Next Article in Journal
Significant Increase in African Water Vapor over 2001–2020
Previous Article in Journal
Improving Soybean Gross Primary Productivity Modeling Using Solar-Induced Chlorophyll Fluorescence and the Photochemical Reflectance Index by Accounting for the Clearness Index
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Calculating Vegetation Index-Based Crop Coefficients for Alfalfa in the Mesilla Valley, New Mexico Using Harmonized Landsat Sentinel-2 (HLS) Data and Eddy Covariance Flux Tower Data

1
New Mexico Water Resources Research Institute, New Mexico State University, Las Cruces, NM 88003-8001, USA
2
Department of Civil Engineering, New Mexico State University, Las Cruces, NM 88003-8001, USA
3
Department of Geography and Environmental Studies, New Mexico State University, Las Cruces, NM 88003-8001, USA
4
USDA SW Climate Hub US Department of Agriculture, Agricultural Research Service, Jornada Experimental Range, Las Cruces, NM 88003-8001, USA
*
Author to whom correspondence should be addressed.
Remote Sens. 2024, 16(16), 2876; https://doi.org/10.3390/rs16162876
Submission received: 25 June 2024 / Revised: 26 July 2024 / Accepted: 1 August 2024 / Published: 6 August 2024
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)

Abstract

The goal of this study is to investigate the usefulness of the relatively new 30 m spatial and <5.7-day temporal resolution Harmonized Landsat Sentinel-2 (HLS) dataset for calculating vegetation index-based crop coefficients (KcVI) for estimating field scale crop evapotranspiration (ETc). Increased spatial and temporal resolution ETc estimates are needed for improving irrigation scheduling, monitoring impacts of water conservation programs, and improving crop yield. The crop coefficient (Kc) method is widely used for estimating ETc. Remote sensing vegetation indices (VI) are highly correlated to Kc and allow the creation of a KcVI but the approach is limited by the availability of high temporal and spatial resolutions. We selected and calculated sixteen commonly used VIs using HLS data and regressed them against field-measured ET for alfalfa in the Mesilla Valley, New Mexico to create linear KcVI models. All models showed good agreement with Kc (r2 > 0.67 and RMSE < 0.15). ETc prediction resulted in an MAE ranging between 0.35- and 0.64-mm day−1, an MSE ranging between 0.20- and 0.75-mm day−1 and an MAPD ranging between 10.0 and 16.5%. The largest differences in predicted ETc occurred early in the growing season and during cutting periods when the spectral signal could be influenced by soil background or irrigation events. The results suggest that applying the KcVI approach to the HLS dataset can help fill in the data gap in remote sensing ET tools. Future work should focus on assessing additional crops and integration into other tools such as the emerging OpenET platform.

1. Introduction

Accurate estimates of evapotranspiration (ET) at both regional and field scales are crucial for gaining insights into the spatiotemporal distribution of agricultural water usage. They are particularly important for regions that suffer from chronic drought, as they can inform water management decisions. Furthermore, these measurements can provide essential context for guiding precision agriculture. In this regard, locally calibrated ET measurements play a vital role in enhancing our understanding of agricultural water use and its impact on the environment.
Remote sensing models for field scale estimates of ET tend to use a physical approach that employs thermal imagery to calculate a surface energy balance to solve for latent energy, which is then converted to evapotranspiration, as a residual in the energy balance (e.g., SEBAL [1], ALEXI/DisALEXI [2], METRIC [3], SSEBop [4]). However, the availability of high spatial (<30 m) and temporal resolution (<8 days) thermal imagery is limited, thus potentially missing several weeks of measurements during the growing period. Conversely, data harmonization of spectral bands between satellites provides high spatial resolution imagery delivered <5.7 days, thus creating new opportunities for the use of empirical methods that utilize the spectral bands for creating vegetation index (VI)-based crop coefficients (KcVI) and provides water managers another tool for decision making. The objective of this research is to better understand the utility of KcVI from the Harmonized Landsat Sentinel (HLS) dataset for estimating ETc at the field scale.
Obtaining measurements of ET across space and time is challenging and field measurements using lysimeters, Bowen ratio, and eddy covariance techniques, or scintillometers, are expensive and require well-trained research personnel who can exploit the demanding accuracy of the measurements. Crop evapotranspiration (ETc) can be estimated using the Kc method, which multiplies a locally calculated standardized reference ET (ETsz) value by an empirically determined Kc value (Equation (1))—the resultant equation is the standardized FAO-56 crop coefficient method [5] (Equation (2)), which allows for the comparison of ETc for different time periods and regions.
K c = E T a E T s z
E T c = E T s z × K c
where
ETsz is reference evapotranspiration (short grass) (mm day−1);
Kc is the crop coefficient (unitless);
ETa is the measured actual evapotranspiration (mm day−1);
ETc is the crop evapotranspiration (mm day−1);
One of the strengths of using the Kc approach is the inclusion of measurements that capture localized weather and plant-specific conditions. ETsz incorporates weather parameters supplied by a nearby weather station to capture the evaporation power of the atmosphere. Kc is a single value representing soil evaporation and plant transpiration [5] and integrates site-specific information such as the farm irrigation and management practices [6] and physiological characteristics of a crop such as light absorption and canopy roughness during different stages of the growth cycle [7]. Kc values are derived from field experiments where the ETa for a specified crop is measured under optimal growing conditions using a weighing lysimeter, soil water balance approach, or eddy covariance method and divided by ETsz (Equation (2)). The FAO-56 report published general Kc values for many crops for three different stages of the growing season. Applying the Kc method can potentially overestimate ETc because of uncertainties in precipitation effectiveness [8], and because Kc values are developed under ideal growing conditions, have morphological and physiological differences for different varieties, thus it is important to modify Kc for local conditions and cultural practices [9]. The FAO method does suggest using a crop stress coefficient (Ks) to adjust for suboptimal growing conditions; however, it requires information on the soil matric potential, hydraulic conductivity, soil type, and the associated field capacity. The published Kc values in FAO-56 do not capture the spatiotemporal variability of Kc at resolutions sufficient for precision agricultural management decisions. In contrast, the sensitivity of the spectral bands of satellite imagery to changes in vegetation health makes remote sensing a viable solution for relating vegetation index (VI) values to Kc and estimating plant ET.
Satellite remote sensing data are widely accepted as a solution for estimating the spatiotemporal distribution of various biophysical attributes of the Earth’s surface, including evapotranspiration. Satellite data are useful because of the availability of free imagery, long periods of historic coverage, and due to most satellite data often having a high level of systematic calibration applied before being distributed [10]. The relationship between electromagnetic radiation (EMR) and vegetation provides the theoretical foundation for remote sensing as a method for measuring ET, where the physical and physiological properties of the plants determine the amount of EMR absorbed, transmitted, reflected, and emitted [11]. In healthy vegetation, EMR wavelengths between 380 and 710 nm, referred to as photosynthetically active radiation, are absorbed by pigments in the chloroplasts, whereas near-infrared wavelengths between 700 and 1200 nm are reflected or transmitted by the spongy mesophyll [12]. As plants decrease chlorophyll production and/or are under stress, there is a reduction in absorption of photosynthetically active radiation [12] which leads to the plants using less water. The relationship between changes in these wavelengths and vegetation characteristics can be realized through the use of VIs.
VIs are dimensionless transformations of measurements of specific wavelengths of electromagnetic energy, typically in the visible, near-infrared, and mid-infrared wavelengths, and are used to highlight features of interest and contain up to 90 percent of spectral information of vegetation [13]. VIs correlate to changes in vegetation health (e.g., water uptake, diseases, nutrients, and pests) and are representative of plant productivity [7]. Because of the similarities to Kc in terms of changes in response to the plant growth cycle and plant stress, KcVI values can be estimated through regression of a VI derived from reflectance data from an in situ radiometer [9] or from satellite and aerial sensors (e.g., [14,15,16]) to ground measured data. Many factors can influence the reflectance values recorded at the sensor including the range of wavelengths the sensor is recording, atmospheric conditions, background soil reflectance, soil moisture, and vegetation disease or stress.
Numerous VIs were developed for different applications. The most commonly applied index is the Normalized Difference Vegetation Index (NDVI) [17]. The NDVI tends to saturate around 0.8 with healthy vegetation and is sensitive to differences in soil background [18]. Several indices were created as variations of the NDVI to incorporate adjustment factors to create a more dynamic response, such as the Renormalized Difference Vegetation Index (RDVI) [19], or to adjust for the soil background effects, such as the Modified Soil-Adjusted Vegetation Index (MSAVI) [20]. Some indices were developed to use only information within the visible spectrum to estimate vegetation conditions such as the Normalized Green Red Difference Index (NGRDI) [21] or the Green-NDVI (GNDVI) [22] and are sensitive to chlorophyll concentration and can have a higher dynamic range compared to the NDVI. Other indices incorporate information from the mid-infrared range because of the response to moisture in vegetation [23] and the ability to indicate drought [24]. While examples exist [25], there are a limited number of studies that compare the effectiveness of multiple VIs for the purpose of calculating Kc. Thus, there is still a need for investigating different VIs for Kc development because the different band combinations and adjustments within the equations provide the potential to improve the dynamic range of the index and reduce background noise, leading to a better representation of Kc. There is also an importance in calculating KcVI with minimal adjustments based on ad hoc manual measurements if these methods are to be operationalized by the agricultural community.
Until recently, a lag in processing, inadequate temporal resolution, and accessibility of data have hindered the operational use of high spatial resolution satellite ET measurements for field scale management decisions. Forage crops such as alfalfa are cut numerous times throughout a growing season and satellites like Landsat may not coincide to capture the average growing conditions [26]. One advancement is the HLS dataset, which capitalizes on the spectral and spatial resolution similarities between Landsat and Sentinel-2 satellites. The result is a dataset that increases the theoretical temporal resolution of data capable of field scale measurements to between 3.2 and 5.7 days [27]. HLS data are available on the internet and have near real-time processes of the most current imagery. The improved temporal resolution of the HLS data makes it an appealing potential solution for use for a KcVI approach for field scale ET measurements.
This study assesses the usefulness of the HLS dataset for creating empirically derived KcVI values and subsequently predicting ETc for an alfalfa (Medicago sativa) crop in the Mesilla Valley of the Rio Grande basin in New Mexico, USA. The objectives of this study are: (1) to create linear regression models from 16 different VIs, and (2) to test the performance of these equations in predicting ETc using regression error metrics.

2. Description of the Study Site and Methodology

2.1. Site Description and Alfalfa Crop

The study site is a 35.2 ha privately owned alfalfa field (32°12′3.07″N, 106°44′09.90″W, central coordinates) located in the Mesilla Valley of southern New Mexico (Figure 1). The Mesilla Valley lies within the Chihuahua Desert and exemplifies agricultural conditions in a semi-arid desert environment. The region’s historical precipitation is less than 230 mm year−1 and 75% of it falls during summer monsoonal months of July, August and September, and the average monthly temperature ranges between 5.5 and 26.6 °C with diurnal temperature of 17.8 °C [28]. The winds in the Valley are moderately strong with occasional gusts in winter and spring and usually light during summer. The potential evaporation of the region exceeds the precipitation, and from 1980 to 2020, the region has experienced an increase of between 135 and 235 mm in evaporative demand [29].
Mesilla Valley is 80 km long and varies in width from a few kilometers to 10 km wide. It is a highly productive agricultural corridor along the Rio Grande in Doña Ana County, New Mexico. Doña Ana County has an area of about 21,400 hectares dedicated to farming and is one of the most productive agricultural regions in New Mexico. The valley supports a variety of forage crops (e.g., alfalfa and winter wheat) and annual row crops (e.g., green chile, onions, and cotton). Over the past few decades, however, the Valley has evolved into one of the largest pecan-producing areas in the United States [30].
The field was laser-leveled and alfalfa was seeded in 2014. The alfalfa field was flood-irrigated 11 times during the 2017 growing season at a 24-day mean interval and received a total of 1874 mm of irrigation and 232 mm of precipitation (Table 1). In 2017, the alfalfa was healthy, uniform, and monotypic. It was harvested seven times during the growing season (Table 2). The average harvest interval for the first five harvests was 12 days and approximately 30 days for the final end-of-season harvest.
The growing season for alfalfa at the study site was determined using growing degree days (GDD) at a base temperature of 5 °C (41 °F) to mark the beginning of the growing season. Accumulation of GDD began on 21 March 2017, or day of the year (DOY 80), when the air temperature reached and stayed above 5 °C (41 °F) for five consecutive days [31]. The end of the growing season was determined to be 7 December 2017 (DOY 341), where the minimum temperature dropped below −4.4 °C (24 °F) for more than two hours [31].
The soil texture at the study site was classified by Boyko et al. [32] as silt loam in the top 30 cm, sandy loam below it up to 60 cm, and then sand up to 180 cm. In 2017, Boyko et al. [32] measured volumetric soil moisture in the upper 30 cm ranging between 20.9% and 45.6% with an average of 40%.

2.2. Evapotranspiration Using the Eddy Covariance Method

All field measurements used in this study are from the previously published work by Boyko, et al. [32]. Climate data for calculating reference evapotranspiration (ETsz) were collected from the Leyendecker III weather station (32°12′21.46″N, 106°44′50.74″W, elevation 1176 m AMSL). The weather station is located approximately 620 m west of the alfalfa field. Data recorded during site maintenance were removed from daily ETsz calculations. The American Society for Civil Engineers Standardized Reference Evapotranspiration [33] was used to calculate ETsz because it is widely used and accepted and provides the ability to transfer crop coefficients to other geographic locations.
Actual evapotranspiration (ETa) was measured by applying the energy balance method to data collected from an eddy covariance flux tower installed in the center of the alfalfa field (Figure 1). Bawazir, et al. [34] present the methodology used for measuring net radiation, soil heat, and sensible heat fluxes and determining latent heat as a residual of the energy balance. Net radiation (Rn) was measured using a NR-Lite (CSI, Logan, UT, USA). Soil heat flux (G) was measured using two HFT3 plates (CSI, Logan, UT, USA) in combination with a CS 616 soil moisture sensor (CSI, Logan, UT, USA) and two CAV averaging soil temperature thermocouples (CSI, Logan, UT, USA). Sensible heat (H) was measured using a CSAT3 (CSI, Logan, UT, USA) three-dimensional sonic anemometer. The equivalent depth of water in mm was calculated by dividing the latent heat flux by the latent heat of vaporization of water (2.45 MJ/kg at 20 °C) and the density of water (1000 kg/m3). An open path LICOR 7500 (LI-COR® Biosciences, Lincoln, NE, USA) was used to directly measure latent heat from May to August as a check for energy balance closure. The fetch distance of the eddy covariance ranged from 2 m to 234 m. The mean energy balance closure was 0.96 for the growing period. See Boyko, et al. [32] for a more detailed description of field measurements and methods carried out.

2.3. Satellite Data

The Harmonized Landsat Sentinel-2 (HLS) dataset is produced to create a virtual constellation of satellites to increase temporal coverage for better monitoring of agricultural and water management among other land monitoring applications [27]. Native Sentinel-2 data have a spatial resolution of 10–20 m and Landsat has a spatial resolution of 30 m. Co-registration of the two satellites is obtained by resampling HLS Sentinel-2 data to 30 m and resampling Landsat data to a common reference grid [27]. The two satellites have similar spectral bands in regard to center wavelengths and bandwidths (Table 3), and both satellites have 12-bit radiometric resolution. The slight differences between satellites require the data to be harmonized and delivered as surface reflectance products so that changes in the spectral responses are due to actual changes in surface reflectance and not a result of the differences between sensors and/or differences in atmospheric conditions. We utilized the HLS data version 2.0 satellite scenes which used Landsat OLI and Sentinel-2 data for the 2017 growing season. These data were acquired through the National Aeronautics and Space Administration (NASA) https://hls.gsfc.nasa.gov/ accessed on 4 January 2023.
Selection criteria were applied within the NASA website to download satellite images for use in the analysis. The selection criteria for the satellite images were acquired during the growing season between 21 March 2017 (DOY 80) and 7 December 2017 (DOY 341), cloud cover of less than 20 percent, and in path/row 33/38 for Landsat and tile T13SCR for Sentinel-2. Further sorting of imagery was performed by visual analysis and assessment of the Fmask band to ensure no clouds were over the study site. After sorting, 26 images (10 Landsat and 16 Sentinel-2) were determined to be useful for analyzing vegetation indices and alfalfa crop coefficients in the Mesilla Valley for the 2017 growing season (Appendix A). The pixels in all but four of the images used in the analysis were assessed to have low or moderate aerosol levels using the Fmask. The average days between satellite passes was 9.1 days and the minimum and maximum days were 1 day and 20 days.

2.4. Spectral Indices

A systematic selection process was used to identify potentially useful vegetation and water indices from the hundreds of spectral indices in use for various earth monitoring applications. A summary of different vegetation indices, their potential uses, advantages, and disadvantages can be found in [35,36,37]. A table of 118 vegetation indices provided by Xue and Su [35] were assessed for inclusion in this study based on whether (1) the spectral bands are available on both HLS satellites; (2) the index does not require manual input, particularly in regards to correction coefficients; and (3) they have some foundation in spectroscopy. From the list of 118 vegetation indices, we selected 16 indices (Table 4). Of the indices that were excluded, 56 used spectral bands not included on Landsat satellites, at least 8 required some form of posteriori input, several indices were duplicated with different names, and some were not peer-reviewed or well-grounded in spectroscopy. It should be noted that several of the citations in [35] were incorrect or did not point to the original source of the indices. Of the 16 selected indices, seven used variations of the relationship between red and near-infrared, four incorporated shortwave infrared wavelengths, three indices used only the bands within the visible spectrum, and two used the relationship between green and near-infrared wavelengths.

2.5. Vegetation Indices-Based Crop Coefficients

Each of the 26 individual HLS images suitable for analysis was processed into 16 vegetation indices and then processed into KcVI values. An area of interest (AOI) of the entire field was digitized into a shapefile using ArcMap version 10.8.1 and 30 cm base map imagery acquired by MAXAR (Westminster, CO, USA) on 14 November 2021 and provided as service layer by ESRI. The shapefile was reduced by a 30 m buffer around the perimeter to reduce the potential influence of edge effects on the spectral indices. Given the uniform irrigation, growth of the field, and cutting cycle, information derived from the imagery pixel values within the AOI was assumed to be representative of the field conditions. Python programming was used to extract zonal statistics of the alfalfa AOI from each vegetation index file including the count, minimum, maximum, mean, and median pixel values. A visual analysis of the histograms revealed the distribution of pixel values was skewed to the left, thus the median value was the better measure of central tendency and used in the subsequent analyses. The median pixel values of each VI from the study field were regressed using an ordinary least squares linear regression against the in situ derived Kc values (i.e., ETa/ETsz) to develop equations for predicting Kc from each VI referred to here as KcVI. The sixteen linear equations were then applied to their respective vegetation index images and then multiplied by the daily ETsz estimate to calculate ETc for each image.
The general process for the study is outlined in Figure 2.

2.6. Statistical Analysis

Adapting the methodology in Hunsaker, et al. [9], common regression metrics were used to assess how well each linear equation predicted the Kc value. The coefficient of determination (r2) was used to measure the proportion in the Kc predicted by each regression equation. The root means square error (RMSE) was used to assess the Euclidean distance from the predicted Kc values to the field-measured Kc values, the lower values indicating lower error. p-values were used to indicate statistical significance of each regression equation. An F-test was used to determine if there was a statistical difference between linear equations produced separately using only Landsat and only Sentinel data for each VI.
The robustness of each linear regression model was assessed using two methods: (1) determining if there is a statistical difference between predicted and measured ETc using a Wilcoxon rank-sum test, and (2) a cumulative ranking of the KcVI model performance in terms of r2. Ten different model runs were performed on the 16 different indices for a total of 160 model outputs with the data partitioned into a 60/40 split of training and testing data using a random selection function in Python—each model run was made up of different sets of randomly selected training and testing data. The Wilcoxon rank-sum test was used to test for statistically significant differences assuming non-parametric data with a 0.05 significance level. The null hypothesis (Ho) assumed no statistical differences between the field-measured and predicted ETc and the alternative hypothesis (Ha) that there is a significant difference. The r2 values for each KcVI regression equation were assigned an ordinal ranking from 1 to 16 for each of the 10 model runs with the higher ranking assigned to the higher r2 values, and were subsequently summed to determine a ranking of which models consistently performed better with the data used in this study.
The relationship between ETc from the eddy covariance tower and the estimated ETc using the spectral indices and the KcVI method were assessed using correlation statistics. The mean absolute error (MAE), the mean squared error (MSE), the mean absolute percent difference (MAPD) relative to the observed mean, and relative root mean squared error (rRMSE) were determined. The MAE assesses the magnitude of error in ETc by using the KcVI equations when comparing the measured ETa, the MSE measures the average squared differences between observed and predicted ET values, the MAPD provides an average of differences between predicted and observed values divided by the observed values, and rRMSE provides a measure of prediction accuracy relative to the range of values for a variable.

2.7. Comparison to OpenET

The agreement between the estimated ETc from the 16 KcVI models and estimated ETa from the OpenET Ensemble model was assessed for 20 additional alfalfa fields located in the Mesilla Valley to address the lack of additional field measurements for comparison. The emerging OpenET platform provides daily, monthly, and annual estimates of ETa for individual field polygons and at a 30 m spatial resolution grid from six remote sensing ET models and a combined Ensemble model [46]. Accuracy assessments of the OpenET models indicate the Ensemble model typically provides a high agreement with eddy covariance data for all croplands (r2 = 0.81, MBE = −0.35 mm day−1) [47], alfalfa (r2 = 0.65, MBE = −12.2 mm month−1) [48], and pecan (r2 = 0.95, MBE = 7 mm month−1) [49]. An additional 20 alfalfa fields were identified using the 2017 USDA Cropland Data Layer and selected to represent the variability in area and distribution throughout the valley. We applied a negative 30 m buffer to the fields to reduce the effects of edge pixel influence on the estimated values. The estimated daily ETa Ensemble model values for each of the 20 alfalfa fields were retrieved through the OpenET Python API.
The 16 KcVI equations were applied to the HLS data for each of the 20 fields to estimate the Kc values and were subsequently multiplied by the corresponding daily ETsz from the Leyendecker III weather station to estimate daily ETc. The data from all 20 fields were combined for each of the 16 KcVI models and then assessed for agreement with the OpenET Ensemble model in terms of r2, MAE, MSE, and MAPD.

3. Results and Discussion

3.1. Regression Equations for KcVI

The linear regressions between the VIs and the Kc values from the in situ field measurements suggest that all of the VIs selected for this study have high correlation coefficients (r2 between 0.67 and 0.90) for the 2017 growing season and predictions of Kc fall within 0.09 to 0.17 RMSE (Table 5). The slope and intercept for the NDVI (1.48 and −0.12) are similar to the results found in the literature. For example, Kamble, et al. [15] estimated Kc for rainfed and irrigated agriculture as 1.4571 × NDVI − 0.1725 and obtained an r2 of 0.91. The linear regression equations derived by Campos, et al. [14] for estimating the basal crop coefficient (Kcb) used the soil-adjusted vegetation index (SAVI) for maize (Kcb = 1.414 × SAVI − 0.02) and soybeans (Kcb = 1.258 × SAVI − 0.006) with r2 values between 0.805- and 0.868. The two best-performing Kc MAE ranged between 0.35- and 0.64-mm day−1, MSE ranged between 0.20- and 0.75-mm day−1, and MAPD ranged 10.0- and 16.5%. The visible atmospherically resistant index (VARI), the normalized green-red difference index (NGRDI), and the normalized difference vegetation index (NDVI) showed the highest correlation and lowest errors when compared to the measured values. Out of the seven VIs that exploit the slope of reflectance between red and near-infrared, the NDVI and the TDVI both had an r2 of 0.88 and the other four equations were less than 0.76. Some of the lowest correlations were found in VIs that corrected for atmospheric and soil background (ARVI, EVI, and MSAVI). The results of an F-test between linear models created separately with Landsat and Sentinel data revealed no significant differences between the linear models of each respective VI, thus suggesting good harmonization between the two sensors (Table A2).
The distribution of field-measured Kc was clustered into low values directly after harvesting (~0.3 to 0.5) and higher values when the alfalfa plants were fully developed (>1.0) during a period typically less than seven days after cutting (Figure 3a). Even with HLS providing a hypothetical return interval of <5 days, there were only a few opportunities during the growing season to capture imagery during early season growth and in the period between cut and canopy closure, and even fewer opportunities due to clouds. This resulted in the inclusion of only seven satellite images during the cut and rapid growth period (Figure 3b). Furthermore, the distribution can be seen on plots of VIs and measured Kc, where the data appear to have two clusters—one related to data acquired during the cut period and another where the alfalfa canopy has closed (Figure 4). Only one data point falls between these clusters.

3.2. Comparisons of Estimated and Measured ETc

The Wilcoxon signed-rank test results of different combinations of training and testing data results suggest the linear model ETc predictions have a closer agreement with the eddy covariance data when data represent all periods of the growth cycle. The 10 model runs of randomized data produced 160 linear models. Of the 160 models, 26 were statistically significantly different (p < 0.05) when compared to measured ETa (Table 6). The model runs that indicated a significant difference were caused by a random selection in the training data of either a majority of low index values (low vegetation cover) or conversely, not including enough low index values. Three models (ExG, NGRDI, and VARI) were more sensitive to changes in training data, which suggests that these models may require larger training datasets. None of the model runs were significantly different between the GNDVI and the field-measured Kc.

3.3. Differences between ET Measurements Based on Vegetation Indices

Modeled results were compared to daily ETa measured from the eddy covariance system to assess the MAE, MSE, and MAPD (Table 7). The MAE was between 0.35- and 0.64-mm day−1 for all VIs, which suggests the use of any of these models for longer time periods of analysis (e.g., monthly or annual) would be appropriate. However, the MAE metric averages out the peaks in higher and lower ET predictions on the daily time scale. This is particularly important to consider during the early growth and cut periods when ET is much lower and the percentages of those errors relative to the measured values can be much higher. The MSE was between 0.20- and 0.75-mm day−1 for all models with the VARI, NGRDI, and NDVI models having better accuracy than the other models. The results of the MAPD suggest the VARI, NGRDI, and NDVI models produce the smallest absolute percent difference from the observed values (<12.0%) on average. Similar results were found by French, et al. [16], where the NDVI was used to calculate Kc for wheat crops in Arizona. French et al. found a MAPD of 28 to 41 percent during the early growth and between 13 and 18 percent during mid-season growth for wheat. The rRMSE was <10% for the NGRDI and the VARI, which are generally considered excellent models; all other models were <20%, which are also considered good models [50,51].
The autocorrelation of the field-derived Kc values used in the VI regression models was based on the ratio of eddy covariance ETa and the weather station ETsz was not an issue impacting the KcVI models. The results of the Wilcoxon signed-rank test, where the data were partitioned into training and testing sets, suggest that there were only significant differences between measured ETa and KcVI predicted ETc when image dates did not capture the early growth stages. Therefore, we argue these comparisons are still accurate, but may not be precise. Future work would benefit from having multiple years of data from different locations.
The difference between higher and lower estimation for the ETa values when compared to eddy covariance measurements was ±2 mm day−1 for all models (Figure 5). For the NDVI, NGRDI, VARI, and VDVI, the predicted values were within approximately ± 1 mm day−1. Although ±2 mm day−1 might be a good estimate in some cases, it is important to consider the percentage error relative to the observed value. All models produced a higher ETc at the first data point and nearly all models produced a lower ETc during the cutting periods by around 1 mm day−1.
The trends of ETc derived using the KcVI followed closely with the measured ETa largest errors in predicting ETc occurred in the early growing season and during cut periods (Figure 6). The lower ET predictions during the cutting periods are likely due to the spectral bands not capturing irrigation events after a cut when soil evaporation is the dominate component of ET and is cited as a downfall in spectral remote sensing approaches for measuring ET [52,53,54].
The resultant equations can be applied to spectral imagery to calculate a crop coefficient and then applied to Equation (2) to estimate daily ETc. An example map of ETc can be seen in Figure 7 using the NGRDI where the range of ET is between 3.3- and 8.2-mm day−1. The lower ET values are mixed pixels of alfalfa and the narrow dirt access roads.

3.4. Agreement with OpenET Ensemble

The daily estimated ETc for the 20 additional alfalfa fields derived from the KcVI models, aside from the VDVI, demonstrated agreement with the daily estimated OpenET Ensemble model ETa (Table 8). The r2 ranged between 0.68 and 0.79 for all the models except the VDVI (r2 = 0.28). The daily MAE ranged between 0.59- and 0.73-mm day−1, which is similar to the values reported by Volk, et al. [47] for OpenET Ensemble compared to eddy covariance data in croplands (0.83 mm day−1). The daily MSE ranged between 0.61- and 1.10 mm day−1 for all models except the VDVI (2.46 mm day−1). The daily MAPD ranged between 24.9% and 34.1% for all the KcVI models except the VDVI (61.9%) compared to the OpenET Ensemble. The higher MAPD in this analysis could be due to the VI-based models being influenced by cutting events, irrigation events following cutting events, and/or the HLS data containing clouds over some of the fields. The rRMSE ranged between 21.5% and 26.7% for all models except for the VDVI, which had an rRMSE of 39.9% suggesting that all but the VDVI were fair-performing models [51]. This analysis only assesses the agreement between the KcVI model estimated ETc and OpenET Ensemble estimated ETa and does not assess the accuracy of either for the 20 additional alfalfa fields. Thus, it is possible that the KCVI estimates better reflect ETc and the errors are from the OpenET Ensemble model estimates. Five of the six models in the OpenET Ensemble depend on thermal data from Landsat ETM+ and OLI, which have larger spans of days without satellite coverage, thus interpolation of these data could miss cutting and irrigation events that are captured by the HLS products and subsequently captured by the KCVI estimates of ETc.

3.5. Evaluation and Next Steps

This study demonstrated promising results for using several different KcVI equations for predicting ETc; however, there are several limitations. The results of this study are specific to alfalfa grown in the Mesilla Valley using the HLS dataset. The results might not be translatable to other sensors with different spectral bandwidths and center wavelengths and not calibrated to surface reflectance. Furthermore, this study used one year of data and, although some of the comparisons separated the data into training and testing sets, there is an inherent autocorrelation between the in situ data and the KcVI-derived ETc. We attempted to address this shortcoming by the inclusion of an analysis that assesses the agreement of the KcVI model estimates with the OpenET Ensemble model estimates for 20 additional alfalfa fields, which provided evidence that the KcVI approach using HLS data could fill in some of the ET data gaps at the field scale left by the lower temporal resolution of models reliant on Landsat thermal data. Future studies would benefit from having additional years of data and multiple fields with instrumentation to measure in situ ETa.
As outlined in Pôças, et al. [55], experimental conditions such as crop type, irrigation management, and climate conditions introduce variability in empirical relationships for developing KcVI. Although there are additional VIs that could be calculated from the addition of red-edge bands included on the sentinel satellite, these bands are not included on Landsat and, therefore, were not considered in this analysis.
In this study, there were not enough clear satellite observations during the cut and early growth periods to assess whether any of the VIs, particularly those that account for soil background, are able to reduce the error in predicted ETc. Although clouds were not present over the study site for the image dates used, three images did have low/moderate aerosol levels and one image had high aerosol levels which can affect VI calculations, thus caution should be used when calculating VIs from HLS data. Where bare soil is present, differences in soil types across locations will likely have an impact on the regression coefficients [9,16]. Future research could address this objective by separating out imagery taken during the early growth periods and running a statistical analysis to assess whether any of the KcVI models more accurately predict ETc during lower vegetation cover. There could be a vegetation cover threshold where an ensemble model could switch between KcVI models based on cover. This would require more image dates during the early growth/cut periods.
Two notable advancements are addressing these issues and creating new opportunities for operational remote sensing ET products. The relatively new OpenET web application provides estimated ETa at the field scale at daily, monthly, and annual time steps for six ET models and an Ensemble model for agricultural areas in the western United States [46]. There is still a need within the OpenET platform to increase the satellite imagery inputs in order to reduce the interpolations between image dates—perhaps the HLS dataset can help improve the temporal coverage. Secondly, as new commercial satellite constellations with multispectral sensors become more common, there advancements in data harmonization that could lead to producing satellite based ET estimates at high spatial and temporal resolution using data such as the Planet Lab constellation of satellites or incorporating data from UAVs [56].

4. Conclusions

We used the HLS dataset to calculate 16 VIs and used linear regression to calculate crop coefficients by regressing field-measured Kc on the different VIs. All the VIs produced linear models that showed high correlations with the measured Kc values and some of the models produced ETa errors close to ±1 mm day−1. The results suggest that the HLS dataset is an important dataset for filling in the data gap with a higher temporal frequency of satellite data to improve the accuracy of satellite ETa estimates and it can reduce the need to interpolate between missing satellite overpasses. This is particularly important for crops like alfalfa that have multiple harvests during a growing season and when missing one or two Landsat images can cause the omission of nearly an entire cutting period. The results also suggest that when crops are past the initial growth period the KcVI approach can be used to provide accurate ET data at monthly and yearly time scales. Another interesting finding was two vegetation indices that used only spectral bands within the visible spectrum for predicting Kc produced the highest agreement with eddy covariance data. This opens the door for harmonizing other higher spatial resolution satellites or aerial sensors that only collect visible spectra and for applying similar methods for calculating Kc. This approach could be further explored for use on properly calibrated unmanned aerial systems (UAS).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/rs16162876/s1, Table S1: Measured ETa, ETsz, and Kc.

Author Contributions

Conceptualization, R.S.; methodology, R.S.; software, R.S.; validation, R.S., A.S.B., M.B., C.S. and A.F.; formal analysis, R.S. and A.S.B.; investigation, R.S.; resources, A.F. and R.S.; data curation, R.S. and A.S.B.; writing—original draft preparation, R.S.; writing—review and editing, A.S.B., M.B., C.S. and A.F.; visualization, R.S.; supervision, A.S.B., M.B., C.S. and A.F.; project administration, A.F.; funding acquisition, A.F., A.S.B. and R.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Agriculture and Food Research Initiative Competitive Grant no. 971 2021-69012-35916 from the USDA National Institute of Food and Agriculture. Additionally, this material is based upon work supported by the U.S. Geological Survey under Grant/Cooperative Agreement No. G21AP10635.

Data Availability Statement

HLS v2 data can be downloaded through the National Aeronautics and Space Administration website at https://hls.gsfc.nasa.gov/ (accessed on 4 January 2023). Data used in calculating the KcVI values are included in the Supplementary Material.

Acknowledgments

We thank Willie Joe Koenig, Dave Lowry and the New Mexico State University Leyendecker Plant Science Center staff, Liam Sabiston, the staff at the New Mexico Water Resources Research Institute, and Kevin Boyko.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Appendix A

Table A1. List of images and their corresponding aerosol level for the pixels over the study site based on the Fmask used in the study for developing the linear regression equations for predicting Kc from the vegetation indices.
Table A1. List of images and their corresponding aerosol level for the pixels over the study site based on the Fmask used in the study for developing the linear regression equations for predicting Kc from the vegetation indices.
DateDOYSensorStudy Site
Aerosol Level
DateDOYSensorStudy Site
Aerosol Level
9 April 201799SentinelLow 5 September 2017248LandsatLow
30 April 2017120LandsatLow11 September 2017254SentinelLow
19 May 2017139Sentinel Low21 September 2017264LandsatLow
1 June 2017152LandsatHigh1 October 2017274SentinelLow
8 June 2017159SentinelLow6 October 2017279SentinelLow
17 June 2017168LandsatLow7 October 2017280LandsatLow
28 June 2017179SentinelLow/Moderate21 October 2017294SentinelLow
18 July 2017199SentinelLow/Moderate23 October 2017296LandsatLow
19 July 2017200LandsatLow26 October 2017299SentinelLow
2 August 2017214SentinelModerate/High10 November 2017314SentinelLow
4 August 2017216LandsatLow15 November 2017319SentinelLow
7 August 2017219SentinelLow20 November 2017324SentinelLow/Moderate
22 August 2017234SentinelLow24 November 2017328LandsatLow
Table A2. Statistical comparison of linear models derived separately from Landsat and Sentinel using F-test where p < 0.05 indicates a significant difference. None of the linear models showed a significant difference, which suggests good data harmonization between the two satellites.
Table A2. Statistical comparison of linear models derived separately from Landsat and Sentinel using F-test where p < 0.05 indicates a significant difference. None of the linear models showed a significant difference, which suggests good data harmonization between the two satellites.
VIF-Statisticp-ValueVIF-Statisticp-Value
ARVI0.610.72MSI1.880.20
EVI1.580.27NDVI1.670.25
EXG1.690.24NGRDI−0.491.00
GEMI2.270.14NMDI0.860.56
GNDVI1.520.28RDVI1.410.32
GRVI0.030.47TDVI2.040.17
II1.100.44VARI−0.181.00
MSAVI2.090.16VDVI1.120.43

References

  1. 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. Ph.D. Thesis, DLO Winand Staring Centre, Wageningen, The Netherlands, 1995; 271p. [Google Scholar]
  2. Anderson, M.C.; Norman, J.M.; Mecikalski, J.R.; Otkin, J.A.; Kustas, W.P. A climatological study of evapotranspiration and moisture stress across the continental United States based on thermal remote sensing: 2. Surface moisture climatology. J. Geophys. Res. Atmos. 2007, 112. [Google Scholar] [CrossRef]
  3. 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]
  4. Senay, G.B. Satellite Psychrometric Formulation of the Operational Simplified Surface Energy Balance (Ssebop) Model for Quantifying and Mapping Evapotranspiration. Appl. Eng. Agric. 2018, 34, 555–566. [Google Scholar] [CrossRef]
  5. 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, Italy, 1998. [Google Scholar]
  6. Djaman, K.; Irmak, S. Actual crop evapotranspiration and alfalfa-and grass-reference crop coefficients of maize under full and limited irrigation and rainfed conditions. J. Irrig. Drain. Eng. 2013, 139, 433–446. [Google Scholar] [CrossRef]
  7. Rocha, J.; Perdigão, A.; Melo, R.; Henriques, C. Remote sensing based crop coefficients for water management in agriculture. In Sustainable Development—Authoritative and Leading Edge Content for Environmental Management; Curkovic, S., Ed.; INTECH: Haarlem, The Netherlands, 2012; pp. 167–192. [Google Scholar]
  8. Allen, R.G.; Clemmens, A.J.; Burt, C.M.; Solomon, K.; O’Halloran, T. Prediction accuracy for projectwide evapotranspiration using crop coefficients and reference evapotranspiration. J. Irrig. Drain. Eng. 2005, 131, 24–36. [Google Scholar] [CrossRef]
  9. Hunsaker, D.J.; Pinter, P.J.; Kimball, B.A. Wheat basal crop coefficients determined by normalized difference vegetation index. Irrig. Sci. 2005, 24, 1–14. [Google Scholar] [CrossRef]
  10. Zanter, K. Landsat 8 (L8) Data Users Handbook; Landsat Science Official Website USGS: Reston, VA, USA, 2016; p. 33.
  11. Sellers, P.; Mintz, Y.; Sud, Y.C.; Dalcher, A. A simple biosphere model (SiB) for use within general circulation models. J. Atmos. Sci. 1986, 43, 505–531. [Google Scholar] [CrossRef]
  12. Jensen, J.R. Introductory Digital Image Processing: A Remote Sensing Perspective, 3rd ed.; Prentice Hall: Upper Saddle River, NJ, USA, 2005; p. 526. [Google Scholar]
  13. Baret, F. TSAVI: A vegetation index which minimizes soil brightness effects on LAI and APAR estimation. In Proceedings of the 12th Canadian Symp. on Remote Sensing and IGARSS’90, Vancouver, BC, Canada, 10–14 July 1989. [Google Scholar]
  14. Campos, I.; Neale, C.M.; Suyker, A.E.; Arkebauer, T.J.; Gonçalves, I.Z. Reflectance-based crop coefficients REDUX: For operational evapotranspiration estimates in the age of high producing hybrid varieties. Agric. Water Manag. 2017, 187, 140–153. [Google Scholar] [CrossRef]
  15. Kamble, B.; Kilic, A.; Hubbard, K. Estimating Crop Coefficients Using Remote Sensing-Based Vegetation Index. Remote Sens. 2013, 5, 1588. [Google Scholar] [CrossRef]
  16. French, A.N.; Hunsaker, D.J.; Sanchez, C.A.; Saber, M.; Gonzalez, J.R.; Anderson, R. Satellite-based NDVI crop coefficients and evapotranspiration with eddy covariance validation for multiple durum wheat fields in the US Southwest. Agric. Water Manag. 2020, 239, 106266. [Google Scholar] [CrossRef]
  17. Rouse, J.W., Jr.; Haas, R.; Schell, J.; Deering, D. Monitoring vegetation systems in the Great Plains with ERTS. In Proceedings of the Third Earth Resources Technology Satellite-l Symposium, Washingdon, WA, USA, 10–14 December 1973. [Google Scholar]
  18. Huete, A.R. A soil-Adjusted vegetation index (SAVI). Remote Sens. Environ. 1988, 25, 295–309. [Google Scholar] [CrossRef]
  19. Roujean, J.-L.; Breon, F.-M. Estimating PAR absorbed by vegetation from bidirectional reflectance measurements. Remote Sens. Environ. 1995, 51, 375–384. [Google Scholar] [CrossRef]
  20. Qi, J.; Chehbouni, A.; Huete, A.R.; Kerr, Y.H.; Sorooshian, S. A Modified Soil Adjusted Vegetation Index. Remote Sens. Environ. 1994, 48, 119–126. [Google Scholar] [CrossRef]
  21. Gitelson, A.A.; Kaufman, Y.J.; Stark, R.; Rundquist, D. Novel algorithms for remote estimation of vegetation fraction. Remote Sens. Environ. 2002, 80, 76–87. [Google Scholar] [CrossRef]
  22. Gitelson, A.A.; Kaufman, Y.J.; Merzlyak, M.N. Use of a green channel in remote sensing of global vegetation from EOS-MODIS. Remote Sens. Environ. 1996, 58, 289–298. [Google Scholar] [CrossRef]
  23. Hardisky, M.; Klemas, V.; Smart, M. The influence of soil salinity, growth form, and leaf moisture on the spectral radiance of Spartina alterniflora canopies. Eng. Remote Sens. 1983, 49, 77–83. [Google Scholar]
  24. Wang, L.; Qu, J.J. NMDI: A normalized multi-band drought index for monitoring soil and vegetation moisture with satellite remote sensing. Geophys. Res. Lett. 2007, 34, L20405. [Google Scholar] [CrossRef]
  25. Gonzalez-Piqueras, J.; Calera, A.; Gilabert, M.A.; Cuesta, A.; De la Cruz Tercero, F. Estimation of crop coefficients by means of optimized vegetation indices for corn. In Proceedings of the Remote Sensing for Agriculture, Ecosystems, and Hydrology V, Barcelona, Spain, 24 February 2004; pp. 110–118. [Google Scholar]
  26. Samani, Z.; Skaggs, R.; Longworth, J. Alfalfa Water Use and Crop Coefficients across the Watershed: From Theory to Practice. J. Irrig. Drain. Eng. 2013, 139, 341–348. [Google Scholar] [CrossRef]
  27. Claverie, M.; Ju, J.; Masek, J.G.; Dungan, J.L.; Vermote, E.F.; Roger, J.-C.; Skakun, S.V.; Justice, C. The Harmonized Landsat and Sentinel-2 surface reflectance data set. Remote Sens. Environ. 2018, 219, 145–161. [Google Scholar] [CrossRef]
  28. Malm, N.R. Climate Guide Las Cruces, 1892–2000; New Mexico State University Agricultural Experiment Station: Las Cruces, NM, USA, 2003. [Google Scholar]
  29. Albano, C.M.; Abatzoglou, J.T.; McEvoy, D.J.; Huntington, J.L.; Morton, C.G.; Dettinger, M.D.; Ott, T.J. A Multidataset Assessment of Climatic Drivers and Uncertainties of Recent Trends in Evaporative Demand across the Continental United States. J. Hydrometeorol. 2022, 23, 505–519. [Google Scholar] [CrossRef]
  30. NMDA. New Mexico Agricultural Statistics 2018 Annual Bulletin; Department of Agriculture State of New Mexicomsc: Las Cruces, NM, USA, 2019.
  31. Sanderson, M.A.; Karnezos, T.; Matches, A. Morphological development of alfalfa as a function of growing degree days. J. Prod. Agric. 1994, 7, 239–242. [Google Scholar] [CrossRef]
  32. Boyko, K.; Fernald, A.G.; Bawazir, A.S. Improving groundwater recharge estimates in alfalfa fields of New Mexico with actual evapotranspiration measurements. Agric. Water Manag. 2020, 244, 106532. [Google Scholar] [CrossRef]
  33. Hoffmann, H.; Nieto, H.; Jensen, R.; Guzinski, R.; Zarco-Tejada, P.; Friborg, T. Estimating evaporation with thermal UAV data and two-source energy balance models. Hydrol. Earth Syst. Sci. 2016, 20, 697. [Google Scholar] [CrossRef]
  34. Bawazir, A.S.; Luthy, R.; King, J.P.; Tanzy, B.F.; Solis, J. Assessment of the crop coefficient for saltgrass under native riparian field conditions in the desert southwest. Hydrol. Process. 2014, 28, 6163–6171. [Google Scholar] [CrossRef]
  35. Xue, J.; Su, B. Significant remote sensing vegetation indices: A review of developments and applications. J. Sens. 2017, 2017, 1–17. [Google Scholar] [CrossRef]
  36. Jafari, R.; Lewis, M.M.; Ostendorf, B. Evaluation of vegetation indices for assessing vegetation cover in southern arid lands in South Australia. Rangel. J. 2007, 29, 39–49. [Google Scholar] [CrossRef]
  37. Bannari, A.; Morin, D.; Bonn, F.; Huete, A. A review of vegetation indices. Remote Sens. Rev. 1995, 13, 95–120. [Google Scholar] [CrossRef]
  38. Kaufman, Y.J.; Tanre, D. Atmospherically Resistant Vegetation Index (Arvi) for Eos-Modis. IEEE Trans. Geosci. Remote Sens. 1992, 30, 261–270. [Google Scholar] [CrossRef]
  39. Huete, A.; Justice, C.; Liu, H. Development of Vegetation and Soil Indices for Modis-Eos. Remote Sens. Environ. 1994, 49, 224–234. [Google Scholar] [CrossRef]
  40. Woebbecke, D.M.; Meyer, G.E.; Von Bargen, K.; Mortensen, D.A. Color Indices for Weed Identification under Various Soil, Residue, and Lighting Conditions. Trans. ASAE 1995, 38, 259–269. [Google Scholar] [CrossRef]
  41. Pinty, B.; Verstraete, M.M. Gemi: A Non-Linear Index to Monitor Global Vegetation from Satellites. Vegetatio 1992, 101, 15–20. [Google Scholar] [CrossRef]
  42. Sripada, R.P.; Heiniger, R.W.; White, J.G.; Weisz, R. Aerial Color Infrared Photography for Determining Late-Season Nitrogen Requirements in Corn. Agron. J. 2005, 97, 1443–1451. [Google Scholar] [CrossRef]
  43. Rock, B.N.; Vogelmann, J.E.; Williams, D.L.; Vogelmann, A.F.; Hoshizaki, T. Remote Detection of Forest Damage. Bioscience 1986, 36, 439–445. [Google Scholar] [CrossRef]
  44. Bannari, A.; Asalhi, H.; Teillet, P.M. Transformed Difference Vegetation Index (Tdvi) for Vegetation Cover Mapping. In Proceedings of the IEEE International Geoscience and Remote Sensing Symposium, Toronto, ON, Canada, 24–28 June 2002. [Google Scholar]
  45. Xiaoqin, W.; Miaomiao, W.; Shaoqiang, W.; Yundong, W. Extraction of Vegetation Information from Visible Unmanned Aerial Vehicle Images. Trans. Chin. Soc. Agric. Eng. 2015, 31, 152–159. [Google Scholar]
  46. Melton, F.S.; Huntington, J.; Grimm, R.; Herring, J.; Hall, M.; Rollison, D.; Erickson, T.; Allen, R.; Anderson, M.; Fisher, J.B. OpenET: Filling a critical data gap in water management for the western United States. JAWRA J. Am. Water Resour. Assoc. 2022, 58, 971–994. [Google Scholar] [CrossRef]
  47. Volk, J.M.; Huntington, J.L.; Melton, F.S.; Allen, R.; Anderson, M.; Fisher, J.B.; Kilic, A.; Ruhoff, A.; Senay, G.B.; Minor, B. Assessing the accuracy of OpenET satellite-based evapotranspiration data to support water resource and land management applications. Nat. Water 2024, 2, 193–205. [Google Scholar] [CrossRef]
  48. Sabiston, L.; Sabie, R.; Buenemann, M.; Stringam, B.; Fernald, A. Comparing field-scale eddy covariance measurements and crop coefficient estimates of alfalfa evapotranspiration to OpenET model estimates and exploring water budget implications in a dryland environment. Irrig. Sci. 2024, 42, 1–18. [Google Scholar] [CrossRef]
  49. Tawalbeh, Z.M.; Bawazir, A.S.; Fernald, A.; Sabie, R.; Heerema, R.J. Assessing Satellite-Derived OpenET Platform Evapotranspiration of Mature Pecan Orchard in the Mesilla Valley, New Mexico. Remote Sens. 2024, 16, 1429. [Google Scholar] [CrossRef]
  50. Despotovic, M.; Nedic, V.; Despotovic, D.; Cvetanovic, S. Evaluation of empirical models for predicting monthly mean horizontal diffuse solar radiation. Renew. Sustain. Energy Rev. 2016, 56, 246–260. [Google Scholar] [CrossRef]
  51. Li, M.-F.; Tang, X.-P.; Wu, W.; Liu, H.-B. General models for estimating daily global solar radiation for different solar radiation zones in mainland China. Energy Convers. Manag. 2013, 70, 139–148. [Google Scholar] [CrossRef]
  52. Glenn, E.P.; Huete, A.R.; Nagler, P.L.; Hirschboeck, K.K.; Brown, P. Integrating remote sensing and ground methods to estimate evapotranspiration. Crit. Rev. Plant Sci. 2007, 26, 139–168. [Google Scholar] [CrossRef]
  53. Anderson, M.C.; Allen, R.G.; Morse, A.; Kustas, W.P. Use of Landsat thermal imagery in monitoring evapotranspiration and managing water resources. Remote Sens. Environ. 2012, 122, 50–65. [Google Scholar] [CrossRef]
  54. Nagler, P.L.; Glenn, E.P.; Nguyen, U.; Scott, R.L.; Doody, T. Estimating riparian and agricultural actual evapotranspiration by reference evapotranspiration and MODIS enhanced vegetation index. Remote Sens. 2013, 5, 3849–3871. [Google Scholar] [CrossRef]
  55. Pôças, I.; Calera, A.; Campos, I.; Cunha, M. Remote sensing for estimating and mapping single and basal crop coefficientes: A review on spectral vegetation indices approaches. Agric. Water Manag. 2020, 233, 106081. [Google Scholar] [CrossRef]
  56. Gao, R.; Yang, Y.; Knipper, K.; Alsina, M.M.; Sanchez, L.A.; Melton, F.; Nieto, H.; Bambach, N.E.; Gao, F.; Alfieri, J. 38HYDRO Estimating 3-m Evapotranspiration Using Planet, OpenET, and Machine Learning Techniques. In Proceedings of the 104th AMS Annual Meeting, Baltimore, MD, USA, 28 January–1 February 2024. [Google Scholar]
Figure 1. The location of the alfalfa study site (purple border) in the Mesilla Valley (red circle), New Mexico, USA.
Figure 1. The location of the alfalfa study site (purple border) in the Mesilla Valley (red circle), New Mexico, USA.
Remotesensing 16 02876 g001
Figure 2. Processing steps used for creating and analyzing vegetation index-based crop coefficients.
Figure 2. Processing steps used for creating and analyzing vegetation index-based crop coefficients.
Remotesensing 16 02876 g002
Figure 3. (a) Distribution of measured Kc values for the 2017 growing season, and (b) satellite acquisitions plotted on the daily measured Kc values with cuts indicated by the vertical dashed lines.
Figure 3. (a) Distribution of measured Kc values for the 2017 growing season, and (b) satellite acquisitions plotted on the daily measured Kc values with cuts indicated by the vertical dashed lines.
Remotesensing 16 02876 g003
Figure 4. Linear regressions between 16 vegetation indices and Kc from in situ data. The red line represents the best-fit line.
Figure 4. Linear regressions between 16 vegetation indices and Kc from in situ data. The red line represents the best-fit line.
Remotesensing 16 02876 g004
Figure 5. Graphs showing higher and lower prediction of ETc using the KcVI method compared to measured ETa. The red line is a visualization of zero deviation.
Figure 5. Graphs showing higher and lower prediction of ETc using the KcVI method compared to measured ETa. The red line is a visualization of zero deviation.
Remotesensing 16 02876 g005
Figure 6. Comparison of predicted crop ET (ETc) using the KCVI method and measured ETa from an eddy covariance tower for alfalfa during the 2017 growing season.
Figure 6. Comparison of predicted crop ET (ETc) using the KCVI method and measured ETa from an eddy covariance tower for alfalfa during the 2017 growing season.
Remotesensing 16 02876 g006
Figure 7. Example of KCVI output map showing estimated ET (mm day−1) using NGRDI for 19 May 2017.
Figure 7. Example of KCVI output map showing estimated ET (mm day−1) using NGRDI for 19 May 2017.
Remotesensing 16 02876 g007
Table 1. Cumulative irrigation and precipitation depths between irrigation events.
Table 1. Cumulative irrigation and precipitation depths between irrigation events.
Irrigation EventDateDOYIrrigation (mm)Precipitation (mm)
13 March 2017622020
221 March 2017801410
34 April 2017941422
43 May 20171231370
516 May 20171361735
611 June 20171621601
77 July 2017188176154
816 August 201722813443
915 September 20172582221
1025 September 201726818815
111 November 201730519911
Table 2. Alfalfa harvesting dates for the study site during the 2017 growing season. DOY is the day of the year.
Table 2. Alfalfa harvesting dates for the study site during the 2017 growing season. DOY is the day of the year.
Cutting EventCut DateDOY
120 April 2017110
22 June 2017153
329 June 2017180
46 August 2017218
54 September 2017247
618 October 2017291
727 November 2017331
Table 3. Spectral characteristics of satellite bands prior to harmonization (nm).
Table 3. Spectral characteristics of satellite bands prior to harmonization (nm).
Landsat OLISentinel-2
Center wavelengthBlue482.04492
Green561.41560
Red654.59665
Near Infrared864.47833
Shortwave Infrared 11608.861614
Shortwave Infrared 22220.02190
BandwidthBlue60.0466
Green57.3336
Red37.4731
Near Infrared28.25106
Shortwave Infrared 184.791
Shortwave Infrared 2175180
Table 4. Equations of spectral indices selected for use in the vegetation index-based crop coefficient index (KcVI) analysis.
Table 4. Equations of spectral indices selected for use in the vegetation index-based crop coefficient index (KcVI) analysis.
NameEquationReference
Atmospherically Resistant Vegetation Index A R V I = ρ N I R ( 2 × ρ r e d + ρ B l u e ) ρ N I R + ( 2 × ρ r e d + ρ B l u e ) [38]
Enhanced Vegetation Index E V I = 2.5 ρ N I R ρ r e d 1 + ρ N I R + 6 ρ r e d 7.5 ρ B l u e [39]
Excess Green Index E x G = 2 × ρ g r e e n ρ r e d ρ B l u e [40]
Global Environment Monitoring Index G E M I = η ( 1 0.25   η ) ρ r e d 0.125 1 ρ r e d η = 2   ρ N I R 2 ρ r e d 2 + 1.5   ρ N I R + 0.5 ρ r e d ρ N I R + ρ r e d + 0.5 [41]
Green NDVI G N D V I = ( ρ N I R ρ g r e e n ) ( ρ N I R + ρ g r e e n ) [22]
Green Ratio Vegetation Index G R V I = ρ N I R ρ g r e e n [42]
Infrared Index I I = ( ρ N I R ρ M i d I R ) ( ρ N I R + ρ M i d I R )   [23]
Modified Soil Adjusted Vegetation Index M S A V I 2 = 1 × ρ N I R + 1 ( 2 × ρ N I R + 1 ) 2 8 × ( ρ N I R ρ r e d ) 2 [20]
Moisture Stress Index M S I = ρ M i d I R ρ N I R [43]
Normalized Green-Red Difference Index N G R D I = ( ρ g r e e n ρ r e d ) ( ρ g r e e n + ρ r e d ) [21]
Normalized Difference Vegetation Index N D V I = ρ N I R ρ r e d ρ N I R + ρ r e d [17]
Normalized Multi-band Drought Index N M D I = ρ N I R ρ 1640 ρ 2130 ρ N I R + ( ρ 1640 ρ 2130 ) [24]
Renormalized Difference Vegetation Index R D V I = ρ N I R ρ r e d ρ N I R + ρ r e d [19]
Transformed Difference Vegetation Index T D V I = 0.5 + ρ N I R ρ r e d ρ N I R + ρ r e d [44]
Visible Atmospherically Resistant Index V A R I = ρ g r e e n ρ r e d ρ g r e e n + ρ r e d + ρ b l u e [21]
Visible-band Difference Vegetation Index V D V I = 2 × ρ g r e e n ρ r e d ρ b l u e 2 × ρ g r e e n + ρ r e d + ρ b l u e [45]
Table 5. Linear regression equations derived from vegetation indices and field-measured Kc values for alfalfa with r2 and RMSE values (dimensionless).
Table 5. Linear regression equations derived from vegetation indices and field-measured Kc values for alfalfa with r2 and RMSE values (dimensionless).
VIKcVIr2RMSEVIKcVIr2RMSE
ARVI1.59 × ARVI + 0.540.720.16MSI−0.89 × MSI + 1.520.850.12
EVI1.28 × EVI + 0.20.740.15NDVI1.48 × NDVI − 0.120.880.10
EXG14.2 × EXG + 0.330.810.13NGRDI1.65 × NGRDI + 0.690.900.09
GEMI1.72 × GEMI − 0.420.670.17NMDI2.16 × NMDI + 0.11 0.760.15
GNDVI2.26 × GNDVI − 0.590.820.13RDVI1.77 × RDVI + 0.070.760.15
GRVI0.11 × GRVI + 0.30.700.16TDVI3.16 × TDVI − 2.530.880.10
II1.29 × II + 0.620.820.13VARI2.17 × VARI + 0.690.900.10
MSAVI1.34 × MSAVI + 0.220.740.15VDVI−4.55 × VDVI + 1.820.760.15
Table 6. Summary of KcVI models that resulted in significant differences in the Wilcoxon signed-rank test between measured ETa and predicted ETc. for 10 model runs for each VI.
Table 6. Summary of KcVI models that resulted in significant differences in the Wilcoxon signed-rank test between measured ETa and predicted ETc. for 10 model runs for each VI.
IndexCountIndexCount
ARVI2MSI1
EVI2NDVI1
EXG3NGRDI3
GEMI2NMDI1
GNDVI0RDVI2
GRVI1TDVI2
II1VARI3
MSAVI2VDVI2
Table 7. Statistical comparison results of predicted ETc using the KcVI method and observed ETa.
Table 7. Statistical comparison results of predicted ETc using the KcVI method and observed ETa.
VIMAE (mm/day)MSE
(mm/day)
MAPD (%)rRMSE (%)VIMAE (mm/day)MSE (mm/day)MAPD (%)rRMSE (%)
ARVI0.620.6415.717.2MSI0.490.3513.112.8
EVI0.600.6115.016.8NDVI0.410.2612.011.0
EXG0.510.4214.013.8NGRDI0.350.2010.09.5
GEMI0.630.7516.318.6NMDI0.570.5114.415.3
GNDVI0.510.3913.913.3RDVI0.580.5714.716.2
GRVI0.640.6016.516.6TDVI0.420.2812.311.3
II0.510.4013.013.5VARI0.360.2110.09.7
MSAVI0.590.6114.816.7VDVI0.520.4216.313.9
Table 8. Statistical comparison results of predicted ETc from the 20 additional alfalfa fields using the KcVI method and estimated ETa from the OpenET Ensemble model.
Table 8. Statistical comparison results of predicted ETc from the 20 additional alfalfa fields using the KcVI method and estimated ETa from the OpenET Ensemble model.
VIr2MAE (mm/day)MSE
(mm/day)
MAPD (%)rRMSE (%)VIr2MAE (mm/day)MSE (mm/day)MAPD (%)rRMSE (%)
ARVI0.710.710.9830.4225.1MSI0.790.590.7124.9121.5
EVI0.760.660.8229.5523.1NDVI0.750.680.8526.4123.4
EXG0.750.690.8726.9923.7NGRDI0.750.680.8430.9523.3
GEMI0.680.731.1033.4426.7NMDI0.750.670.8634.9623.6
GNDVI0.710.730.9929.6525.3RDVI0.760.660.8227.9123.0
GRVI0.730.730.9434.1624.7TDVI0.740.690.8725.9423.8
II0.780.610.7427.8821.9VARI0.750.690.8631.2023.6
MSAVI0.760.670.8429.6423.3VDVI0.281.192.4661.8739.9
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sabie, R.; Bawazir, A.S.; Buenemann, M.; Steele, C.; Fernald, A. Calculating Vegetation Index-Based Crop Coefficients for Alfalfa in the Mesilla Valley, New Mexico Using Harmonized Landsat Sentinel-2 (HLS) Data and Eddy Covariance Flux Tower Data. Remote Sens. 2024, 16, 2876. https://doi.org/10.3390/rs16162876

AMA Style

Sabie R, Bawazir AS, Buenemann M, Steele C, Fernald A. Calculating Vegetation Index-Based Crop Coefficients for Alfalfa in the Mesilla Valley, New Mexico Using Harmonized Landsat Sentinel-2 (HLS) Data and Eddy Covariance Flux Tower Data. Remote Sensing. 2024; 16(16):2876. https://doi.org/10.3390/rs16162876

Chicago/Turabian Style

Sabie, Robert, A. Salim Bawazir, Michaela Buenemann, Caitriana Steele, and Alexander Fernald. 2024. "Calculating Vegetation Index-Based Crop Coefficients for Alfalfa in the Mesilla Valley, New Mexico Using Harmonized Landsat Sentinel-2 (HLS) Data and Eddy Covariance Flux Tower Data" Remote Sensing 16, no. 16: 2876. https://doi.org/10.3390/rs16162876

APA Style

Sabie, R., Bawazir, A. S., Buenemann, M., Steele, C., & Fernald, A. (2024). Calculating Vegetation Index-Based Crop Coefficients for Alfalfa in the Mesilla Valley, New Mexico Using Harmonized Landsat Sentinel-2 (HLS) Data and Eddy Covariance Flux Tower Data. Remote Sensing, 16(16), 2876. https://doi.org/10.3390/rs16162876

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