Skip to main content
Scientific Reports logoLink to Scientific Reports
. 2023 Sep 2;13:14465. doi: 10.1038/s41598-023-40499-6

Assessment of atmospheric emissivity models for clear-sky conditions with reanalysis data

Luis Morales-Salinas 1,, Samuel Ortega-Farias 2, Camilo Riveros-Burgos 3,4, José L Chávez 5, Sufen Wang 6, Fei Tian 6, Marcos Carrasco-Benavides 7, José Neira-Román 7, Rafael López-Olivari 8, Guillermo Fuentes-Jaque 1,9
PMCID: PMC10475081  PMID: 37660172

Abstract

Atmospheric longwave downward radiation (Ld) is one of the significant components of net radiation (Rn), and it drives several essential ecosystem processes. Ld can be estimated with simple empirical methods using atmospheric emissivity (εa) submodels. In this study, eight global models for εa were evaluated, and the best-performing model was calibrated on a global scale using a parametric instability analysis approach. The climatic data were obtained from a dynamically consistent scale resolution of basic atmospheric quantities and computed parameters known as NCEP/NCAR reanalysis (NNR) data. The performance model was evaluated with monthly average values from the NNR data. The Brutsaert equation demonstrated the best performance, and then it was calibrated. The seasonal global trend of the Brutsaert equation calibrated coefficient ranged between 1.2 and 1.4, and the K-means analysis identified five homogeneous zones (clusters) with similar behavior. Finally, the calibrated Brutsaert equation improved the Rn estimation, with an error reduction, at the worldwide scale, of 64%. Meanwhile, the error reduction for each cluster ranged from 18 to 77%. Hence, Brutsaert’s equation coefficient should not be considered a constant value for use in εa estimation, nor in time or location.

Subject terms: Climate sciences, Environmental sciences

Introduction

Thermal atmospheric emissivity (εa) is a parameter mainly used to estimate the atmospheric downward longwave radiation (Ld), which is used to determine net radiation (Rn). In this way, the Rn plays a significant role in modeling natural phenomena such as vegetation evapotranspiration rates, snowmelt, and frost occurrence1. The longwave radiation emitted by the atmosphere occurs at wavelengths between 4 and 100 μm in the electromagnetic spectrum and is influenced mainly by water vapor, carbonic anhydride, ozone, and clouds2.

The Ld is one of the significant components of the Rn model applied to forests3, affecting several essential ecosystem processes, such as photosynthetic rate, plant respiration, and primary productivity. However, the Ld term is not easy to measure. Although pyrgeometers are used to measure it, their high cost4,5 often limits their inclusion in automatic weather stations.

Then, the Ld term can be estimated using simple empirical methods adjusted based on direct measurements to solve this problem. In this regard, the literature shows that reliable estimations of Ld can be obtained using a statistical adjustment or calibration based on measured air temperature and relative humidity data measured at the surface (at screen height). However, local calibration is required6 for accurate results. Another approach is to use radiosonde data and atmospheric radiative transfer models, but this information limits their application to a specific date and place713.

Although an extensive database is needed, estimating εa through statistical models could be significantly improved using different parameterizations based on atmospheric conditions. This database should be able to represent a range of atmospheric conditions for every location in the world to decrease the error of statistical estimations of εa.

Therefore, local calibrations of the coefficients involved in the empirical εa models are necessary. Research indicates that the majority of equations used to estimate εa are only valid for clear-sky days, reaching more accurate results when considering daily or climatological averages. Clear-sky days are defined by the absence of visible clouds in the sky. In this sense, the clear-sky conditions are defined when the clear-sky index (Global solar radiation/Extraterrestrial solar radiation) is approximately greater than 0.914,15. In order to achieve this, it is possible to use existing databases such as NCEP/NCAR reanalysis to obtain more accurate estimations of the emissivity at the Earth’s surface16,17.

According to the literature, using semiempirical approaches to estimate εa has inherent errors linked to instrument-based measurement deviation or uncertainty. Therefore, the use of a properly calibrated εa model is a viable alternative for estimating more accurately at specific locations using meteorological variables such as air temperature and relative humidity6,1820. In this regard, there is plenty of information in the literature about the estimation of Rn using empirical and semiempirical εa models based mainly on ground-based instruments for specific locations of different roughness surfaces worldwide2036.

In this study, the performance of eight models was evaluated to determine their accuracy in the estimation of atmospheric emissivity for different locations worldwide. Climatic data from NCEP/NCAR reanalysis (NNR) were obtained from a dynamically consistent scale resolution of basic atmospheric quantities and computed parameters. Among the evaluated models, the one with the best statistical performance was calibrated on a global scale using a parametric instability analysis approach. In this way, one of the main contributions of the study at hand was to improve the Rn computation over homogeneous latitude areas globally, reducing the need for local calibration of atmospheric emissivity.

Information derived from NCEP/NCAR reanalysis data (NNR)

Exploratory analysis for the NCEP/NCAR reanalysis data (1948–2020) across the world revealed that the air temperature (ta) varied between – 37 °C and 49 °C, with an average of 17 °C. Additionally, the actual vapor pressure (ea) values ranged from 0.01 to 21.9 kPa with an average value of 5.02 kPa; meanwhile, εa averaged 0.73 and varied between 0.34 and 0.97. Moreover, the variable with the most significant variation coefficient was ta, presenting a value of 447.7%, ea with 107.6%, and εa reached an 18.5% variation.

Figure 1 shows the observed values of εa obtained using the NNR data throughout the year’s seasons. In Fig. 1, a spatial pattern can be seen due to the formation of homogeneous groups or units (clusters) based on latitude. Moreover, this trend was also observed on a monthly scale (data not shown). These clusters have temporal variability related to atmospheric dynamics. Additionally, εa presents homogeneous values overseas and in oceans; likewise, the poles show the same trend but with different absolute values. On the continents, εa have variations related to the topography, land use, and closeness to seas and oceans. That trend was also observed in other study37, which indicated that uncertainties in the computation of land surface temperature, can be highly influenced by the spatial variability of the ground.

Figure 1.

Figure 1

Maps of climatological world atmospheric emissivity (εa) for (a) winter, (b) spring, (c) summer, and (d) autumn, calculated from NCEP/NCAR reanalysis data. This figure was obtained with R software83.

At the spatial resolution scale of the NNR, climatological variables merely correspond to a spatial trend as the NNR's spatial estimate being approximately 250 km. Comparing this data with surface weather station networks at this resolution is difficult since they correspond merely to an average in each grid element or “pixel”38,39. However, several studies have established the agreement between grid and ground-based observations for variables such as solar radiation, air temperature, relative humidity, precipitation, and pressure40. Wind speed exhibits the largest biases in space recorded, compared with other reanalysis products like ERA-40, ERA-Interim, or ERA5. Although the NNR presents significant differences, the spatial calculation resolution is different, adding additional elements, and making it challenging to make direct comparisons38,4044. It is noteworthy that evapotranspiration calculated from NNR data is comparable to those calculated from observations at most weather stations39,45.

Performance of atmospheric emissivity models

Descriptive analysis for the εa evaluation using the eight models is presented in Table 1. This table shows that the estimated values of εa were between 0.22 and 0.99, and the average values ranged from 0.61 to 0.83. Moreover, the Bastiaanssen model exhibited the lowest variation, with a variation coefficient of 1.1%; meanwhile, the Brutsaert model showed the highest variation coefficient, with a value of 28.1%. The other models obtained an intermediate variation with values from 6.4 to 20.1%.

Table 1.

Descriptive statistics of estimated atmospheric emissivity (dimensionless) for each model.

Model Minimum Maximum Median Average Standard deviation Variation coefficient (%)
Bastiaanssen 0.71 0.76 0.76 0.75 0.0083 1.1
Prata 0.67 0.85 0.71 0.72 0.0512 7.1
Idso 0.70 0.89 0.74 0.75 0.0480 6.4
Brutsaert 0.22 0.85 0.66 0.61 0.1701 28.1
Idso and Jackson 0.74 0.99 0.82 0.83 0.0737 8.9
Swinbank 0.37 0.89 0.68 0.66 0.1330 20.1
Brunt 0.61 0.83 0.69 0.69 0.0627 9.1
Angstrom 0.65 0.82 0.72 0.72 0.0602 8.3

Table 2 presents the performance of the eight models, showing that the minimum and maximum values of the root mean square error (RMSE) were 0.097 and 0.216, respectively. Meanwhile, the coefficient of determination (r2) ranged from 0.45 to 0.69, while the Akaike information criterion (AIC) presented a minimum value of − 2,354,713 and a maximum of − 1,745,909. Considering all statistical parameters such as systematic error (BIAS), mean absolute error (MAE), RMSE, normalized RMSE (nRMSE), coefficient of determination (r2), index of agreement (d) and AIC, the Idso and Jackson model presented the poorest performance (BIAS = 0.097, MAE = 0.143, RMSE = 0.216, nRMSE = 34%, r2 = 0.45, d = 0.25, and AIC = − 1,745,909), while the εa Brutsaert model had the best performance (BIAS = − 0.127, MAE = 0.128, RMSE = 0.152, nRMSE = 23.9%, r2 = 0.76, d = 0.80, and AIC = − 2,354,713).

Table 2.

Comparative statistics for the performance of the eight models for estimating atmosphere emissivity, using processed NCEP/NCAR reanalysis data.

Model BIAS MAE RMSE nRMSE r2 d P AIC
Bastiaanssen 0.019 0.099 0.131 20.6 0.50 0.23 **  − 1,996,608
Prata  − 0.011 0.084 0.103 16.2 0.57 0.68 **  − 1,949,235
Idso 0.020 0.076 0.105 16.6 0.59 0.65 **  − 1,973,122
Brutsaert 0.127 0.128 0.152 23.9 0.69 0.80 **  − 2,354,713
Idso and Jackson 0.097 0.143 0.216 34.0 0.45 0.25 **  − 1,745,909
Swinbank  − 0.070 0.089 0.108 17.0 0.66 0.85 **  − 2,058,430
Brunt  − 0.041 0.088 0.101 16.0 0.64 0.75 **  − 2,058,430
Angstrom  − 0.010 0.079 0.097 15.3 0.60 0.74 **  − 2,018,959

BIAS, MAE and RMSE are the systematic error, mean absolute error, and root mean square error, respectively. The units are dimensionless. nRMSE is the normalized root mean square and corresponds to a percentage. r2 is the coefficient of determination, and d is the index of agreement (dimensionless). The AIC is the Akaike information criterion (dimensionless).

Global calibration for the best model

The better performance of the Brutsaert model for estimating εa simplifies the global calibration process, considering that only one parameter remains dependent, so that the exponent’s hypothesis is invariant.

Supplementary Fig. 1 shows the seasonal behavior of εa, revealing a consistent linear and positive regression with (ea/ta), independent of the season. The plot shows two separated data tendencies during the winter season with the upper right side of the graph being the most important as it concentrates more points over a linear trend.

Only a trend was evident for the spring season, with a higher concentration of points from 0.7 to 1.0. For the summer months, the graph presents the most irregular linear regression of all seasons, with the cluster of points concentrated in the top portion of Supplementary Fig. 1c, in a range of 0.8 to 1.0.

Finally, in the autumn season, two linear regressions can be identified with different slopes. However, the most important trend is located in the upper portion of the graph, where εa are concentrated in a range from 0.8 to 1.0.

The performance summary of the parametric instability analysis through the geographically weighted regression (GWR) of the spatial variation for Brutsaert equation parameters, is presented in Supplementary Table 1. The negative BIAS values show that the GWR coefficients underestimated the spatial variability of the Brutsaert equation parameters. Furthermore, the BIAS depicted a random behavior. The monthly mean RMSE was approximately 0.022 (dimensionless) with an estimation error of 1.5%. The RMSE values for autumn and winter were above the mean, while the RMSE values for spring and summer were below the mean.

As a result, the months that were closer to the average maximum temperature showed a more accurate estimation of the parameters in the Brutsaert equation compared to the colder months near the average minimum temperature. AIC values showed a similar pattern to RMSE; thus, warmer months resulted in a better AIC value than colder months. The Nash–Sutcliffe efficiency (NSE) index, d, and r2 indices had values near 1, indicating a good fit of the GWR coefficients for each month.

Figure 2 shows the global seasonal trend of calibrated Brutsaert model coefficient and the cluster resulting from the K-means analysis. The calibrated coefficient value ranged between 1.2 and 1.4, considering the four seasons and the five zones with similar behavior. In this sense, the Brutsaert model coefficient did not present a unique value for the entire world. The predominant zone was related to the Ecuador line, which covered a critical zone of the study area.

Figure 2.

Figure 2

World spatial distribution of the calibrated coefficient of Brutsaert for (a) winter, (b) spring, (c) summer, and (d) autumn. Additionally, the homogeneous areas or clusters are presented (e), referenced for the Northern Hemisphere. This figure was obtained with R software83.

Additionally, the austral and boreal zones presented differentiation from the rest of the world. The monthly mean empirical coefficient of the Brutsaert model is shown in Supplementary Table 2. Moreover, the performance of the uncalibrated and calibrated Brutsaert equations in computing net radiation for each cluster is presented in Table 3. Here, a RMSE reduction was observed at a worldwide scale of 64%, while in Cluster 2, a RMSE decrease of approximately 77% was observed. However, Cluster 3, which mainly represents a significant portion of the land, only reached an RMSE reduction of 18%.

Table 3.

Comparison of statistical indices for evaluating the Brutsaert equation effect on net radiation computation.

Zone Model BIAS (W m–2) MAE (W m–2) RMSE (W m–2) nRMSE (%) NSE d r2
Cluster 1 Uncalibrated Brutsaert’s – 37.5 37.5 40.1 46.4 0.78 0.95 0.99
Calibrated Brutsaert’s  − 4.5 6.1 7.5 8.7 0.99 1.00 1.00
Cluster 2 Uncalibrated Brutsaert’s  − 26.9 27.1 30.9 36.3 0.87 0.97 0.99
Calibrated Brutsaert’s  − 2.3 5.5 7.0 8.2 0.99 1.00 0.99
Cluster 3 Uncalibrated Brutsaert’s  − 3.8 9.1 11.4 22.5 0.95 0.99 0.96
Calibrated Brutsaert’s 1.7 7.2 9.3 18.2 0.97 0.99 0.97
Cluster 4 Uncalibrated Brutsaert’s  − 16.9 17.1 20.1 25.8 0.93 0.98 0.99
Calibrated Brutsaert’s 7.6 9.0 10.6 13.6 0.98 1.00 0.99
Worldwide Uncalibrated Brutsaert’s  − 15.2 18.1 23.6 32.0 0.90 0.98 0.97
Calibrated Brutsaert’s  − 0.2 6.5 8.4 11.3 1.00 1.00 0.99

BIAS, MAE, and RMSE are the systematic error, mean absolute error, and root mean square error, respectively. The nRMSE is the normalized root mean square error, and its unit is %. The NSE is the Nash–Sutcliffe model efficiency index, d is the index of agreement, and r2 is the coefficient of determination (dimensionless). The information is only presented for four out of the five clusters.

It is important to note that the spatial resolution of the model used is approximately 250 km, which allows the calculation of meteorological variables that correspond to large climatic regions across the Earth.

The spatial configuration of the variables is influenced by factors such Earth's topography, oceans, and land surface cover, which affects variables such as albedo and surface emissivity. The Seasonal dependence is strongly associated with the Earth's trajectory in its solar orbit, affecting the incident energy of short and long waves, adjusting to the solar declination29,46,47.

Improvements of estimated net radiation

In high latitudes (C1 in Fig. 2), a calibrated Brutsaert model coefficient demonstrated good agreement between the observed and estimated values of Rn (Fig. 3b) values using a calibrated Brutsaert model coefficient (Cluster 1 in Table 3). Low error values were observed for BIAS, MAE, and RMSE, − 4.5, 6.1, and 7.5 W m–2, respectively.

Figure 3.

Figure 3

Comparison of net radiation computed using uncalibrated (a,c,e,g) and calibrated (b,d,f,h) Brutsaert’s equation in the estimated worldwide clusters: Cluster 1 (a,b); Cluster 2 (c,d); Cluster 3 (e,f); and Cluster 4 (g,h).

At the same time, the NSE, d, and the r2 had values close to 1.0 and a value of nRMSE equal to 8.7%. In this context, estimated Rn for latitudes greater than 60° N under all-sky conditions from MODIS imagery, was validated using data from 82 sites and eight different observation networks27.

These authors found acceptable accuracy values of RMSE ranging between 15.04 and 23.66 W m–2, whereas the values of r2 and BIAS were between 0.51 and 0.85 and between − 0.08 and 0.27 W m2, respectively. However, for Alaska’s Arctic tundra summer conditions (at USA sites Fen, Tussock, and Heath, the latitude of 68° N), the estimated Rn for all-sky conditions aligned well with the observed values, presenting an average RMSE of 23 W m–2 and r2 value equal to 0.99 using the remote sensing thermal-based two-source energy balance model48.

Errors were found using a similar value for the original Brutsaert coefficient (1.25 ± 0.009), which was consistent with our study for summer condition (S Table 1). When estimating the Rn in the geographic areas corresponding to Cluster 2 (Table 3), latitudes are between 40 and 60°N (C2 in Fig. 3), and lower errors were observed, with values of 49, 10, and 7% for BIAS, MAE and RMSE, respectively. Simulations of Rn in different zones, locations, and vegetation surfaces have been performed within a range of latitudes34, 41–60°N in Canada (Eloria, Ontario, with a latitude of 43°N) using an empirical Rn-Model observed an average MAE, r2 and d of approximately 28 W m–2, 0.98 and 0.99, respectively. At the same latitude but different locations in Avignon, France30 (latitude 43°N), the Rn was evaluated using a semiempirical model based on Stefan–Boltzmann under grass cover, observing a RMSE of 34 W m-2 with a calibrated Brutsaert’s coefficient of 1.31.

Also, the Rn has been estimated for clear-sky and all-wave net radiation combined visible and shortwave infrared (VSWIR) and thermal infrared (TIR) remote sensing data at a location in Montana, USA (Fort Peck, latitude 48° N)49, observing that the component-based approach presented a BIAS, RMSE and r2 of 76.7, 2.0 and 0.87, respectively. Using a direct estimation approach, the BIAS, RMSE, and r2 values were 52.3, − 1.5, and 0.94, respectively. Furthermore, Rn estimations from solar shortwave radiation measurements and conventional meteorological observations (or satellite retrievals) were conducted at 24 different sites50. Three of these sites were located at latitudes over 42° N (Fort Peck, MT in grass cover; Sioux Falls, SD in grass cover; and Wind River, WA in temperate evergreen forest cover). Thus, the errors obtained at those sites were 17.2, 20.8, and 16.0 W m–2 for RMSE and 3.5, 2.6, and 4.6 W m–2 for BIAS, respectively. For mid-latitudes (C3 in Fig. 3), the estimation of Rn (Fig. 3f), using a calibrated Brutsaert’s coefficient (Cluster 3 in S Table 2), presented a statistical mean deviation lower than 9.3 W m–2. The evaluation of the different Rn models presented in the literature worldwide are mainly inserted between latitudes 42°N and 40°S. In this context, Rn was estimated using MODIS data for clear sky days with an empirical εa equation51,52, and this study covered most of Oklahoma and the southern part of Kansas, USA (latitude from approximately 34.5°to 38.5°N and longitude from 95.3° to 99.5°W).

Thus, the comparison between observed and simulated Rn presented 59 W m–2, 74 W m–2, and 0.89 for BIAS, RMSE, and r2, respectively. On the other hand, for the climate of Baghdad, Iraq (latitude 33°N) in natural prairie grass, there was a good agreement between observed and estimated Rn with a simple empirical approach for all clear skies53, with an average RMSE value equal to 28 W m–2 and r2 value of 0.984.

However, in a semiarid climate covered by sparse vegetation near Tabernas54, Almería, Spain (37°N) good agreement was obtained between the observed and simulated Rn, with a mean RMSE value of 47 W m–2 and r2 value of 0.96. In another study conducted on an olive vegetation surface in southwestern Sicily, Italy (37°N)55, an RMSE value of 35.4 W m–2 was obtained in the Rn. They used a semiempirical model based on the Stefan–Boltzmann law that included estimating longwave radiation incorporating the original Brutsaert’s coefficient value. Furthermore, in a wetland in the Paynes Prairie Preserve (29° N) in north-central Florida, USA56, the Rn was estimated using GOES satellite data. Thus, the Rn was best characterized when the GOES solar and GOES longwave radiation products were combined, reaching average RMSE, NSE, and r2 values equal to 14.1 W m–2, 0.92, and 0.95, respectively.

In another experience, in the Upper Blue Nile Basin, Ethiopia (7.5°–12.5° N)57 the Rn distribution was estimated from satellite MODIS and automatic weather station, obtaining a reduction in the mean bias (MB) and RMSE values of 76.43 and 17.71%, respectively, by implementing the new recalibrated Brutsaert equation.

Furthermore, the solar shortwave radiation data and meteorological observations (or satellite retrievals) from 24 different sites (USA and China) were used to estimate Rn50, where 21 of them were in latitudes between 32° and 41°N under various land covers (grass, pastures, wheat, rangeland, crop, forest, native prairie, and desert). Across all sites and land covers, the BIAS varied between 19.7 and 27.8 W m–2, while the RMSE ranged from 12.8 to 21.0 W m–2.

For grass cover, a semiempirical model based on the Stefan–Boltzmann law was evaluated to estimate Rn in Talca, Chile (35°S)30, obtaining an RMSE of 42 W m–2 with a calibrated Brutsaert’s εa coefficient equal to 1.31. Furthermore, for olive tree cover (Pencahue Valley site, Pencahue, Chile, latitude 35°S and CIFA Experimental Station site, Córdoba, Spain, latitude 37.8°N) and vineyard cover (Pencahue Valley site, Pencahue, Chile, latitude 35°S), the estimation of Rn was observed with RMSE and MAE values below 45.0 and 31.0 W m–2, respectively25,35,36. For latitudes from 41°S to 60°S (C4 in Fig. 3), the estimation of Rn (Fig. 3h), using a calibrated Brutsaert’s coefficient (Cluster 4 in Table 3), presented a statistical mean deviation lower than 10.6 W m–2.

In this case, there is limited literature about the estimation of Rn in southern latitudes. Thus, the approaches presented in this study are promising for improving the estimation of Rn by incorporating a calibrated Brutsaert’s εa equation coefficient, broken down by homogeneous latitudes separated into five zones around the world (Figs. 2e and 3). Additionally, the errors found in this study are lower and similar in the same case compared to those found in the existing literature. In this sense, it is necessary to further evaluate εa estimates under a larger number of land cover types, different vegetation architectures of surface roughness lengths, and other characteristics than the Brutsaert εa equation coefficient values obtained in this research.

Conclusions

A spatially explicit approximation method for calculating atmospheric emissivity (εa) has been investigated to improve the estimation of downward longwave radiation during the day and further enhance radiation calculations.

The study evaluated eight models globally to estimate air emissivity using the NCEP/NCAR Reanalysis database, which corresponds to spatial trends in each variable used, mainly due to the calculation resolution, which is 2.5° in latitude and longitude. The results showed that the Brutsaert εa model had the best performance (BIAS = − 0.127, MAE = 0.128, RMSE = 0.152, nRMSE = 23.9%, r2 = 0.76, d = 0.80, AIC = – 2,354,713), and it was calibrated using geographically weighted regression (GWR) for global use.

The calibrated values considerably improved the calculation of the components of the surface energy balance, reducing calculation errors in net radiation from 25.2 W m–2 (nRMSE = 32.6%) to 8.6 W m–2 (nRMSE = 12.0%). The study indicates that the Brutsaert εa model should not be considered to have a constant coefficient value in time or space. It is advisable to use the coefficients found in this work to minimize errors when calculating net radiation. Using a sinusoidal equation or spline-type interpolation to reproduce the temporal variability of the coefficients for each day of the year is recommended when using the average monthly coefficients of the Brutsaert equation to estimate the emissivity of the atmosphere at a daily level.

Methods

Due to the different climatic conditions, the entire world was used as the study area to achieve an adequate model evaluation and calibration. Observed climatic data were obtained from a dynamically consistent scale resolution of basic atmospheric quantities and computed parameters known as NCEP/NCAR reanalysis data (NNR). These data were produced by the US National Centers for Environmental Predictions (NCEP) and the National Center for Atmospheric Research (NCAR) based in Boulder, CO, USA16.

NCEP/NCAR reanalysis data

The NNR data of global climatic information cover the period from 1948 to the present. Its spatial resolution is 2.5° longitude and 2.5° latitude with a temporal resolution of one month, one day, or six hours, and diagnosed diabatic heating of 17 vertical isobaric levels from 1000 to 10 hPa5860. The NNR data were developed by the synergy of processes such as quality control, assimilation, interpolation, observed data acquired by ground and sea stations, planes, satellites, and atmospheric soundings, together with simulations of atmospheric general circulation models using the Climate Data Assimilation System (CDAS)58.

The data used in this research were based on the “Surface” and “Surface flux” sections and their upward solar radiation flux.

Atmospheric emissivity parameterizations

Below are the equations used to estimate εa with meteorological variables such as ta and actual vapor pressure (ea). The exception is the Bastiaanssen model21 because it estimates εa at a daily scale for any condition of cloudiness, only depending on atmospheric transmissivity (τsw). The Bastiaanssen model was calibrated61 and used in the satellite-based energy balance for mapping evapotranspiration with an internalized calibration (METRIC) model62. The eight evaluated models are the following2,12,13,21,52,6365:

εa=0.85-Lnτsw0.09,
εa=1-1+ξ·e-1.2+3.0·ξ,
εa=0.70+0.0000595·ea·e1500/Ta,
εa=1.24·ea/Ta17,
εa=1-0.261·e-0.000777·(273-Ta)2,
εa=0.0000092·Ta2,
εa=0.605+0.048·ea,
εa=0.83-0.18·10-0.067·ea,

where εa is the atmospheric emissivity (dimensionless) for clear-sky conditions based on air temperature (Ta, K), water vapor pressure (ea, hPa), atmospheric transmissivity (τSW, dimensionless), and altitude (z, meters above sea level). The z was obtained from the WorldClim data66 with a spatial resolution of 1 km. Moreover, ea was estimated as follows29,67:

ea=6.108·RH100·e17.27·tata+237.3,

where ea is the actual water vapor pressure (hPa), ta is the air temperature (°C), and RH is the relative humidity (%). Also, the τSW and ξ were calculated according to:

τsw=0.75+0.00002·z,
ξ=46.5·ea/ta,

The observed values of εa were calculated as follows:

εa=Ldσ·To4,

where To corresponds to the average temperature of the whole air profile (K) measured by a meteorological station, and Ld is the atmospheric longwave downward radiation (W m–2).

For this study, the To and Ld data was obtained from the reanalysis databases. On the other hand, the estimated values of εa from the eight models were calculated using ta and RH obtained from the same reanalysis databases.

The evaluation of the goodness of fit for each model was conducted using the monthly average values of the NCEP/NCAR reanalysis (NNR) data.

Statistical analysis

The evaluation of the goodness of fit for each model was conducted using the monthly average values of the NNR data through the determination of the systematic error68 (BIAS), mean absolute error68 (MAE), root mean square error68 (RMSE), normalized root mean square error69 (nRMSE), and coefficient of determination70 (r2) (Table 3). Additionally, the index of agreement (d) was used7075, as well as the Akaike information criterion (AIC)7678. Also, the Nash–Sutcliffe efficiency (NSE) index79 was used, and it can range from − ∞ to 1. An efficiency of 1 (NSE = 1) corresponds to a perfect match of modeled data to the observed data. An efficiency of 0 (NSE = 0) indicates that the model predictions are as accurate as the mean of the observed data, whereas an efficiency less than zero (NSE < 0) occurs when the observed mean is a better predictor than the model. Essentially, if the model efficiency is closer to 1, the model is more accurate. NSE is equivalent to the coefficient of determination (r2), thus ranging between 0 and 1.

Global calibration

The εa model with the best performance was adjusted globally using geographically weighted regression (GWR). For this analysis, the dependent variable was εa, and the independent variables were ea and Ta. The GWR was carried out using NNR data, where information on each pixel was extracted from the grids, generating a vector type point layer for every month.

The GWR is based on weighted least squares80, considering the distance between each point, and it is described with the following equation8082:

yi=a0ui,vi+kakui,vixik+δi,

where (ui, vi) corresponds to the coordinates of the ith point in the space, yi is the dependent variable value, x is the kth independent variable in the ith point, a0 and ak are the regression parameters in the ith point, and δi is the error in the ith point. The ak(ui, vi) coefficients were estimated as follows:

akui,vi=XT·Wui,vi·X-1·XT·Wui,vi·Y,

where the dependent and independent variables are in the Y and X matrices, respectively.

All calculations, statistical analyses, and figures were processed using R software83 and the libraries “raster”84, “rgdal”85, “hexbin”86, “hydroGOF”87, “topmodel”88, and “GWmodel89,90.

Net radiation improvements

Rn was computed at a global scale to evaluate the impact of the εa calibrated model on the traditional method of calculating net radiation. The Rn is the sum of downward (incoming) and upward (outgoing) shortwave and longwave radiation, which is also a measure of the available energy at an underlying surface. It is also the fundamental parameter that governs the climate of the planetary boundary layer and is the driving force for processes such as evapotranspiration, air and soil heating, and photosynthesis. The net radiation over the terrestrial surface can be calculated as follows21:

Rn=1-αR+L-L-1-εsL,

where Rn is the estimated net radiation (W m-2); R is the downward shortwave solar radiation (W m–2); L and L are the downward and upward longwave radiation, respectively (W m–2); α is the surface albedo (dimensionless); and εs is the surface emissivity (dimensionless). The components of the incoming and outgoing longwave radiation, respectively, are given by:

L=εaσTa4,
L=εsσTs4,

where εa is atmospheric emissivity (dimensionless); Ta is air temperature (K); Ts is the land surface temperature (K), which was obtained from the monthly mean MOD11C3 product91; and σ is the Stefan–Boltzmann constant (5.67 × 10–8 W m–2 K–4).

εs can be calculated from a simple linear regression using the normalized difference vegetation index or NDVI92, which is necessary to estimate the land surface temperature (LST). The values of εs were calculated as follows25:

εs=0.9585+0.0357·NDVI,

where the NDVI is obtained from the monthly mean MOD13A3 product93.

εa is determined according to Brutsaert’s64 method, where the observed Rn was computed with the shortwave and longwave radiation from the NCEP/NCAR reanalysis. Meanwhile, the estimated Rn was obtained using the uncalibrated and calibrated parameters of Brutsaert’s equation and calculating the longwave radiation using L and L equations.

Supplementary Information

Acknowledgements

This study was supported by the International and regional cooperation and exchange projects of the National Natural Science Foundation of China (51961125205), and the Chilean government through National Agency for Research and Development (ANID)/PCI (NSFC190013 and FSEQ210004) and ANID/FONDECYT (No 1161809).

Author contributions

L.M.S and S.O.F conceived the study. L.M.S. designed the research. L.M.S., C.R.B. and G.F.J. did the computations. L.M.S, S.O.F, C.R.B., J.L.C, M.C.B., J.N.R., R.L.O., and G.F.J. analyzed the results. L.M.S and C.R.B. wrote the paper with comments from S.O.F, J.L.C, S.W., F.T., M.C.B., J.N.R., and R.L.O.

Data availability

The datasets used and/or analysed during the current study available from the corresponding author on reasonable request.

Competing interests

The authors declare no competing interests.

Footnotes

Publisher's note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

The online version contains supplementary material available at 10.1038/s41598-023-40499-6.

References

  • 1.Flerchinger GN, Xaio W, Marks D, Sauer TJ, Yu Q. Comparison of algorithms for incoming atmospheric long-wave radiation. Water Resour. Res. 2009;45:1–14. [Google Scholar]
  • 2.Idso SB, Jackson RD. Thermal radiation from the atmosphere. J. Geophys. Res. 1969;74:5397–5403. [Google Scholar]
  • 3.Marthews TR, Malhi Y, Iwata H. Calculating downward longwave radiation under clear and cloudy conditions over a tropical lowland forest site: An evaluation of model schemes for hourly data. Theor. Appl. Climatol. 2012;107:461–477. [Google Scholar]
  • 4.Sellers WD. Physical Climatology. University of Chicago Press; 1965. [Google Scholar]
  • 5.Duarte HF, Dias NL, Maggiotto SR. Assessing daytime downward longwave radiation estimates for clear and cloudy skies in Southern Brazil. Agric. For. Meteorol. 2006;139:171–181. [Google Scholar]
  • 6.Choi M, Jacobs JM, Kustas WP. Assessment of clear and cloudy sky parameterizations for daily downwelling longwave radiation over different land surfaces in Florida, USA. Geophys. Res. Lett. 2008 doi: 10.1029/2008GL035731. [DOI] [Google Scholar]
  • 7.Wright J. Emisividad infrarroja de la atmosfera medida en Heredia, Costa Rica. Top. Meteor. Ocean. 1999;6:44–51. [Google Scholar]
  • 8.Dilley AC, O’Brien DM. 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]
  • 9.Andreas EL, Ackley SF. On the differences in Ablation seasons of Arctic and Antarctic Sea Ice. J. Atmos. Sci. 1982;39:440–447. [Google Scholar]
  • 10.Idso SB. 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]
  • 11.Clark, G. & Allen, C. The estimation of atmospheric radiation for clear and cloudy skies. In 2nd National Passive Solar Conference (AS/ISES), 675–678 (1978).
  • 12.Swinbank WC. Long-wave radiation from clear skies. Q. J. R. Meteorol. Soc. 1963;89:339–348. [Google Scholar]
  • 13.Brunt D. Notes on radiation in the atmosphere. I. Q. J. R. Meteorol. Soc. 1932;58:389–420. [Google Scholar]
  • 14.Marty C, Philipona R. The clear-sky index to separate clear-sky from cloudy-sky situations in climate research. Geophys. Res. Lett. 2000;27:2649–2652. [Google Scholar]
  • 15.Calbó J, González J-A, Pagès D. A method for sky-condition classification from ground-based solar radiation measurements. J. Appl. Meteorol. 2001;40:2193–2199. [Google Scholar]
  • 16.Gulev SK, Zolina O, Grigoriev S. Extratropical cycone variability in the Northern Hemisphere winter from the NCEP/NCAR reanalysis data. Clim. Dyn. 2001;17:795–809. [Google Scholar]
  • 17.Von Randow RCS, Alvalá RCS. Estimativa da radiação de onda longa atmosférica no pantanal sul mato-grossense durante os períodos secos de 1999 e 2000. Rev. Bras. Meteorol. 2006;21:398–412. [Google Scholar]
  • 18.Silva JB, et al. Evaluation of methods for estimating atmospheric emissivity in Mato-Grossense Cerrado. Ambient. e Agua Interdiscipl. J. Appl. Sci. 2019;14:1. [Google Scholar]
  • 19.Herrero J, Polo MJ. Parameterization of atmospheric longwave emissivity in a mountainous site for all sky conditions. Hydrol. Earth Syst. Sci. 2012;16:3139–3147. [Google Scholar]
  • 20.Crawford TM, Duchon CE. An improved parameterization for estimating effective atmospheric emissivity for use in calculating daytime downwelling longwave radiation. J. Appl. Meteorol. 1999;38:474–480. [Google Scholar]
  • 21.Bastiaanssen WGM. 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; 1995. [Google Scholar]
  • 22.Kjaersgaard JH, et al. Comparison of the performance of net radiation calculation models. Theor. Appl. Climatol. 2009;98:57–66. [Google Scholar]
  • 23.Ortega-Farias S, Poblete-Echeverría C, Brisson N. Parameterization of a two-layer model for estimating vineyard evapotranspiration using meteorological measurements. Agric. For. Meteorol. 2010;150:276–286. [Google Scholar]
  • 24.Irmak S, Mutiibwa D, Payero JO. Net radiation dynamics: performance of 20 daily net radiation models as related to model structure and intricacy in two climates. Trans. ASABE. 2010;53:1059–1076. [Google Scholar]
  • 25.López-Olivari R, Ortega-Farías S, Morales L, Valdés H. Evaluation of three semi-empirical approaches to estimate the net radiation over a drip-irrigated olive orchard. Chil. J. Agric. Res. 2015;75:341–349. [Google Scholar]
  • 26.Parry CK, et al. Comparison of a stand-alone surface renewal method to weighing lysimetry and eddy covariance for determining vineyard evapotranspiration and vine water stress. Irrig. Sci. 2019;37:737–749. [Google Scholar]
  • 27.Chen J, He T, Jiang B, Liang S. Estimation of all-sky all-wave daily net radiation at high latitudes from MODIS data. Remote Sens. Environ. 2020;245:111842. [Google Scholar]
  • 28.Riveros-Burgos C, Ortega-Farías S, Morales-Salinas L, Fuentes-Peñailillo F, Tian F. Assessment of the clumped model to estimate olive orchard evapotranspiration using meteorological data and UAV-based thermal infrared imagery. Irrig. Sci. 2021;39:63–80. [Google Scholar]
  • 29.Allen RG, Pereira LS, Raes D, Smith M. Crop Evapotranspiration-Guidelines for Computing Crop Water Requirements-FAO Irrigation and Drainage Paper 56. FAO; 1998. p. 6541. [Google Scholar]
  • 30.Ortega-Farias S, Antonioletti R, Olioso A. Net radiation model evaluation at an hourly time step for mediterranean conditions. Agronomie. 2000;20:157–164. [Google Scholar]
  • 31.Al-Riahi M, Al-Jumaily K, Kamies I. Measurements of net radiation and its components in semi-arid climate of Baghdad. Energy Convers. Manag. 2003;44:509–525. [Google Scholar]
  • 32.Almeida AC, Landsberg JJ. Evaluating methods of estimating global radiation and vapor pressure deficit using a dense network of automatic weather stations in coastal Brazil. Agric. For. Meteorol. 2003;118:237–250. [Google Scholar]
  • 33.Samani Z, Bawazir AS, Bleiweiss M, Skaggs R, Tran VD. Estimating daily net radiation over vegetation canopy through remote sensing and climatic data. J. Irrig. Drain. Eng. 2007;133:291–297. [Google Scholar]
  • 34.Sentelhas PC, Gillespie TJ. Estimating hourly net radiation for leaf wetness duration using the Penman-Monteith equation. Theor. Appl. Climatol. 2008;91:205–215. [Google Scholar]
  • 35.Carrasco M, Ortega-Farías S. Evaluation of a model to simulate net radiation over a Vineyar cv. Cabernet Sauvignon. Chil. J. Agric. Res. 2008;68:156–165. [Google Scholar]
  • 36.Berni JA, Zarco-Tejada PJ, Sepulcre-Cantó G, Fereres E, Villalobos F. Mapping canopy conductance and CWSI in olive orchards using high resolution thermal remote sensing imagery. Remote Sens. Environ. 2009;113:2380–2388. [Google Scholar]
  • 37.Simó G, Martínez-Villagrasa D, Jiménez MA, Caselles V, Cuxart J. Impact of the surface-atmosphere variables on the relation between air and land surface temperatures. In: Vilibić I, Horvath K, Palau JL, editors. Meteorology and Climatology of the Mediterranean and Black Seas. Springer International Publishing; 2019. pp. 219–233. [Google Scholar]
  • 38.Rusticucci MM, Kousky VE. A comparative study of maximum and minimum temperatures over Argentina: NCEP-NCAR reanalysis versus station data. J. Clim. 2002;15:2089–2101. [Google Scholar]
  • 39.Kubota M, Iwabe N, Cronin MF, Tomita H. Surface heat fluxes from the NCEP/NCAR and NCEP/DOE reanalyses at the Kuroshio Extension Observatory buoy site. J. Geophys. Res. Ocean. 2008;113:1–14. [Google Scholar]
  • 40.Mooney PA, Mulligan FJ, Fealy R. Comparison of ERA-40, ERA-Interim and NCEP/NCAR reanalysis data with observed surface air temperatures over Ireland. Int. J. Climatol. 2011;31:545–557. [Google Scholar]
  • 41.Thomas SR, Nicolau S, Martínez-Alvarado O, Drew DJ, Bloomfield HC. How well do atmospheric reanalyses reproduce observed winds in coastal regions of Mexico? Meteorol. Appl. 2021;28:1–13. [Google Scholar]
  • 42.Yu L, et al. Validation of ECMWF and NCEP-NCAR reanalysis data in Antarctica. Adv. Atmos. Sci. 2010;27:1151–1168. [Google Scholar]
  • 43.Raziei T, Parehkar A. Performance evaluation of NCEP/NCAR reanalysis blended with observation-based datasets for estimating reference evapotranspiration across Iran. Theor. Appl. Climatol. 2021;144:885–903. [Google Scholar]
  • 44.de Lima JAG, Alcântara CR. Comparison between ERA Interim/ECMWF, CFSR, NCEP/NCAR reanalysis, and observational datasets over the eastern part of the Brazilian Northeast Region. Theor. Appl. Climatol. 2019;138:2021–2041. [Google Scholar]
  • 45.Yang S-K, Hou Y-T, Miller AJ, Campana KA. Evaluation of the earth radiation budget in NCEP–NCAR reanalysis with ERBE. J. Clim. 1999;12:477–493. [Google Scholar]
  • 46.Iqbal M. An Introduction to Solar Radiation. Academic Press Inc.; 1983. [Google Scholar]
  • 47.Becker S. Calculation of direct solar and diffuse radiation in Israel. Int. J. Climatol. 2001;21:1561–1576. [Google Scholar]
  • 48.Cristóbal J, et al. Estimation of surface energy fluxes in the Arctic tundra using the remote sensing thermal-based Two-Source Energy Balance model. Hydrol. Earth Syst. Sci. 2017;21:1339–1358. [Google Scholar]
  • 49.Wang D, Liang S, He T, Shi Q. Estimating clear-sky all-wave net radiation from combined visible and shortwave infrared (VSWIR) and thermal infrared (TIR) remote sensing data. Remote Sens. Environ. 2015;167:31–39. [Google Scholar]
  • 50.Wang K, Liang S. Estimation of daytime net radiation from shortwave radiation measurements and meteorological observations. J. Appl. Meteorol. Climatol. 2009;48:634–643. [Google Scholar]
  • 51.Bisht G, Venturini V, Islam S, Jiang L. Estimation of the net radiation using MODIS (moderate resolution imaging spectroradiometer) data for clear sky days. Remote Sens. Environ. 2005;97:52–67. [Google Scholar]
  • 52.Prata J. 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]
  • 53.Holtslag AAM, Van Ulden AP. A simple scheme for daytime estimates of the surface fluxes from routine weather data. J. Clim. Appl. Meteorol. 1983;22:517–529. [Google Scholar]
  • 54.Alados I, Foyo-Moreno I, Olmo FJ, Alados-Arboledas L. Relationship between net radiation and solar radiation for semi-arid shrub-land. Agric. For. Meteorol. 2003;116:221–227. [Google Scholar]
  • 55.Cammalleri C, et al. Actual evapotranspiration assessment by means of a coupled energy/hydrologic balance model: Validation over an olive grove by means of scintillometry and measurements of soil water contents. J. Hydrol. 2010;392:70–82. [Google Scholar]
  • 56.Jacobs JM, Anderson MC, Friess LC, Diak GR. Solar radiation, longwave radiation and emergent wetland evapotranspiration estimates from satellite data in Florida, USA. Hydrol. Sci. J. 2004;49:461–476. [Google Scholar]
  • 57.Tegegne EB, et al. Estimation of the distribution of the total net radiative flux from satellite and automatic weather station data in the Upper Blue Nile basin, Ethiopia. Theor. Appl. Climatol. 2021;143:587–602. [Google Scholar]
  • 58.Kalnay E, et al. The NCEP/NCAR 40-year reanalysis project. Bull. Am. Meteorol. Soc. 1996;77:437–471. [Google Scholar]
  • 59.Kistler R, et al. The NCEP-NCAR 50-year reanalysis: Monthly means CD-ROM and documentation. Bull. Am. Meteorol. Soc. 2001;82:247–267. [Google Scholar]
  • 60.Trenberth KE, Fasullo JT, Mackaro J. Atmospheric moisture transports from ocean to land and global energy flows in reanalyses. J. Clim. 2011;24:4907–4924. [Google Scholar]
  • 61.Allen, R. G., Hartogensis, O. & de Bruin, H. A. R. Long-wave radiation over alfafa during the RAPID field campaign in southern Idaho (Research Report, Kimberly, Univ. of Idaho, Id, 2000).
  • 62.Allen RG, 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]
  • 63.Idso SB, Jackson RD, Pinter PJ, Jr, Reginato RJ, Hatfield JL. Normalizing the stress-degree-day parameter for environmental variability. Agric. Meteorol. 1981;24:45–55. [Google Scholar]
  • 64.Brutsaert W. On a derivable formula for long-wave radiation from clear skies. Water Resour. Res. 1975;11:742–744. [Google Scholar]
  • 65.Angstrom A. A study of the radiation of the atmosphere. Smithson. Misc. Collect. 1915;65:159. [Google Scholar]
  • 66.Hijmans RJ, Cameron SE, Parra JL, Jones PG, Jarvis A. Very high resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 2005;25:1965–1978. [Google Scholar]
  • 67.Tetens O. Uber einige meteorologische Begriffe. Z. Geophys. 1930;6:297–309. [Google Scholar]
  • 68.Morales-Salinas L, et al. Monthly calibration of Hargreaves-Samani equation using remote sensing and topoclimatology in central-southern Chile. Int. J. Remote Sens. 2017;38:7497–7513. [Google Scholar]
  • 69.Hanha SR. Air quality model evaluation and uncertainty. JAPCA. 1988;38:406–412. doi: 10.1080/08940630.1988.10466390. [DOI] [PubMed] [Google Scholar]
  • 70.Legates DR, McCabe GJ. Evaluating the use of ‘goodness-of-fit’ measures in hydrologic and hydroclimatic model validation. Water Resour. Res. 1999;35:233–241. [Google Scholar]
  • 71.Deeg J, et al. Extrasolar planet detection by binary stellar eclipse timing: Evidence for a third body around CM Draconis. Astron. Astrophys. 2008;480:563–571. [Google Scholar]
  • 72.Meek DW, Howell TA, Phene CJ. Concordance correlation for model performance assessment: An example with reference evapotranspiration observations. Agron. J. 2009;101:1012–1018. [Google Scholar]
  • 73.Willmott CJ, Robeson SM, Matsuura K. A refined index of model performance. Int. J. Climatol. 2012;32:2088–2094. [Google Scholar]
  • 74.Willmott C, Matsuura K. Advantages of the mean absolute error (MAE) over the root mean square error (RMSE) in assessing average model performance. Clim. Res. 2005;30:79–82. [Google Scholar]
  • 75.Willmott CJ, et al. Statistics for the evaluation and comparison of models. J. Geophys. Res. 1985;90:8995. [Google Scholar]
  • 76.Akaike H. Information theory and an extension of the maximum likelihood principle. In: Csaki BNPF, editor. 2nd International Symposium on Information Theory. Akadémiai Kiadó; 1973. pp. 267–281. [Google Scholar]
  • 77.Burnham KP, Anderson DR. Model Selection and Inference: A Practical Information-Theoretic Approach. Springer-Verlag; 2002. [Google Scholar]
  • 78.Sakamoto Y, Ishiguro M, Kitagawa G. Akaike Information Criterion Statistics. D. Reidel; 1986. p. 26853. [Google Scholar]
  • 79.Gupta HV, Kling H. On typical range, sensitivity, and normalization of Mean Squared Error and Nash-Sutcliffe Efficiency type metrics. Water Resour. Res. 2011;47:2–4. [Google Scholar]
  • 80.Fotheringham AS, Charlton M, Brunsdon C. Measuring spatial variations in relationships with geographically weighted regression BT. In: Fischer MM, Getis A, editors. Recent Developments in Spatial Analysis: Spatial Statistics, Behavioural Modelling, and Computational Intelligence. Springer Berlin Heidelberg; 1997. pp. 60–82. [Google Scholar]
  • 81.Fotheringham AS, Brunsdon C, Charlton M. Quantitative Geography: Perspectives on Spatial Data Analysis. Sage; 2000. [Google Scholar]
  • 82.Fotheringham AS, Brunsdon C, Charlton M. Geographically Weighted Regression: The Analysis of Spatially Varying Relationships. Wiley; 2003. [Google Scholar]
  • 83.R Core Team. R: A language and environment for statistical computing version 4.1.3. https://cran.r-project.org/bin/windows/base/old/4.1.3 (2022).
  • 84.Hijmans, R. raster: Geographic data analysis and modeling (2021).
  • 85.Bivand, R., Keitt, T. & Rowlingson, B. rgdal: Bindings for the ‘Geospatial’ data abstraction library (2021).
  • 86.Carr, D., Lewin-Koh, N. & Maechler, M. hexbin: Hexagonal binning routines (2021).
  • 87.Zambrano-Bigiarini, M. Goodness-of-fit functions for comparison of simulated and observed hydrological time series (2020).
  • 88.Buytaert, W. topmodel: Implementation of the Hydrological Model TOPMODEL in R (2018).
  • 89.Lu B, Harris P, Charlton M, Brunsdon C. The GWmodel R package: Further topics for exploring spatial heterogeneity using geographically weighted models. Geo-spat. Inf. Sci. 2014;17:85–101. [Google Scholar]
  • 90.Gollini, I., Lu, B., Charlton, M., Brunsdon, C. & Harris, P. GWmodel: An R package for exploring spatial heterogeneity using geographically weighted models. https://arXiv.org/arXiv.1306.0413, (2013).
  • 91.Wan, Z., Hook, S. & Hulley, G. MOD11C3 MODIS/terra land surface temperature/emissivity monthly L3 global 0.05Deg CMG V006. NASA EOSDIS land processes DAAC (2015).
  • 92.Tucker CJ. Red and photographic infrared linear combinations for monitoring vegetation. Remote Sens. Environ. 1979;8:127–150. [Google Scholar]
  • 93.Didan, K. MOD13A3 MODIS/terra vegetation indices monthly L3 Global 1 km SIN grid V006. NASA EOSDIS land processes DAAC (2015).

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Data Availability Statement

The datasets used and/or analysed during the current study available from the corresponding author on reasonable request.


Articles from Scientific Reports are provided here courtesy of Nature Publishing Group

RESOURCES