Next Article in Journal
Polar Aerosol Vertical Structures and Characteristics Observed with a High Spectral Resolution Lidar at the ARM NSA Observatory
Previous Article in Journal
Basin-Scale Sea Level Budget from Satellite Altimetry, Satellite Gravimetry, and Argo Data over 2005 to 2019
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Technical Note

Evaluation of Satellite Images and Products for the Estimation of Regional Reference Crop Evapotranspiration in a Valley of the Ecuadorian Andes

by
Fernando Oñate-Valdivieso
*,
Arianna Oñate-Paladines
and
Deiber Núñez
Department of Civil Engineering, Hydrology and Climatology Research Group, Universidad Técnica Particular de Loja, C/. Marcelino Champagnat S/N, Loja 1101608, Ecuador
*
Author to whom correspondence should be addressed.
Remote Sens. 2022, 14(18), 4630; https://doi.org/10.3390/rs14184630
Submission received: 20 July 2022 / Revised: 7 September 2022 / Accepted: 9 September 2022 / Published: 16 September 2022

Abstract

:
The capabilities of images and products obtained by four satellite sensors with different spatial and temporal resolutions (LANDSAT 8, ASTER, MODIS and SENTINEL 3) were analyzed as inputs for the calculation of the reference crop evapotranspiration (ETo). The FAO Penman-Monteith equation was: (a) used in a traditional way, to calculate local values of ETo using data recorded in the existing meteorological stations in the study area; and (b) applied to estimate the regional ETo using the images and products obtained through remote sensing. The capabilities of the satellite products selected were evaluated by means of cross-validation, comparing the values obtained by the meteorological stations to the corresponding values extracted from the regional evapotranspiration maps. The coefficient of determination (R2), the Nash–Sutcliffe efficiency (NSE) and the root mean square error (RMSE) were calculated. The results of the study show that there is a direct correlation between the spatial resolution and the accuracy of the ETo estimates, the products of LANDSAT 8 being those that reached the greatest accuracy. Nevertheless, for continuous ETo monitoring, SENTINEL 3 is preferred.

Graphical Abstract

1. Introduction

The spatial and temporal variation of evapotranspiration has great importance for hydrological, environmental, forestry and agricultural applications. Evapotranspiration (ET) represents the loss of water from a surface through the simultaneous processes of evaporation and transpiration [1]. ET is generally measured using ground-based instruments, such as lysimeters, tank evaporators, or eddy covariance systems. However, ground measurements have limited ability to provide representative values because measurement network densities tend to be low [2], reducing their practical applicability. In contrast, the application of indirect methods based on empirical and semi-empirical equations that use meteorological data present greater relevance. Widely accepted methodologies include those proposed by Thornwaite and Mather; Turc, Blanney and Criddle; and Penman and Monteith [3].
Reference crop evapotranspiration (ETo) is defined as the rate of evapotranspiration from a hypothetical reference crop with an assumed crop height of 0.12 m, a fixed daily canopy resistance of 70 s m−1 and an albedo of 0.23, closely resembling the evaporation from an extensive surface of green grass cover of uniform height, actively growing and adequately watered [4].
ETo depends on the characteristics of the soil and vegetation, so it can present notable spatial variation. However, the direct and indirect methods have limitations in representing this variation because they are not spatially distributed.
ETo is usually calculated by applying the FAO Penman-Monteith equation, which requires information on temperature, solar radiation, wind speed and relative humidity. These variables are usually recorded through meteorological station networks. However, in the Andes, these networks lack sufficient density and the necessary instruments to record all required parameters. To compensate for this, remote sensing strategies for monitoring and studying ETo at different spatial and temporal scales have been used for several decades [5,6,7,8,9,10,11], showing that it is possible to determine ETo in large areas by applying techniques based on remote sensing even where climate data are not available.
Methodologies for calculating ET using remote sensing data can be grouped into the following categories: (a) surface energy balance methods [12,13,14,15], (b) surface temperature and vegetation index methods [16,17,18,19], (c) methods based on the Penman–Monteith equation [20,21,22], (d) methods based on the Priestley–Taylor equation [23,24], (e) empirical methods [25] and (f) water balance-based methods [26,27].
Currently there are different kinds of satellite data, with various spatial, temporal and radiometric resolutions, applicable for the calculation of ETo. Satellite images are an excellent source for constant and frequent observation of the Earth’s surface in near real time [28]. In this work, the images and products obtained by remote sensing were analyzed to evaluate their potential for calculation of the reference crop evapotranspiration in an Andean zone in Ecuador.

2. Materials and Methods

2.1. Study Area

The study area is located in the province of Loja (Ecuador) between the coordinate points 79°13’29.8”W, 03°55´35.9”S and 79°12´08.8”W, 04°08´16.4”S. The city of Loja (240,000 inhabitants) is located in the valley, surrounded by large areas of forest and grassland. The valley is limited by the eastern and western branches of the Andes mountain range. The study area has an altitude range of 2102 to 3400 m [29]. Its climate is sub-humid equatorial temperate, with an average temperature of 16.6 °C and an average annual rainfall of 909 mm, presenting a dry season between May and November and a rainy season between December and April [30]. The location of the study area is presented in Figure 1.

2.2. The Penman–Monteith Equation

The FAO Penman-Monteith equation [4] is the most widely used equation for the estimation of the reference crop evapotranspiration. The FAO Penman-Monteith equation may be expressed as:
E T O = 0.408 Δ R n G + γ 900 T + 273 U 2 e s e a Δ + γ 1 + 0.34 U 2
where ETo is the reference evapotranspiration (mm d−1), T is the daily air temperature at 2 m height (°C), Δ is the slope of the vapor pressure curve (kPa °C−1), γ is the psychrometric constant (kPa °C−1), (esea) is the saturation vapor pressure deficit (kPa), Rn is the net radiation flux on the crop surface (Mj m2 day −1), G is the sensible heat flux into the soil (Mj m2 d−1) and U2 is the wind speed at 2 m above ground (m s−1).
The FAO Penman-Monteith equation is used in this work with two different approaches: (a) to calculate the ETo in a traditional way, using data recorded in the meteorological stations available in the study area and (b) to estimate the regional ETo using images and products obtained by remote sensing. The values calculated at the meteorological stations were used to validate the results obtained using satellite data.

2.3. Information Collected

Three image sets from the LANDSAT 8 Operational Land Imager (OLI) Thermal Infrared Sensor (TIRS), two image sets from the Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) and six product sets from the Moderate Resolution Imaging Spectroradiometer (MODIS) and Land Surface Temperature and Emissivity (MOD11)—Vegetation Indices (MOD13) were collected for the study area from the USGS EarthExplorer server [31]. Additionally, four image sets from the SENTINEL 3 Ocean and Land Color Instrument (OLCI) Sea and Land Surface Temperature Radiometer (SLSTR) were obtained from the ESA Copernicus portal [32]. Their acquisition dates, spectral bands and spatial resolutions are listed in Table 1.
LANDSAT 8 and ASTER images were atmospherically and topographically corrected using the Atmospheric/Topographic Correction for Mountainous Terrain (ATCOR) software developed by the German Aerospace Center, Wessling, Germany [33]. MODIS and SENTINEL 3 data do not require such corrections. All images in Table 1 are geometrically matched to each other.

2.4. Calculation of ETo Using Meteorological Data

There are five Davis Vantage Pro weather stations in the study area. These stations are part of a network owned by Universidad Técnica Particular de Loja (Table 2), which collects information on maximum, minimum and average air temperatures (Ta); relative humidity (HR); wind speed (U2); and shortwave solar radiation (Rs) with a temporal resolution of 10 min.
Equation (1) was applied to calculate the daily reference crop evapotranspiration using the recorded variables. Further details on the calculation of ETo can be found in [4,34].

2.5. Calculation of ETo Using Remote Sensing Data

For LANDSAT 8 and ASTER, the surface temperature (Ts) was calculated by applying the following equation [35]:
T s = T b 1 + λ T b α l n ε
where Ts is the surface temperature (K), Tb is the brightness temperature (K), λ is the effective wavelength of the thermal band (μm), α =   h   c / σ 1.438 × 10 2   mK , h is Planck’s constant (6.626 × 10−34 J s−1), c is the speed of light (2.998 × 108 m s−1), σ is the Stefan Boltzmann constant (1.38 × 10−23 J K−1) and ε is the surface emissivity.
The emissivity was estimated using the following equation [36]:
ε = ε V PV + ε SU ( 1     PV )
where ε is the emissivity of the surface, εV is the emissivity of the vegetation (0.985), εSU is the emissivity of the soil (0.960) and PV is the vegetation proportion, which can be determined by the following equation [37]:
P V =   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 NDVImax and NDVImin are the maximum and minimum values of the normalized difference vegetation index.
Band 9 of the images obtained by the SLSTR sensor was used for SENTINEL 3. In the case of MODIS, the temperature maps were obtained from the MOD11A1 product.
The air temperature (Ta) was calculated using multiple regression equations, applying the methodology proposed by [38]. The independent variables considered were the soil temperature (Ts) (calculated with Equation (2)) and the elevation (z) extracted from a 30 m resolution DEM that was downloaded from the Alaska Satellite Facility Distributed Active Archive Center (ASF DAAC). The Ta values registered in the existing meteorological stations in the study area (Table 2) were considered as dependent variables.
The instantaneous net radiation was calculated by applying the following equation [39]:
R n i = 0.77 R s + 0.98 R l σ T s 4
where Rni is the instantaneous net surface radiation (w m−2), Rs is the incoming shortwave solar radiation (w m−2), Ts is the surface temperature (K) and σ is the Stefan Boltzmann constant (5.67 × 10−8 W m−2 K−4).
To determine the daily net radiation Rnd, the equation from [7] was used:
R n d = C d R n i
C d = 1 α R s d + ε R l d σ T s d 4 1 α R s i + ε R l i σ T s i 4
where Rsd is the daily incoming shortwave solar radiation (W m−2), Rsi is the instantaneous incoming shortwave solar radiation (W m−2), Rld is the daily instantaneous downward longwave radiation (w m−2), Rli is the instantaneous downward longwave radiation (w m−2), Tsd is the daily surface temperature (K), Tsi is the instantaneous surface temperature (K), α is the albedo and ε is the emissivity.
Albedo (α) was determined using the following equation [7]:
α = 1 2 ρ R E D + ρ N I R
where ρ R E D and ρ N I R   are, respectively, the RED and NIR reflectances.
Rsd and Rsi were calculated using information recorded at weather stations. The descending longwave radiation Rld was estimated using the following equation [2,40]:
R l d = 1.24 e a T a 1 7 σ T a 4  
e a = 6.108   e 17.27 T a 273.15 T a 35.85 H R 100
where ea is the actual vapor pressure (HPa), and Ta is the air temperature (K).
The soil heat flux was calculated using the following equation [41]:
G = 0.0038 α + 0.0074 α 2 1 0.98 N D V I 4 T s 273.15 α R n d
Equation (1) was applied to calculate the regional ETo using the results obtained from the remote sensing products described in this section; wind speed, radiation and relative humidity data, recorded at the meteorological stations, were spatialized using the ordinary Kriging interpolation method [29,41] through the use of geographic information systems (GISs). Kriging is a method that originates in geological and mining applications that assumes there is a non-random relationship between the points that changes spatially. The geostatistical interpolation consists in making an optimal prediction of field values in a place where no data is available by using known data to elaborate a field variogram (covariance), thus providing a prediction and its error considering certain weighting factors [42]
The data to be interpolated were adjusted satisfactorily to a normal distribution, observing the existence of a second-order trend. With these considerations, it was decided to perform an ordinary Kriging interpolation without any transformation, with second-order trend removal and adjustment of a spherical semivariogram, which, according to [43], should have been adequate due to the nature of the data.

2.6. Product Evaluation

The results obtained using the data from LANDSAT 8, SENTINEL 3, ASTER and MODIS were evaluated through cross-validation. The ETo was calculated at a regional scale for each selected date, leaving one station at a time out of the calculations. The point value corresponding to the spatial location of the station not considered was extracted to relate it to the one calculated in the traditional way. The coefficient of determination (R2), the Nash–Sutcliffe efficiency (NSE) and the root mean square error (RMSE) [41] were calculated to quantitatively assess the accuracy of the ETo estimates.

3. Results and Discussion

An inverse relationship exists between the elevation above sea level (z) and the air temperature, as can be observed in the regression equations for each date considered (Table 3). On the other hand, a more variable relationship between air temperature and ground temperature (Ts) can also be observed, which was produced by the particular conditions of the moment of observation.
High variability was observed in the ETo values (Table 4), which was attributed to the influence of local conditions caused by the presence of marked altitude differences that lead to differences in temperature, wind speed, incident radiation, etc., causing low uniformity in the ETo values.
Additionally, no seasonal behavior was observed in the study area. This was because it is located on the equator, in an inter-Andean valley influenced by the climatic regime of the Pacific Ocean (wet season from December to April) and the climatic regime of the Amazonia (humid almost all year round), with average temperature fluctuations of no more than 2 °C. This produces uniform climatic conditions to some extent throughout the year [30].
Some examples of the surface temperature calculated using the products obtained by remote sensing are shown in Figure 2. The influence of the spatial resolution of the images can be observed, with the LANDSAT results being the most detailed. On the other hand, the spatial distribution of the temperature shows that the highest values of Ts were at the lowest elevations (in the valley of the study area) where the city of Loja is located. The surrounding areas of the city are covered by vegetation and present a higher albedo than the valley, consequently registering a lower surface temperature. The valley is surrounded by mountains, so the lowest Ts values are recorded in the periphery of the study area.
Similar behavior was observed in the air temperature maps presented in Figure 3, where the distribution of Ta is more uniform. Additionally, the inverse proportionality between the elevation and Ta can be observed.
The spatial variation of the net radiation is presented in Figure 4. The highest values are located in the areas with the highest elevation, which are covered by forests in good condition. The spatial variability of the ETo is high (Figure 5) due to the variable conditions of temperature, net radiation, wind speed, etc.
The ETo values for the selected dates at the existing stations in the study area calculated using remote sensing data (Table 5) are similar to those presented in Table 4. On the dates in which information from two sensors is available, it is possible to see values that have a certain similarity, with the exception of 5 July 2017, in which the values calculated using ASTER and MODIS data present much greater differences. This may have originated in the spatial resolution of the sensors, since MODIS presents larger pixels in contrast to ASTER, which has a better level of detail thanks to its better spatial resolution.
The absolute values of the differences between the ETo calculated with station data and the ETo calculated with remote sensing data are presented in Table 6. On the days in which MODIS data were used, a greater absolute values for the differences were observed, reaching a maximum mean value of 0.61 mm with a standard deviation of 0.42 mm (01/17/17). Differences were observed for the dates in which LANDSAT 8 and ASTER data were used with values of 0.15 mm and deviations of 0.08 mm and 0.16 mm.
The smallest values of the average and standard deviation, were observed in the Villonaco station (Table 7), in which, again, LANDSAT 8 and ASTER showed lower average values—0.02 mm and 0.03 mm, respectively—while the average values obtained by Sentinel 3 and MODIS were significantly higher (0.28 mm and 0.18 mm, respectively). The resolution of each sensor could cause the difference between these values.
The comparison between the remote sensing data ETo and the station data ETo (Figure 6) presented a good fit, especially when data from LANDSAT 8 and ASTER were used. On the other hand, a large dispersion was observed for the results obtained with MODIS; the detail level of the satellite images and products is decisive.
All validation parameters for the ETo estimates (R2, NSE and RMSE) obtained from satellite products were in very satisfactory ranges (Table 8). ASTER and LANDSAT 8 products presented the best results, which can be attributed to their good spatial resolution. This also occurred with the other products, the precision of which decreased as their spatial resolution decreased. The effect of resolution was notable since the radiometric response for the same portion of the territory differs depending on the size of the pixel. This is because, with larger pixels, greater numbers of land covers will be considered to average the information, which causes a decrease in the precision of the ETo estimates.
In contrast to this, the temporal resolutions of SENTINEL 3 and MODIS are significantly higher than those of LANDSAT 8 and ASTER (see Table 1). This implies the availability of a greater number of images. Although the accuracy obtained by SENTINEL 3 and MODIS was lower than that obtained by LANDSAT 8 and ASTER (see Table 8), their resolution is acceptable and, therefore, they constitute an important alternative in cases where permanent monitoring of the ETo is required or in areas characterized by frequent cloud cover.

4. Conclusions

The results show that the precision of the calculations is proportional to the spatial resolution. Higher spatial resolutions reach close-to-optimum values for the coefficient of determination (R2), Nash–Sutcliffe efficiency (NSE) and root mean square error (RMSE). This translates into very satisfactory ETo estimates when using LANDSAT 8 and ASTER products.
The temporal resolutions of MODIS and SENTINEL 3 products is significantly higher than those of the LANDSAT 8 and ASTER products. Therefore, although the precision of ETo estimations using MODIS and SENTINEL 3 is lower than that obtained with LANDSAT 8 and MODIS, they can be used in applications that require continuous monitoring of ETo.
Finally, LANDSAT 8 sensor products result in greater precision in ETo estimations, while SENTINEL 3 products constitute the best alternative if continuous monitoring of ETo is preferred over precision.
Although several of the parameters required by the Penman–Monteith equation can be estimated through remote sensing, data measured at meteorological stations are still required to estimate ETo accurately. This could limit the application of the proposed methodology in places where meteorological networks have little coverage. Future works could focus on the estimation of these parameters based on remote sensing in order to reduce dependence on data measured in the field. Furthermore, exploring the applicability of different methodologies for calculating ETo through remote sensing could be relevant. Finally, the study of the relationship between the regional ETo, the climatic variations and the multi-temporal change of land use/cover could potentially contribute to understandings of the implications of climate change in a given territory, assisting in the definition of possible scenarios and outlining adaptation actions.

Author Contributions

Conceptualization, F.O.-V. and A.O.-P.; methodology, F.O.-V. and A.O.-P.; validation, A.O.-P., D.N. and F.O.-V.; formal analysis, F.O.-V.; investigation, A.O.-P.; resources, D.N.; writing—original draft preparation, F.O.-V., A.O.-P. and D.N.; writing—review and editing, F.O.-V.; visualization, D.N.; project administration, F.O.-V.; All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Walker, E.; García, G.A.; Venturini, V. Estimación de la evapotranspiración real en zonas de llanura mediante productos de humedad de suelo de la misión SMAP. Rev. Teledetección 2018, 52, 17–26. [Google Scholar] [CrossRef]
  2. Sur, C.; Kang, S.; Kim, J.; Choi, M. Remote sensing-based evapotranspiration algorithm: A case study of all sky conditions on a regional scale. GISci. Remote Sens. 2015, 52, 627–642. [Google Scholar] [CrossRef]
  3. Marini, F.; Santamaría, M.; Oricchio, P.; Di Bella, C.M.; Basualdo, A. Estimación de evapotranspiración real (ETR) y de evapotranspiración potencial (ETP) en el sudoeste bonaerense (Argentina) a partir de imágenes MODIS. Rev. Teledetección 2017, 48, 29–41. [Google Scholar] [CrossRef]
  4. Allen, R.; Pereira, L.; Raes, D.; Smith, M. Crop. Evapotranspiration: Guidelines for Computing Crop Water Requeriments; FAO Irrigation and Drainage Paper No. 56; Food and Agriculture Organization of the United Nations: Rome, Italy, 1998; p. 338. [Google Scholar]
  5. Sánchez, M.; Chuvieco, E. Estimación de la evapotranspiración del cultivo de referencia, ET0, a partir de imágenes NOAA-AVHRR. Rev. Teledetección 2000, 14, 11–21. [Google Scholar]
  6. Rivas, R.; Caselles, V. A simplified equation to estimate spatial reference evaporation from remote sensing-based surface temperature and local meteorological data. Remote Sens. Environ. 2004, 93, 68–76. [Google Scholar] [CrossRef]
  7. Sobrino, J.A.; Gómez, M.; Jiménez-Muñoz, J.C.; 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]
  8. Tasumi, M.; Trezza, R.; Allen, R.G.; Wright, J.L. Satellite-Based Energy Balance for Mapping Evapotranspiration with Internalized Calibration (METRIC)-Applications. J. Irrig. Drain. Eng. 2008, 133, 395–406. [Google Scholar]
  9. Venturini, V.; Islam, S.; Rodríguez, L. Estimation of evaporative fraction and evapotranspiration from MODIS products using a complementary based model. Remote Sens. Environ. 2008, 112, 132–141. [Google Scholar] [CrossRef]
  10. Barraza, V.; Restrepo-Coupe, N.; Huete, A.; Grings, F.; Van Gorsel, E. Passive microwave and optical index approaches for estimating surface conductance and evapotranspiration in forest ecosystems. Agric. For. Meteorol. 2015, 213, 126–137. [Google Scholar] [CrossRef]
  11. Knipper, K.; Hogue, T.; Scott, R.; Franz, K. Evapotranspiration estimates derived using multi-platform remote sensing in a semiarid region. Remote Sens. 2017, 9, 184. [Google Scholar] [CrossRef]
  12. Anderson, M.C.; Norman, J.M.; Diak, G.R.; Kustas, W.P.; Mecikalski, J.R. A two-source time-integrated model for estimating surface fluxes using thermal infrared remote sensing. Remote Sens. Environ. 1997, 60, 195–216. [Google Scholar] [CrossRef]
  13. Bastiaanssen, W.G.M.; Menenti, M.; Feddes, R.A.; Holtslag, A.A.M. A remote sensing surface energy balance algorithm for land (SEBAL): 1. Formulation. J. Hydrol. 1998, 212, 198–212. [Google Scholar] [CrossRef]
  14. 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]
  15. Sánchez, J.M.; Scavone, G.; Caselles, V.; Valor, E.; Copertino, V.A.; Telesca, V. Monitoring daily evapotranspiration at a regional scale from Landsat-TM and ETM+data: Application to the Basilicata region. J. Hydrol. 2008, 351, 58–70. [Google Scholar] [CrossRef]
  16. Jiang, L.; Islam, S. Estimation of surface evaporation map over southern Great Plains using remote sensing data. Water Resour. Res. 2001, 37, 329–340. [Google Scholar] [CrossRef]
  17. Nishida, K.; Nemani, R.R.; Glassy, J.M.; Running, S.W. Development of an evapotranspiration index from aqua/MODIS for monitoring surface moisture status. IEEE Trans. Geosci. Remote Sens. 2003, 41, 493–501. [Google Scholar] [CrossRef]
  18. Tang, Q.H.; Peterson, S.; Cuenca, R.H.; Hagimoto, Y.; Lettenmaier, D.P. Satellite-based near- real-time estimation of irrigated crop water consumption. J. Geophys. Res. Atmos. 2009, 114, D05114. [Google Scholar] [CrossRef]
  19. Castañeda-Ibáñez, C.; Martínez-Menes, M.; Pascual-Ramírez, F.; Flores-Magdaleno, H.; Fernández-Reynoso, D.; Esparza-Govea, S. Estimación de coeficientes de cultivo mediante sensores remotos en el distrito de riego río Yaqui, Sonora, México. Agrociencia 2015, 49, 221–232. [Google Scholar]
  20. Mu, Q.; Heinsch, F.A.; Zhao, M.; Running, S.W. Development of a global evapotranspiration algorithm based on MODIS and global meteorology data. Remote Sens. Environ. 2007, 111, 519–536. [Google Scholar] [CrossRef]
  21. Mu, Q.Z.; Zhao, M.S.; Running, S.W. Improvements to a MODIS global terrestrial evapotranspiration algorithm. Remote Sens. Environ. 2011, 115, 1781–1800. [Google Scholar] [CrossRef]
  22. Zhang, K.; Kimball, J.S.; Mu, Q.; Jones, L.A.; Goetz, S.J.; Running, S.W. Satellite based analysis of northern ET trends and associated changes in the regional water balance from 1983 to 2005. J. Hydrol. 2009, 379, 92–110. [Google Scholar] [CrossRef]
  23. Fisher, J.B.; Tu, K.P.; Baldocchi, D.D. Global estimates of the land-atmosphere water flux based on monthly AVHRR and ISLSCP-II data, validated at 16 FLUX- NET sites. Remote Sens. Environ. 2008, 112, 901–919. [Google Scholar] [CrossRef]
  24. Miralles, D.G.; Holmes, T.R.H.; De Jeu, R.A.M.; Gash, J.H.; Meesters, A.G.C.A.; Dolman, A.J. Global land-surface evaporation estimated from satellite- based observations. Hydrol. Earth Syst. Sci. 2011, 15, 453–469. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, K.C.; Wang, P.; Li, Z.Q.; Cribb, M.; Sparrow, M. A simple method to estimate actual evapotranspiration from a combination of net radiation, vegetation index, and temperature. J. Geophys. Res. Atmos. 2007, 112, D15107. [Google Scholar] [CrossRef]
  26. Zeng, Z.Z.; Piao, S.L.; Lin, X.; Yin, G.D.; Peng, S.S.; Ciais, P.; Myneni, R.B. Global evapotranspiration over the past three decades: Estimation based on the water balance equation combined with empirical models. Environ. Res. Lett. 2012, 7, 014026. [Google Scholar] [CrossRef]
  27. Long, D.; Longuevergne, L.; Scanlon, B.R. Uncertainty in evapotranspiration from land surface modeling, remote sensing, and GRACE satellites. Water Resour. Res. 2014, 50, 1131–1151. [Google Scholar] [CrossRef]
  28. Kumar, U.; Rashmi; Chatterjee, C.; Raghuwanshi, N.S. Comparative Evaluation of Simplified Surface Energy Balance Index-Based Actual ET against Lysimeter Data in a Tropical River Basin. Sustainability 2021, 13, 13786. [Google Scholar] [CrossRef]
  29. Oñate-Valdivieso, F.; Oñate-Paladines, A.; Collaguazo, M. Spatiotemporal Dynamics of Soil Impermeability and Its Impact on the Hydrology of An Urban Basin. Land 2022, 11, 250. [Google Scholar] [CrossRef]
  30. Oñate-Valdivieso, F.; Fries, A.; Mendoza, K.; Gonzalez-Jaramillo, V.; Pucha-Cofrep, F.; Rollenbeck, R.; Bendix, J. Temporal and spatial analysis of precipitation patterns in an Andean region of southern Ecuador using LAWR weather radar. Meteorol. Atmos. Phys. 2018, 130, 473–484. [Google Scholar] [CrossRef]
  31. U.S. Geological Survey (USGS). Earthexplorer. Available online: https://earthexplorer.usgs.gov (accessed on 18 November 2021).
  32. Copernicus Open Access Hub. Available online: https://scihub.copernicus.eu/dhus/#/home (accessed on 5 December 2021).
  33. Richter, R.; Schläpfer, D. Atmospheric/Topographic Correction for Satellite Imagery: ATCOR-2/3 User Guide, DLR-IB 565-01/15; German Aerospace Center: Wessling, Germany, 2015. [Google Scholar]
  34. Jensen, M.E.; Burmann, R.D.; Allen, R.G. Evaporation and Irrigation Water Requirements. In ASCE Manual and Reports on Engineering Practice; American Society of Civil Engineers: New York, NY, USA, 2016; Volume 70, p. 360. [Google Scholar]
  35. Li, Y.Y.; Liu, Y.; Ranagalage, M.; Zhang, H.; Zhou, R. Examining land use/land cover change and the summertime surface urban heat island effect in fast-growing greater hefei, China: Implications for sustainable land development. ISPRS Int. J. Geo-Inf. 2020, 9, 568. [Google Scholar] [CrossRef]
  36. Valor, E.; Caselles, V. Mapping land surface emissivity from NDVI: Application to European, African, and South American areas. Remote Sens. Environ. 1996, 57, 167–184. [Google Scholar] [CrossRef]
  37. Meng, X.; Cheng, J.; Zhao, S.; Liu, S.; Yao, Y. Estimating Land Surface Temperature from Landsat-8 Data using the NOAA JPSS Enterprise Algorithm. Remote Sens. 2019, 11, 155. [Google Scholar] [CrossRef]
  38. Ninyerola, M.; Pons, X.; Roure, J.M. A methodological approach of climatological modelling of air temperature and precipitation. Int. J. Climatol. 2000, 20, 1823–1841. [Google Scholar] [CrossRef]
  39. Carmona, F.; Rivas, R.; Kruse, E. Estimating daily net radiation in the FAO Penman-Monteith method. Theor. Appl. Climatol. 2017, 129, 89–95. [Google Scholar] [CrossRef]
  40. 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]
  41. Oñate-Valdivieso, F.; Uchuari, V.; Oñate-Paladines, A. Large-Scale Climate Variability Patterns and Drought: A Case of Study in South—America. Water Resour. Manag. 2020, 34, 2061–2079. [Google Scholar] [CrossRef]
  42. Karavitis, C.A.; Alexandris, S.G.; Tsesmelis, D.E.; Athanasopoulos, G. Application of the Standardized Precipitation Index (SPI) in Greece. Water 2011, 3, 787–805. [Google Scholar] [CrossRef]
  43. Goovaerts, P. Geostatistical approaches for incorporating elevation into the spatial interpolation of rainfall. J. Hydrol. 2000, 228, 113–129. [Google Scholar] [CrossRef]
Figure 1. Location of study area.
Figure 1. Location of study area.
Remotesensing 14 04630 g001
Figure 2. Surface temperature (°C) for the different satellite products.
Figure 2. Surface temperature (°C) for the different satellite products.
Remotesensing 14 04630 g002
Figure 3. Air temperature (°C) for the different satellite products.
Figure 3. Air temperature (°C) for the different satellite products.
Remotesensing 14 04630 g003
Figure 4. Daily net radiation (Rn) (MJ m−2 day−1) for the different products.
Figure 4. Daily net radiation (Rn) (MJ m−2 day−1) for the different products.
Remotesensing 14 04630 g004
Figure 5. Maps of ETo (mm day−1) for the different satellite products.
Figure 5. Maps of ETo (mm day−1) for the different satellite products.
Remotesensing 14 04630 g005
Figure 6. Comparison between remote sensing data ETo and the station data ETo.
Figure 6. Comparison between remote sensing data ETo and the station data ETo.
Remotesensing 14 04630 g006
Table 1. Satellite products considered.
Table 1. Satellite products considered.
ResolutionSatellite Data
LANDSAT 8SENTINEL 3ASTERMODIS.
Radiometric (bits)16-8–1612
Spatial (m)30 (Coastal/Aerosol)
30 (B,G,R,NIR,SWIR)
100 (TIR)
500 (S1-S60)
1000 (S7-S9)
15 (VNIR)
30 (SWIR)
90 (TIR)
250 (B1-2) 500(B3-7) 1000 (B8-36)
Spectral (bands)1111 (SLSTR)
21 (OLCI)
1436
Temporal (days)162 (SLSTR)
1 (OLCI)
161
Acquisition dates7 January 2017
19 August 2017
20 September 2017
7 May 2017
24 June 2017
14 July 2017
20 September 2017
7 May 2017
24 June 2017
17 January 2017
13 February 2017
29 March 17
14 July 2017
14 August 2017
19 September 2017
Table 2. Weather stations in the study area.
Table 2. Weather stations in the study area.
StationLongitudeLatitudeElevation (masl)
Militar79°13′03.1″W03°56′52.6″S2033
Jipiro79°11′23.5″W03°58′15.7″S2218
Tecnico79°14′59.6″W03°59′14.9″S2377
Ventanas79°14′46.0″W04°01′18.9″S2816
Villonaco79°16′09.9″W03°59′10.6″S2952
Table 3. Multiple regression equations to calculate Ta.
Table 3. Multiple regression equations to calculate Ta.
DateSensorEquation to Calculate Ta (°C)
7 January 2017LANDSAT 8Ta = 31.901 − 0.00786z + 0.0893Ts
17 January 2017MODISTa = 79.961 − 0.01546z − 1.1131Ts
13 February 2017MODISTa = 27.239 − 0.00675z + 0.1313Ts
29 March 2017MODISTa = 38.777 − 0.008317z − 0.133Ts
7 May 2017ASTERTa = 28.963 − 0.00585z + 0.01104Ts
SENTINEL 3Ta = 34.6518 − 0.00733z − 0.1604Ts
24 June 2017ASTERTa = 28.991 − 0.00663z + 0.05198Ts
SENTINEL 3Ta = 21.7586 − 0.00625z + 0.3755Ts
14 July 2017SENTINEL 3Ta = 31.7595 − 0.00771z + 0.0356Ts
MODISTa = 30.696 − 0.00723z + 0.0220Ts
14 August 2017MODISTa = 40.953 − 0.00875z − 0.2319Ts
19 August 2017LANDSAT 8Ta = 25.822 − 0.00414z − 0.0000201Ts
19 September 2017MODISTa = 8.785 − 0.004197z − 0.5991Ts
20 September 2017LANDSAT 8Ta = 34.122 − 0.00837z + 0.0667Ts
SENTINEL 3Ta = 31.288 − 0.00783z + 0.1384Ts
Table 4. ETo values for the selected dates at the existing stations in the study area.
Table 4. ETo values for the selected dates at the existing stations in the study area.
DateMilitarJipiroTecnicoVentanasVillonaco
7 January 20172.801.522.000.905.94
17 January 20173.543.774.230.403.15
13 February 20173.863.723.084.073.55
29 March 20172.463.002.941.983.05
7 May 20171.912.282.931.744.56
24 June 20172.743.232.442.975.15
14 July 20172.063.213.313.204.86
14 August 20173.453.633.493.364.85
19 August 20172.612.903.073.313.07
19 September 20173.734.003.792.802.75
20 September 20173.844.704.093.593.43
Table 5. ETo values for the selected dates at the existing stations in the study area calculated using remote sensing data.
Table 5. ETo values for the selected dates at the existing stations in the study area calculated using remote sensing data.
DateSensorMilitarJipiroTecnicoVentanasVillonaco
7 January 2017LANDSAT2.561.672.201.005.99
17 January 2017MODIS2.783.113.060.372.70
13 February 2017MODIS3.453.372.473.372.96
29 March 2017MODIS2.252.882.441.753.05
7 May 2017ASTER2.012.022.041.564.53
SENTINEL 31.991.502.541.354.21
24 June 2017ASTER2.653.152.003.095.19
SENTINEL 32.872.952.282.734.68
14 July 2017SENTINEL 32.853.123.113.094.56
MODIS2.273.352.963.014.86
14 August 2017MODIS3.373.653.013.034.80
19 August 2017LANDSAT2.332.802.633.323.08
19 September 2017MODIS3.243.532.952.262.75
20 September 2017LANDSAT3.464.613.523.493.57
SENTINEL 33.654.233.503.103.43
Table 6. Absolute values of differences between ETo calculated with station data and ETo calculated with remote sensing data.
Table 6. Absolute values of differences between ETo calculated with station data and ETo calculated with remote sensing data.
DateSensorMilitarJipiroTecnicoVentanasVillonacoAVGSTD
7 January 2017LANDSAT 80.240.150.200.100.050.150.08
17 January 2017MODIS0.760.661.170.030.450.610.42
13 February 2017MODIS0.410.350.610.700.590.530.15
29 March 2017MODIS0.210.120.500.230.000.210.18
7 May 2017ASTER0.100.260.890.180.030.290.35
SENTINEL 30.080.780.390.390.350.400.25
24 June 2017ASTER0.090.080.440.120.040.150.16
SENTINEL 30.130.280.160.240.470.260.13
14 July 2017SENTINEL 30.790.090.200.110.300.300.29
MODIS0.210.140.350.190.000.180.13
14 August 2017MODIS0.080.020.480.330.050.190.20
19 August 2017LANDSAT 80.280.100.440.010.010.170.19
19 September 2017MODIS0.490.470.840.54-0.590.17
20 September 2017LANDSAT 80.380.090.570.10-0.290.23
SENTINEL 30.190.470.590.49-0.440.17
AVG 0.300.270.520.250.20
STD 0.230.050.080.040.05
Table 7. Average and standard deviation of the absolute values of the differences between ETo calculated with station data and ETo calculated with remotely sensed data (Table 6), by sensor and station.
Table 7. Average and standard deviation of the absolute values of the differences between ETo calculated with station data and ETo calculated with remotely sensed data (Table 6), by sensor and station.
SensorParameterMilitarJipiroTecnicoVentanasVillonaco
LANDSAT 8Average0.300.110.400.070.02
STD0.070.030.190.050.03
ASTERAverage0.100.170.670.150.03
STD0.010.130.320.040.01
SENTINEL 3Average0.300.410.340.310.28
STD0.330.290.200.170.09
MODISAverage0.360.290.660.340.18
STD0.250.240.300.250.28
Table 8. Validation parameters.
Table 8. Validation parameters.
SensorNSER2RMSE
LANDSAT 80.9940.9680.060
ASTER0.9880.9380.067
SENTINEL 30.9850.8940.110
MODIS0.9790.8650.160
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Oñate-Valdivieso, F.; Oñate-Paladines, A.; Núñez, D. Evaluation of Satellite Images and Products for the Estimation of Regional Reference Crop Evapotranspiration in a Valley of the Ecuadorian Andes. Remote Sens. 2022, 14, 4630. https://doi.org/10.3390/rs14184630

AMA Style

Oñate-Valdivieso F, Oñate-Paladines A, Núñez D. Evaluation of Satellite Images and Products for the Estimation of Regional Reference Crop Evapotranspiration in a Valley of the Ecuadorian Andes. Remote Sensing. 2022; 14(18):4630. https://doi.org/10.3390/rs14184630

Chicago/Turabian Style

Oñate-Valdivieso, Fernando, Arianna Oñate-Paladines, and Deiber Núñez. 2022. "Evaluation of Satellite Images and Products for the Estimation of Regional Reference Crop Evapotranspiration in a Valley of the Ecuadorian Andes" Remote Sensing 14, no. 18: 4630. https://doi.org/10.3390/rs14184630

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