDVI, α, T a Absolute LSTThe process of determining the wet and dry conditions is albedo dependentNo need for additional data
Dependency on LST is less than other models
Capable of regulating variations of LST in proportion to albedo
Temperatures for dry and wet conditions can be determined by the image itself
Does not solve H
Applicable for homogenous areas with constant atmospheric conditions
Dependency of extreme LSTs on location
Day, MonthBasin
METRICLST, NDVI, LAI, W, RH, Ta, ShSurface–air temperature gradientInternal calibration by anchor pixels through a subjective procedure
Using the air surface temperature gradient instead of absolute LST
Defining a linear relationship between LST and dT
Minimum ground data requirement
Able to use slope and aspect in heterogeneous lands
Solves all the energy fluxes
Physical concept
Internal calibration
Uses a soil water balance to determine the hot pixel
Time-consuming
Cost-intensive
Uncertainties due to user’s judgment in selecting anchor pixels
ET underestimation in arid and non-agricultural areas
Day, Month, Season, YearField to regional
SSEBLST, NDVI, DEM, W, RH, Ta, Sh Surface–air temperature gradientJoint use of LST and reference ETSimple
Cost-effective
Operational model
Provides the rapid ET estimates over large regions
Does not solve H and G
ET underestimation in surfaces with low albedo and ET overestimation in surfaces with high G and high albedo
High sensitivity to LST
Day,
Month,
Season
Basin to regional
SSEBopLST, Ta, Albedo, NDVISurface–air temperature gradientSelecting the reference pixels by a non-subjective procedure
ET is linearly scaled between evapotranspiration of hot and cold pixels (in proportion to LSTs of hot and cold pixels),
dT is obtained from a linear relationship with LST
Simple, cost-effective, and operational model
Low computational effort
Selects the reference pixels automatically
Does not solve H and G
Does not use surface and aspect in heterogeneous lands
Day, Month, Season, YearBasin to regional
Two-SourceTSEBLST, W, Ta,
LAI
LST observations at one or two viewing anglesEnergy fluxes can be determined by one or two LST observationsUsable at conditions with high VPD and low meteorological dataDetermination of Priestly–Taylor and Penman–Monteith coefficients Day, Month, SeasonLocal to regional
TSTIMLST, LAI, WTemporal changes in LSTTime-integrated structure
Uses LST data at two times (t1 and t2)
No need for measuring T a
Less sensitive to systematic biases of LST
Requires LST data with a high temporal resolutionDay, Month, Season, YearRegional to continental
ALEXILST, W,
LAI, LULC
Temporal changes in LSTTime-integrated structure
Considering temporal variations of LST
Satellite-based
Generates maps of daily energy fluxes as well as soil moisture over large scalesRequires LST data with a high temporal resolutionDayContinental
DTDLST, LAI,
Ta, W
Two LST observations at two timesTime-integrated structureSimple structure with a few input data
Operational model
Less sensitive to errors of absolute LST and meteorological data,
No need for modeling boundary layer development
Determines the Priestley–Taylor coefficient
Requires LST data with a high temporal resolution
DayRegional
TTMENDVI, LST, α , Ta, WLST observation at a single viewing angleVI/LST space-based
Uses a trapezoidal framework with specified boundary conditions
Patch approach for estimating ET
Simple structure with few input data requirements
No need for computing the resistance network
Misestimates energy fluxes due to the temperature-dependent cold edge
Different selections of wet/dry edgesNeeds a flat surface with a large number of pixels
DayRegional
HTEMLAI, Albedo, Ta,
LST, ea,
NDVI, W
LST observation at one viewing angleUses a trapezoidal framework for estimating energy fluxesHybrid dual-source scheme of the patch and layer methods
No need for computing the resistance network
Inappropriate in complex surfaces
Different selections of wet/dry edges
Needs a flat surface with a large number of pixels
DayRegional
ETEMLLST, Albedo, LULC, LAI, NDVI, Crop height, W,
T a , VPD
Surface–air temperature differenceBased on a theoretical VI/LST space for each pixelAppropriate for complex and heterogeneous conditions
No need for computing the resistance network
Too many inputsDayLocal to regional
NDVI = Normalized Difference Vegetation Index; LAI = Leaf Area Index; α = surface albedo; T a = air temperature; W = wind speed; ea = actual vapor pressure; es = saturated vapor pressure; RH = relative humidity; Tpbl = average temperature of planetary boundary layer; hpbl = height of planetary boundary layer; LULC = Land Use Land Cover; Sh = sunshine hours.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Taheri, M.; Mohammadian, A.; Ganji, F.; Bigdeli, M.; Nasseri, M. Energy-Based Approaches in Estimating Actual Evapotranspiration Focusing on Land Surface Temperature: A Review of Methods, Concepts, and Challenges. Energies 2022, 15, 1264. https://0-doi-org.brum.beds.ac.uk/10.3390/en15041264

AMA Style

Taheri M, Mohammadian A, Ganji F, Bigdeli M, Nasseri M. Energy-Based Approaches in Estimating Actual Evapotranspiration Focusing on Land Surface Temperature: A Review of Methods, Concepts, and Challenges. Energies. 2022; 15(4):1264. https://0-doi-org.brum.beds.ac.uk/10.3390/en15041264

Chicago/Turabian Style

Taheri, Mercedeh, Abdolmajid Mohammadian, Fatemeh Ganji, Mostafa Bigdeli, and Mohsen Nasseri. 2022. "Energy-Based Approaches in Estimating Actual Evapotranspiration Focusing on Land Surface Temperature: A Review of Methods, Concepts, and Challenges" Energies 15, no. 4: 1264. https://0-doi-org.brum.beds.ac.uk/10.3390/en15041264

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