Next Article in Journal
Machine Learning Classification of Roasted Arabic Coffee: Integrating Color, Chemical Compositions, and Antioxidants
Next Article in Special Issue
Spatio-Temporal Distribution and Trends of Major Agricultural Crops in Romania Using Interactive Geographic Information System Mapping
Previous Article in Journal
Competitive Firms’ Optimal Decisions on Entering Accessory Market
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Spatio-Temporal Description of the NDVI (MODIS) of the Ecuadorian Tussock Grasses and Its Link with the Hydrometeorological Variables and Global Climatic Indices

by
Jhon Villarreal-Veloz
1,2,
Xavier Zapata-Ríos
1,2,*,
Karla Uvidia-Zambrano
1 and
Carla Borja-Escobar
1
1
Department of Civil and Environmental Engineering, Escuela Politécnica Nacional, Ladrón de Guevara E11-253, Quito P.O. Box 17-01-2759, Ecuador
2
Center for Research and Water Resources Studies (CIERHI), Escuela Politécnica Nacional, Ladrón de Guevara E11-253, Quito P.O. Box 17-01-2759, Ecuador
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(15), 11562; https://doi.org/10.3390/su151511562
Submission received: 8 May 2023 / Revised: 25 June 2023 / Accepted: 28 June 2023 / Published: 26 July 2023
(This article belongs to the Special Issue The Detection and Application of Remote Sensing)

Abstract

:
This study examined the changes in tussock grass greenness over 18 years (2001–2018) using NDVI data from 10 key areas of the Páramo ecosystem in the Ecuadorian Andes. In addition, the study investigated the influence of hydrometeorological variables (precipitation, soil temperature, and water availability) and climatic indices (AAO, MEI, MJO, NAO, PDO, El Niño 1 + 2, 3, 3.4, and 4) on greenness dynamics. The spatial and temporal variations of NDVI were studied, applying several analysis and indicators, such as: the standard deviation, z-score anomalies, Sen slope, Mann–Kendall test, and time integrated-NDVI (TI-NDVI). Linear and multilinear correlations were used to evaluate the influence of hydrometeorological variables and climatic indices on the greenness of tussock. The findings of the study show that Páramo, located in the Inter-Andean valley above 2° S, is the most productive, followed by those located in the Royal Range (eastern cordillera). The anomalies and trends of NDVI on the Royal Range tended to be greening over time. NDVI showed a moderate multilinear correlation with precipitation and soil temperature, and a strong response to water availability. Finally, NDVI was weakly linearly related to the climatic indices, the most representative being the MJO, and slightly related to ENSO events. Understanding the regional and global-scale variables that control tussock grasses’ phenology will help to determine how present and future climate changes will impact this ecosystem.

1. Introduction

Along the Andes Mountain range in its main north–south direction, gradients in climate are observed because this mountain chain is exposed to different patterns of atmospheric circulation and various global climatic phenomena. Consequently, the western and eastern slopes of the Andes are characterized by contrasting climatic asymmetries [1]. This climatological variability in the region is manifested by variability in the seasonal and spatial patterns of precipitation and air temperature [2]. In the equatorial Andes, moisture sources originate from the Amazon Basin and the tropical Atlantic from the east, and from the Pacific Ocean from the west [3,4]. In this section of the Andes, precipitation amounts are greatest north of the equator, on eastern slopes, and at low to medium elevations [4]. In the inter-Andean valleys, located between the western and eastern cordillera (Royal Range), two rainy seasons are common from February to May and October to November as a consequence of the two moisture sources, together with the Inter-Tropical Convergence Zone oscillation [5]. Interannual and interdecadal climate variability is also observed because of large-scale phenomena such as the El Niño Southern Oscillation (ENSO) [2].
The knowledge of the climate in the Ecuadorian Andes is largely based on meteorological observations derived from field stations [2,4,6]. However, a limited number of stations with a short observation period are located at elevations above 3000 m.a.s.l. The National Institute of Meteorology and Hydrology of Ecuador (INAMHI) shows in its database a total of 34 meteorological stations (13.6% of all stations in the country) with elevations above 3000 m.a.s.l. However, only 50% of these stations are currently in operation. On the other hand, information derived from regional and global climate models has allowed an understanding of the climate at regional scales in the Andes [7]. Unfortunately, these models still report climate variability in the Andes with high uncertainty due to their coarse spatial resolution and inability to capture the strong climate gradients and topographical heterogeneity in the region [8,9]. Consequently, there are still relevant processes in mountain areas that are not fully understood. For example, there are interactions between climate and vegetation [9]. This knowledge gap includes how changes in climate variability (e.g., ENSO) influence ecohydrological processes and the spatial extent of ENSO influence in the Andes [4,10,11]. Regrettably, mountain ecosystems such as the high Andes are expected to be some of the most affected by climate change [12]. Consequently, there is a need to investigate these issues and fill existing knowledge gaps.
The seasonal variation of vegetation observed by remote sensing has been considered as a simple proxy for many ecosystem processes that take place on the planet [13,14,15]. The structures and processes that occur in terrestrial ecosystems are highly dependent on climate [16]. Specifically, the variability of the spectral indices of the land surface vegetation has been very useful in understanding aboveground net primary production [17,18,19], phenology [14], canopy structure [20], agricultural processes and silvicultural [21], soil carbon stocks [17,22,23], water availability [17], droughts [17,24], energy cycles [25,26], and biodiversity [27,28,29], among others. A specific application has been the understanding of the relationship between vegetation responses to climate and large circulation patterns [30,31,32,33]. Therefore, the seasonal variation in vegetation over the land surface observed by remote sensing is considered a good indicator of environmental change [34] and a broad indicator of climate variability [32,35]. Temporal changes within an ecosystem can be monitored by the greenness variability referred to in this study as the vegetation dynamics [36]. As a result, satellite information has been an important tool for understanding vegetation dynamics at regional and global scales [37,38,39].
A large number of satellite sensors have been used to obtain vegetation indices. The Moderate Resolution Imaging Spectroradiometer (MODIS) is one of the most common sources of information. MODIS has provided data for ecosystem monitoring with appropriate spatial and temporal resolutions and improved geometric and radiometric properties [40,41]. The spatial greenness dynamics were derived using MODIS Normalized Difference Vegetation Index (NDVI) time series data [32]. This vegetation index is the most widely used indicator for monitoring vegetation from satellites [42]. Positive and negative NDVI trends have been found in various vegetation classes in southern South America, including the Andes. The Multivariate El Niño Index of the Southern Oscillation (ENSO) and the Index of Antarctic Oscillation (AAO) explain the variability in annual phenology in some regions of southern South America [32]. In tropical South America, it was observed that the NDVI responds slightly to ENSO events [43], as does the NDVI over the Colombian Andes [44]. In Ecuador, Silvia [45] studied the temporality of NDVI and concluded that the vigor of the plant is greater in the Amazon region, followed by the Andean and coastal regions. In addition, NDVI studies on wetlands in the southern coastal region of Ecuador have demonstrated a direct relationship between NDVI and biomass with precipitation and air temperature. In contrast, in a small central Andean sector of Ecuador, satellite information (NDVI, CWSI: Crop Water Stress Index and SAVI: Soil Adapted Vegetation Index) made it possible to determine the best form of irrigation (drip irrigation) for Opuntia ficus [46].
The tropical Andes are a global conservation hotspot because of their large number of endemic plants and the degree of threat to their ecosystems [47]. Páramos belong to one of these ecosystems and extend from 2800 to 4700 m.a.s.l [48]. These extend between latitudes 11° N and 8° S over 35,000 km2 [49]. This ecosystem is essential because it provides important environmental services, such as water regulation and supply, carbon storage, soil formation, biodiversity conservation, climate change mitigation, recreational services, and food, among others [1,50]. In addition, it houses endemic animals, a diversity of flora, and indigenous cultures that contribute to planetary ecosystem diversity [51]. Rapid population growth, deforestation, mining and subsistence farming, drainage of wetlands, frequent burning of vegetation, and changes in land cover are threatening the high Andean ecosystems in the Andes [52,53].
Existing studies on the Páramo throughout the Ecuadorian Andes lack a detailed analysis of the response of vegetation to different global climatic factors. A quantitative evaluation of the changes in vegetation has the potential to define the interaction between climate and the response of Páramo vegetation. In this study, the Ecuadorian Andes were defined at 1° N and 5° S latitudes. This section of the Andes is characterized by rapid changes in elevation over short distances. NDVI was utilized, which provides continuous vegetation monitoring with spatial and temporal consistency, allowing a comparison of Vegetation Index (VI) dynamics between different study areas within a region of interest. Frequent and regular updates of the vegetation index and its high spatial and temporal resolutions are among the advantages of using VIs over other hydrometeorological variables.
Throughout the world, strong relationships have been found between vegetation indices and climatic variables such as precipitation and air temperature [54,55]. Satellite-derived precipitation and air temperature have also been available since the late 1970s. However, these estimates are still inaccurate for mountainous topographies [56,57]. Nevertheless, this relationship is dependent on geographic location; therefore, analyses of different ecosystems should be performed. It is widely recognized that meteorological variables such as precipitation and air temperature have a significant influence on the spectral indices of vegetation [58]. However, few studies have examined these patterns in the Ecuadorian Andes.
This study postulates that the dynamics of vegetation in the Páramo ecosystem, located at high elevations in the Andes responds to climatic conditions and water availability. In addition, global weather patterns such as El Niño Oscillation should influence the greenness dynamics of this ecosystem. This postulate is based on the case of tropical glaciers which have been widely studied as indicators of climate variability in mountain ecosystems [59,60], and the widely recognized impact of ENSO on the dynamics of tropical glaciers [61]. Therefore, the main objective of this research is to characterize the seasonal patterns and the interannual variability of the dynamics of vegetation greenness in the Páramo ecosystem using MODIS-NDVI in a period of time between the years 2001 to 2018. In the context of this general objective, the following research questions were formulated: (i) What is the spatial and temporal variability of NDVI at different locations within tussock grasses throughout the equatorial Andes? (ii) How do the vegetation indices respond to changes in rainfall, soil temperature and water availability? (iii) Is there any temporal trend in the NDVI time series over the 18 years of observations? (iv) How does NDVI respond to different global climate indices like MEI, AAO and El Niño indices?

2. Materials and Methods

2.1. Study Area

Ecuador is located in the north-western corner of South America (Figure 1a), and is crossed by the Andes Mountain range, which runs from north to south. The Andes in Ecuador are made up of two main mountain ranges: The Royal Range or Eastern Range and the Western Range. Both mountain ranges hydrologically divide Ecuador into two main basins: the Pacific Ocean and Amazon/Atlantic. In addition, the Andes divide the territory of Ecuador into three regions (Coastal, Andean and Amazon), which are characterized by their own seasonality (Table 1). The mountain range extends to 6° latitude (1.2° N–4.8° S), with an approximate area of 60,400 km2, which corresponds to 24% of the surface of Ecuador. The elevation of this region is in a range from 1500 to 6310 m.a.s.l. [62]. In the Andean region there is a population of approximately 6.1 million inhabitants with an economy based on various activities: (i) agriculture, where the main products are potatoes, strawberries, wheat, barley and carrots; (ii) cattle raising, reaching 48.4% of Ecuador’s production [63]; and (iii) mining, where the main extracted resources are gold, silver, bronze, antimony, lead, zinc and platinum [64]. Additionally, the Andes have an ideal environment for the diversification of multiple ecosystems because of their geomorphological characteristics, soil types, precipitation patterns, and temperature gradients [65]. Within this territory, 6 main ecosystems have been identified: Páramo, Humid Puna, Dry Puna, Cloud Forest, Seasonal Andean Forest, and Dry Andean Forest [66].
The Páramos develop in humid and sub-humid regions and in climatic conditions characterized by low or freezing temperatures and high radiation because of their location at high elevations [1,48]. Although the climate in this ecosystem is fairly uniform, extreme daily changes are common [48]. The moisture sources originate from the Amazon region and/or the warm and cold ocean currents of the Pacific Ocean [3,4,48]. Precipitation patterns are generally bimodal with a central dry season and have a typical range of 800–2000 mm/year [6,48,67]. Andosols with a deep A horizon and high concentrations of organic matter are typically found in Páramos [68,69]. The soils are dark in color, acidic, have a high cation exchange capacity, and are generally poorly developed [67].
Focusing on the vegetation of the Páramo ecosystem, Harling [70] defined three types of Páramo in Ecuador: (i) tussock, (ii) cushions, and (iii) desert Páramos. The tussock grasses are between 3000 and 4000 m.a.s.l (Figure 2) and are dominated by Calamagrostis, Festuca and Stipa, with the presence of meadows and pads. The cushion Páramos occupy areas above 4000 m.a.s.l. where the grasslands are replaced by Azorella pedunculata, meadows, thickets, and rosettes represented by Espeletia pycnophylla. Desert or sandy Páramos that present restrictive conditions for vegetation growth are located at elevations greater than 4300 m.a.s.l. In the present study, tussock grass Páramos have been selected as an ideal sentinel ecosystem for the analysis of the dynamics of the vegetation and its relationship with the climate, showing its predominance in the Ecuadorian Andes, due to the fact that they extend from north to south in an approximate area of 9650 km2, equivalent to 64% of the surface of the Ecuadorian Páramo.
Table 1. Rainy and dry seasons for the three regions of continental Ecuador [71].
Table 1. Rainy and dry seasons for the three regions of continental Ecuador [71].
CoastHighlandsAmazon
Average annual air temperature (°C)25.512.721.8
Average annual rainfall (mm/year)892.9798.93449.4
Rainy seasonJanuary to AprilMar to April and October to NovemberAlmost the whole year 1
Dry seasonJune to DecemberMay to September and December to JanuaryAugust to January
1 Heavy rains occur from May to August.

2.2. Time Series of Vegetation Indices, Climate Information and Global Teleconnection Indices

2.2.1. NDVI Dataset

NDVI is an indicator of vegetation response that has been used in multiple ecosystem studies [44,72,73,74,75]. NDVI is sensitive to chlorophyll in leaves and is calculated in terms of reflectance at red and near-infrared wavelengths, according to Equation (1) [13].
NDVI = ρ N I R ρ r e d ρ N I R + ρ r e d
where ρ N I R and ρ r e d are the bidirectional surface reflectances within the MODIS bands [13]. An interpretation of the type of vegetation cover from a range of NDVI values is shown in Table 2.
The NDVI information for the Ecuadorian Páramos was obtained using the MODIS Collection 6 product MOD13Q1 available between 2001 and 2018. MOD13Q1 is a spatial, spectral and radiometric representation of surface vegetation conditions [77]. NDVI products were generated from MODIS data using the 16-day peak composite method and combined with the MODIS bidirectional reflectance distribution function to obtain a higher percentage of clear-sky data that minimizes noise. The MODIS images have a spatial resolution of 0.002° (~250 m) and were downloaded from the Distributed Active Archive Center (DAAC) Level-1 and Atmosphere Archive & Distribution System (LAADS; https://ladsweb.modaps.eosdis.nasa.gov/ (accessed on 4 March 2019)). Subsequently, these images were projected onto WGS84 using ArcGIS 10.4. In this ArcGIS analysis, it was possible to generate 414 mosaics and create a database for the Ecuadorian Páramos (h9v9, h10v8, h10v9) based on the Páramos shown on the land use coverage map at the national scale, developed by the Ministry of the Environment of Ecuador [78].
Consecutively, the Páramo mosaics were analyzed with the R-cran 3.5.3 software and the images were filtered with the reliability layer of the MOD13Q1 product to maintain only cells with good quality and eliminate negative values. Thus, an annual NDVI time series was created with a total of 414 values per cell corresponding to the 18 years of study as a result of the integration of 23 images per year (1 image every 16 days). Missing information was spatially and temporally filled in by pixels that showed at least 50% of the total temporal information. Spatial approximation was carried out using the Ordinary Kriging method, which generates a reliable performance in the prediction of missing information on the images [79]. To fill in the information with this method, the experimental variogram was calculated, a theoretical variogram was selected between Gaussian, exponential, pentaspherical or wave for each image, and the NDVI values were predicted on the missing spatial information. Finally, the temporal approximation was performed through the Savitzky—Golay filter within the TIMESAT 3.3 software [30,80].

2.2.2. Satellite Climate Information and Water Availability

The University of California, Santa Barbara—Climate Hazards Group InfraRed Precipitation with Station (UCSB-CHIRPS) project provides precipitation data with a spatial resolution of 0.01° (~1 km). This information is based on satellite infrared rainfall and rain gauge station observations [81]. Monthly precipitation data for the study area was downloaded from http://chg.geog.ucsb.edu/data/chirps/ (accessed on 25 Febrary 2019). These images were then stitched, cropped, and resampled (bilinear interpolation) at a spatial resolution of 250 m to match the extent and spatial resolution of NDVI. Precipitation images were then converted to ASCII files to create a data matrix.
Land Surface Temperature and Emissivity (LST&E) were derived from the MODIS-MOD11B3 product with a spatial resolution of 0.055° (~6 km). The information in this product is based on the split window method, where the result is an accurate soil surface temperature [82]. The monthly temperature information was downloaded from the portal https://lpdaac.usgs.gov/products/mod11b3v006/ (accessed on 16 August 2019), it was transformed into Celsius degrees and processed in the same way as precipitation.
Water discharge rates at specific points in the Andes were used as indicators of water availability in the region. Information on water availability for 12 high-altitude basins was provided by the INAMHI (Table 3). The Páramo ecosystem is the dominant land cover type within the selected hydrographic basins, with coverage exceeding 50% of its surface (Figure 1).

2.2.3. Global Teleconnection Indices

Planetary-scale waves are the product of atmospheric processes regulated by mountain ranges and the boundaries between land and water masses. These global phenomena generate anomalies in the climatic seasons and, in long-term periods, influence large geographic regions [83]. Several global climatic anomalies were selected, including the Antarctic Oscillation (AAO), ENSO Multivariate Index (MEI), and Madden–Julian Oscillation (MJO). Table 4 shows a summary of the teleconnection rates studied and the digital repositories where they were downloaded.

2.3. Methods

A NDVI spatio-temporal analysis was performed and summarized within 10 defined zones within the study area. The zones were divided with respect to their predominant terrain orientation and location, resulting in the salient climatic influence of the Pacific Ocean, Amazon, or a hybrid regime (Figure 3). These zones are defined by the summit lines of the western and eastern mountain ranges as well as by the inter-Andean alley delimited by both mountain ranges. In addition, a differentiation is made between the northern and southern parts of the Páramo, defined by latitude 2° S.

2.3.1. Spatio-Temporal Analysis of NDVI

A productivity indicator called Time Integration—NDVI (TI-NDVI) was applied in this study. TI-NDVI is defined as the area under the curve of the NDVI time series [84]. The TI-NDVI was calculated for each cell to describe the spatial distribution of the productivity of the tussock grasses. In addition, the variability of vegetation dynamics per cell was calculated using the standard deviation (Equation (2)). The NDVI anomalies were computed using the z-score (Equation (3)) to illustrate the spatial and temporal distributions. Additionally, the monthly anomalies were smoothed with a moving average of a 36-month average to partially remove the inter-annual variability of the tussock grasses. The moving average was calculated using Equation (4), where i corresponds to the month within the study period. The Mann–Kendall test [85,86] and its slope with 95% confidence were calculated on the NDVI time series to determine whether there was a spatial trend for each cell. The magnitude of the trend was estimated using Sen’s slope where a positive slope value indicates a monotonic increase or an increasing trend, while a negative value of the slope indicates a monotonic decrease or a decreasing trend [87]. The maximum value of TI-NDVI during a quarter analysis was developed to determine the period (December–January–Febrary: DJF; March–April–May: MAM; June–July–August: JJA; September–October–November: SON) where the maximum NDVI value generally occurs. This last analysis requires comparing the sum of the TI-NDVI within each quarter to select the period of greatest occurrence that presents the maximum annual productivity.
S D x , y = N D V I x , y , t M E A N N D V I N 1
Z   S C O R E x , y , t = N D V I x , y , t M E A N N D V I S D N D V I
M E A N Z   S C O R E i = i i + 36 Z   S C O R E i 36
where NDVI(x,y,t) is the NDVI value at a given position and time. MEAN(NDVI) is the mean NDVI value of the database. N is the number of months during the study period. SD(NDVI) is the standard deviation of the NDVI database. Z SCORE(x,y,t) is the time anomaly of each cell.

2.3.2. NDVI Analysis—Climatic Variables and Water Availability

Precipitation and soil temperature are the main variables controlling phenology and ecosystem production [12]. A Pearson correlation coefficient with 5% significance was calculated to understand the relationship between vegetation and climatic variables [88]. A simple and multivariate correlation was performed between NDVI and precipitation or soil temperature, and both climatic variables. The correlation was calculated at different time levels: monthly, bi-monthly, quarterly, four-monthly, half-yearly and annually. Similarly, water availability is directly related to vegetative growth to a greater or lesser extent. Therefore, a simple linear regression model was used to determine the relationship between the mean monthly discharge and mean monthly NDVI time series.

2.3.3. NDVI Analysis—Global Climate Indices

The Global Teleconnection Indices (GTI) allowed statistical analysis between them and the NDVI time series. Table 4 shows the GTIs that were selected to understand their influence on the grassland Páramos. Initially, the Pearson correlation with 5% significance between NDVI and GTI was calculated to later calculate the cross-correlation (p value < 0.05) with a delay of one period to the past of the GTI for the monthly, semi-annual and annual series.

3. Results

Tussock grasses are located along 9650 km2 in the western and royal mountain range between 3000 and 4000 m.a.s.l. (Figure 2). Sixty-eight percent of this vegetation drains towards the Amazon basin and the remaining 32% flows into the Pacific Ocean. A total of 1242 MODIS images were downloaded and processed from 2001 to 2018. After filtering the MODIS images and filling in the data gaps, only a quarter of the surface extent of the tussock Páramo ecosystem resulted in good quality NDVI data, corresponding to a total of 39,192 cells. Seventy-three percent (73.1%) of the total observations were in the northern part (north of 2° S latitude) and 26.9% in the southern part. Approximately fifty-one percent (51.3%) of all observations are in the Western Range and 41.5% in the Royal Range.

3.1. Spatio-Temporal Analysis of NDVI

Table 5 shows the mean elevation, mean annual precipitation, mean monthly ground surface temperature, mean values, median, standard deviation, and temporal trend analysis of the NDVI calculated and summarized for the defined zones within the study area (Figure 3). The average greenness level of the grasslands (Figure 1) shows that the grassland Páramo ecosystem of the Western Range and the northern Inter-Andean Valley are slightly greener than their corresponding ones to the south. On the other hand, the eastern mountain range shows the same greenness both to the north and to the south (4, 5, 9 and 10 zones).
The average zone variability of the tussock grasses greenness indicated by means of standard deviation can be seen in Table 6—Column 4. This shows little variability in greenness in general, where the grasslands located to the north are more variable than those in the south, with higher values in the Amazon basin. On the other hand, the southern grasslands showed the greatest variability in the western mountain range.
The mean zone z-score anomalies of the tussock grasses are shown in Table 6—Column 5. The negative (positive) anomalies represent areas of lower (higher) vegetation with respect to the mean of tussock grasses. Positive and negative anomalies are distributed in space by 46.5% and 56.5%, respectively. In the northern part of the grasslands, there are positive anomalies with the exception of the western mountain range (1 and 2 zones), which presents a negative and neutral anomaly for the western and eastern parts, respectively. In contrast, the southern tussock grasses in the western mountain-range (6 and 7 zones) have the largest negative anomalies, whereas the western part of the western mountain range being the dominant anomaly. On the contrary, the northern and southeastern mountain ranges (4, 5, 9 and 10 zones) present positive anomalies, with the northern part showing the highest productivity.
The monthly NDVI anomalies of the study area and the 36-month moving average are shown in Figure 4. The moving average shows the tendency of grasslands to be greener over time, with maximum anomalies of up to 0.15 SD. However, during the period 2009–2011, the grasslands experienced a decrease in vegetation greenness, values close to 0. Similarly, tussock grasses showed higher (lower) greenness in the face of La Niña (El Niño) events.
Table 6—Column 6 shows the mean Mann–Kendall trend test results for each zone with a significance level of 5% for the period 2001–2018. The Tau coefficient of the trend test ranges from −1 to 1 and is read as Spearman’s coefficient [89]. The trend results were discretized as described by Yan [90]. The trend test shows that 56.26% of the study area shows statistically significant trend results. For the entire study area, a trend of 5.75% negative and 50.51% positive was observed. Within all zones, there is a very weak positive trend. Of these, the greatest trends are found in the eastern mountain range, both in the northern and southern tussock grasses.
Figure 5 shows the quarter where the maximum NDVI productivity values of the tussock grasses usually occur. According to the summary of results presented in Table 6, it is observed that in general the grasslands present their maximum annual quarterly productivities in the MAM period, and only the southern inter-Andean Valley (zone 7) presents it in JJA.

3.2. NDVI Analysis—Climatic Variables and Water Availability

The CHIRPS product shows a distribution of 147–2214 mm/year of precipitation over the grassland Páramo ecosystem. The highest precipitation occurs over the Amazon basin around Antisana volcano (section 1) and towards the southwest of La Merced de Buenos Aires city at the northern end of the Western Range. The standard deviation of the precipitation is low over the entire study area. However, the highest SD values for precipitation are observed around Guagua Pichincha volcano. On the other hand, the soil temperature has a range of 6.2–25.0 °C where the highest values are shown in the lower limit of the grassland and the lowest are found around the peak flow line. In addition, the smallest standard deviations of soil temperature are found in the Pacific basin and the largest in the Amazon basin.
Table 7 shows the spatial average of the simple and multivariable correlations between NDVI and precipitation, temperature and both variables. It can also be seen in Table 7 the amount of significant data out of the total for each level of correlation (monthly, bimonthly, quarterly, four-monthly, half-yearly and annual). The six-month correlations show the strongest Pearson correlations in combination with a high percentage of significant data. In such a way, the six-monthly correlations have been averaged for each zone and are shown in Table 8.
Table 8—Column 6 summarizes the Pearson coefficients of the multivariate correlations between NDVI, precipitation, and temperature for the 10 zones. The correlations are weak (11.79%), moderate (17.68%), strong (15.06%) and very strong (0.70%). The grasslands, with the exception of the west of the western mountain range present a moderate positive correlation (weak positive correlation with precipitation). Negative correlations were obtained between NDVI and temperature. Above latitude 2° S, in zones 1,2, 4, there are moderate negative correlations, unlike the weak negative correlation of the inter-Andean alley and the east of eastern mountain range. The multivariate correlation of NDVI with precipitation and temperature shows moderate positive correlations, and in the case of the west of the western mountain range, close to a strong correlation. On the other hand, the number of significant correlations of the southern eastern mountain range in terms of temperature is negligible, so the results do not represent these areas.
There are clear differences in the water discharge in the hydrographic basins that drain the western and eastern mountain ranges (Figure 6). For example, basins located in the western mountain range showed increases in discharge around April and May and decreases around September and October. Some basins that drained the Western Range even exhibited a bimodal pattern (for example, H0064, H0793, H0333, H0337, and H0896). While the basins located in the eastern mountain range were characterized by a maximum value in the middle of the year (for example, H0722, H0788, H0787, H0789, and H0790). It was also evident that the water discharge in the Royal Range is greater than that in basins that drain the Western Range. The peak discharge and base flow in the basins draining the Royal Range doubled the peak discharge and base flow in the basins draining the Western Range (Figure 6).
The temporal variability of NDVI is closely associated with water availability, as indicated by the surface water discharge. Table 9 shows the Pearson correlation coefficients (R2) resulting from a linear regression model between the mean monthly surface water discharge and the mean monthly variability of the NDVI. There was a moderately explanation for the relationship between the discharge recorded in the western watersheds and the mean NDVI time series. The relationship was the weakest between the surface water discharge recorded in the eastern basins and the mean NDVI time series. One of the reasons is that the data set for both indices has more data available along the Western Range than along the Royal Range. Therefore, western VI values predominate in both datasets. Later, the mean monthly surface water discharge was correlated with the mean NDVI time series for each group indicated by the blue line in Figure 1.

3.3. NDVI Analysis—Global Climate Indices

In the same way as the analysis between NDVI and climatic variables, the Global Teleconnection Indices have been analyzed at different levels of temporal information (monthly, bimonthly, quarterly, four-monthly, half-yearly, and annual). From this analysis, the results at the semester level present the highest correlations in combination with the amount of significant information. The semi-annual Pearson correlation between the NDVI and GTI is summarized in Table 10, where the AAO, MJO and NAO indices have the lowest percentages of significant data. Of these indices, MJO presents higher average correlation coefficients than the other indices. On the other hand, the cross-correlation analysis of a GTI lag for the monthly, semi-annual and annual series has lower Pearson coefficients than the analysis without time shift.

4. Discussion

4.1. Spatio-Temporal Analysis of NDVI

The analysis shows that tussock grasses have a similar average behavior throughout the study area with grasslands above 2° S latitude showing higher productivity, higher precipitation, and lower surface temperature. The mean, median, z–score and TI–NDVI indicate that the northern inter-Andean alley (zone 3) is the most productive zone, followed by the Royal Range (Table 6). Zone 3 high productivity is a consequence of grassland response to rainfall, water availability, and soil temperature (Table 5). Another variable that can contribute to productivity is elevation as has been indicated by [91,92].
The seasonality of the average productivity of the NDVI generally presents its maximum peak in the period March–April–May, corresponding to the first and most intense cycle of precipitation in the region [43,93]. On the other hand, the maximum annual productivity of the Southern Inter-Andean Valley (zone 8) occurs in June–July–August, corresponding to the summer (winter) of the Ecuadorian highlands (Amazon). Zone 8 can also be affected by the humidity of the Amazon, which presents an increase in its precipitation, which in turn is influenced by the Atlantic Ocean [4]. In other words, the grasslands of zone 8 have a relationship with the Amazon and the surface temperature of the Atlantic Ocean [4,94].
The NDVI variability (Table 6) shows that zones above 2° S latitude are more variable than their southern counterparts. Of all the zones, zone 5 has the highest variability (0.078) in the study area, which also corresponds to the highest variability of precipitation (186.3 mm/year) and soil temperature (4.0 °C) in the study area (Table 5). These results are in agreement with those of studies on the Royal Range in Colombia, which show the greatest variation in NDVI is linked to variations in precipitation and air temperature. On the other hand, the tussock grasses of zones 7 to 10 show less variability, indicating the influence of moisture transported from the Amazon [95]. Unlike zone 6, which has a similar variability with its counterpart to the north (zone 1) and shows no influence of humidity from the Amazon due to the blockade of the Western Range [96].
The spatial and temporal anomalies of NDVI (Table 6) shown that the greatest anomalies occur in the tussock grasses of the northern areas and of these, the anomalies of the Western Range are positive and for the Royal Range, positive. In addition, these results are in agreement with those obtained in the Mann–Kendall and Sen’s Slope analysis, which show the Royal Range with the largest positive trends. That is, the anomalies and trends show that within the 2001–2018 period, there is an increase in the greenness of the tussock grasses on the Royal Range due to a possible effect of CO2 fertilization [97,98]. On the other hand, it is observed that zones 3 and 8 have the lowest greening tendencies (zone 3: τ = 0.09; zone 8: τ = 0.08) together with the highest positive anomaly and a slight negative anomaly (zone 3: z-score = 0.26; zone 8: z-score = −0.05), respectively. This observation can be interpreted as the maximum extent of productivity in these areas and the small positive trends may be the response to global CO2 fertilization [98]. Finally, in a peculiar way, it can be pointed out that the anomalies of Figure 4 shows negative anomalies (−0.1 SD) in the period 2001–2004 throughout the study area that corresponds to the event of extreme drought recorded in this period in Quito, the capital city [94].

4.2. NDVI Analysis—Climatic Variables

The results of the linear and multilinear Pearson correlations at different NDVI temporal levels with precipitation, temperature, and water availability show only a partial explanation of the behavior of tussock grasses. It was observed that vegetation has positive linear correlations with precipitation, negative with temperature and high R-squared with discharge flows. In addition, the multilinear correlation of NDVI with precipitation and temperature show higher Pearson coefficients than the linear correlations with these individual variables. The results show that the vegetation correlations at large temporal scales (6 and 12 months) have higher average coefficients, except for the linear correlation of NDVI and precipitation at 12 months (Table 7). These results are confirmed by the study by Vega-Jácome [99] on the Peruvian Andes, which shows higher correlations at the same time scales. Zones 1 and 6 showed the highest Pearson coefficients and the highest rates of significance data (Table 8), which are in agreement with the results obtained in wetlands in Peru [100]. On the other hand, it has been observed that NDVI has a better correlation with evaporation [101], the high mountain vegetation in Peru has a greater influence of water availability, temperature, cloudiness, solar radiation, slope angle and elevation [102,103,104], and the biomass of the Antisana grasslands is inversely related to elevation [91].
Laraque et al. [105] showed the spatial regimes of discharge in Ecuador are related to precipitation patterns. Of the two main slopes of the Andes, the slope towards the Amazon has fewer fluctuations in water flow [106]. In addition, the Pacific slope during the first half of the year shows greater water availability, whereas the Amazon slope has peak flows in the middle of the year. The R-squared between the monthly mean NDVI and monthly mean discharge shows no linear relationships. These relationships are different from the correlations of NDVI and precipitation, which show higher correlation coefficients in the Western Range. Likewise, the coefficients of determination between discharge and NDVI show, that at the time of increasing discharge (April–May, peak rainfall), the level of greenness also increases [105,107]. Thus, water availability is a dominant factor in the greenness of grasslands, as corroborated by the results of the high Andean vegetation [99,100].
Extensive studies have been conducted on different types of vegetation and their relationship with temperature on a global scale. There is a relationship between the root, stem, and leaf biomass [108]. Other studies on trees located at elevations higher than ~1700 m.a.s.l. showed a positive trend between their diameter and temperature [109]. In the case of Ecuadorian tussock grasses, the relationship with temperature is negative (weak and moderate), whereas greenness decreases as the temperature rises. For both precipitation and water availability, the greatest relationships were observed throughout the Western Range.

4.3. NDVI Analysis—Global Climate Indices

Various studies on the Andes have confirmed that climatic variables are directly or indirectly linked to global indices [74,110,111,112,113,114,115,116,117,118,119]. The study correlated greenness with various global climate indices, among which the North Atlantic Oscillation (NAO) had the highest Pearson coefficients. This index is related to the trade winds that originate in the north tropical Atlantic Ocean, which previous studies have concluded to be strongly correlated with precipitation in Colombia [112,117]. The results showed an increase in the average Pearson coefficients in the direction of the Pacific Ocean in correspondence with the direction of the trade winds.
On the other hand, all the indices that are based on sea surface temperature of the Pacific Ocean (ENSO, MEI, PDO, El Niño 1 + 2, 3, 3.4 and 4) showed their main incidence in the climatology of the Colombian coasts, Ecuador and Peru [2,3,4,44,115,120,121,122,123,124]. Of these indices, PDO showed weak correlations with NDVI. Studies have determined that the temperature of the Andes is associated with the positive phase of the P DO [125] and ENSO events [113,114,115,116]. In addition, the strength of these ENSO events is influenced by the phases of the PDO [74,111], and coastal (Andean) precipitation is linked to El Niño 1 + 2 (El Niño 3.4) [115,124]. Similarly, the Ecuadorian tussock grasses showed a slight response to ENSO events (Figure 4), corroborating the results observed from NDVI in South America and Colombia [43,44] and refuting those obtained in Ecuador and Peru [110,126]. Subsequently, it was observed that the tussock grasses around snow-capped volcanoes have positive NDVI anomalies that are linked to the warm phase of ENSO (March–April), where a negative balance is produced in the glaciers, which promotes their melting and increases the availability of water; thus, the greenness in the tussock grasses increases (Figure 1) [127].
Finally, the Antarctic low-pressure westerly winds measured by the AAO showed weak correlations with NDVI and a small number of cells with significant data, of which the largest are over the Royal Range, corroborating the results of Leeuwen et al. [32]. In contrast, the MJO is considered as one of the main modulators of climatic variability in the tropics [128]. Studies in the Colombian Andes have determined that Rossby and Kelvin waves are linked to precipitation and temperature [119,123,129]. However, the results showed that NDVI not only shows weak correlations with this index, but also that the Ecuadorian Andes do not correlate significantly with the MJO.

5. Conclusions

The study examined the spatial and temporal variability of NDVI in 10 equatorial tussock grassland areas. The grasslands displayed consistent characteristics, with the inter-Andean Valley above 2° S latitude, and the Royal Range being the most productive. Peak productivity coincided with maximum rainfall (MAM). In zone 8 (southern inter-Andean Valley), productivity responded to Amazon moisture influenced by Atlantic Ocean temperature, with a peak during JJA.
The NDVI of the Ecuadorian tussock grasslands showed correlations with precipitation and soil temperature, and multivariate analysis revealed stronger associations. Higher precipitation and water availability corresponded to increased greenness, whereas higher soil temperatures were associated with lower NDVI. Anomalies exhibited a northeasterly increase, with negative anomalies in the Western Range owing to the topographical barrier of the Royal Range. The Royal Range had the strongest positive tendencies, influenced by moisture from the Amazonian region. Positive trends were observed throughout the study area, and potentially linked to global CO2 fertilization. The linear relationships between NDVI and teleconnection indices were generally weak, with NAO demonstrating the highest ratios. The grasslands showed a slight response to El Niño and La Niña events, warranting further investigation using non-linear models.

Author Contributions

Conceptualization, X.Z.-R.; Methodology, J.V.-V. and X.Z.-R.; Validation, J.V.-V.; Formal analysis, J.V.-V., K.U.-Z. and C.B.-E.; Investigation, J.V.-V. and X.Z.-R.; Writing—original draft, J.V.-V. and X.Z.-R.; Writing—review & editing, J.V.-V. and X.Z.-R.; Supervision, X.Z.-R.; Project administration, X.Z.-R. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support provided by the Escuela Politécnica Nacional (National Polytechnic School) for the development of the project PIJ 17-05: “Los patrones climáticosglobales y su influencia en la respuesta temporal y espacial de índices espectrales de la vegetación delpáramo en el Ecuador” (Global climate patterns and their influence on temporal and spatial responses ofthe Ecuadorian Páramo vegetation’s spectral indices).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are openly available in the following data repositories in: NDVI—MODIS: https://ladsweb.modaps.eosdis.nasa.gov/ (accessed on 4 March 2019), UCSB-CHIRPS: http://chg.geog.ucsb.edu/data/chirps/ (accessed on 25 Febrary 2019), LST&E-MODIS: https://lpdaac.usgs.gov/products/mod11b3v006/ (accessed on 16 August 2019), AAO: www.cpc.ncep.noaa.gov/products/precip/Cwlink/daily_ao_index/aao/ (accessed on 21 February 2020), MEI: psl.noaa.gov/enso/mei/ (accessed on 21 February 2020), MJO: www.cpc.ncep.noaa.gov/products/precip/CWlink/daily_mjo_index/proj_norm_order.ascii (accessed on 30 January 2020), NAO: www.cpc.ncep.noaa.gov/products/precip/CWlink/pna/nao.shtml (accessed on 23 January 2020), PDO: www.ncdc.noaa.gov/teleconnections/pdo/ (accessed on 24 January 2020), El Niño 1 + 2: psl.noaa.gov/data/correlation/nina1.data (accessed on 22 January 2020), El Niño 3: psl.noaa.gov/data/correlation/nina3.data (accessed on 22 January 2020), El Niño 3: psl.noaa.gov/data/correlation/nina4.data (accessed on 22 January 2020), El Niño 3.4: psl.noaa.gov/data/correlation/nina34.data (accessed on 22 January 2020).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lauer, W. Ecoclimatological Conditions of the Paramo Belt in the Tropical High Mountains. Int. Mt. Soc. Stable 1981, 1, 209–221. [Google Scholar] [CrossRef]
  2. Garreaud, R.D.; Vuille, M.; Compagnucci, R.; Marengo, J. Present-Day South American Climate. Palaeogeogr. Palaeoclim. Palaeoecol. 2009, 281, 180–195. [Google Scholar] [CrossRef]
  3. Harden, C.P. Human Impacts on Headwater Fluvial Systems in the Northern and Central Andes. Geomorphology 2006, 79, 249–263. [Google Scholar] [CrossRef]
  4. Vuille, M.; Bradley, R.S.; Keimig, F. Climate Variability in the Andes of Ecuador and Its Relation to Tropical Pacific and Atlantic Sea Surface Temperature Anomalies. J. Clim. 2000, 13, 2520–2535. [Google Scholar] [CrossRef]
  5. Jørgensen, P.M.; León-Yánez, S. Catalogue of the Vascular Plants of Ecuador; Missouri Botanical Garden Press: St. Louis, MO, USA, 1999; Volume 75. [Google Scholar]
  6. Pourrut, P.; Pouyaud, B. El Agua En El Ecuador: Clima, Precipitaciones, Escorrentía; Corporación Editora Nacional: Quito, Ecuador, 1995. [Google Scholar]
  7. Junquas, C.; Heredia, M.B.; Condom, T.; Ruiz-Hernández, J.C.; Campozano, L.; Dudhia, J.; Espinoza, J.C.; Menegoz, M.; Rabatel, A.; Sicart, J.E. Regional Climate Modeling of the Diurnal Cycle of Precipitation and Associated Atmospheric Circulation Patterns over an Andean Glacier Region (Antisana, Ecuador). Clim. Dyn. 2022, 58, 3075–3104. [Google Scholar] [CrossRef]
  8. Buytaert, W.; de Bievre, B. Water for Cities: The Impact of Climate Change and Demographic Growth in the Tropical Andes. Water Resour. Res. 2012, 48. [Google Scholar] [CrossRef] [Green Version]
  9. Viviroli, D.; Archer, D.R.; Buytaert, W.; Fowler, H.J.; Greenwood, G.B.; Hamlet, A.F.; Huang, Y.; Koboltschnig, G.; Litaor, M.I.; López-Moreno, J.I.; et al. Climate Change and Mountain Water Resources: Overview and Recommendations for Research, Management and Policy. Hydrol. Earth Syst. Sci. 2011, 15, 471–504. [Google Scholar] [CrossRef] [Green Version]
  10. Hamel, P.; Riveros-Iregui, D.; Ballari, D.; Browning, T.; Célleri, R.; Chandler, D.; Chun, K.P.; Destouni, G.; Jacobs, S.; Jasechko, S.; et al. Watershed Services in the Humid Tropics: Opportunities from Recent Advances in Ecohydrology. Ecohydrology 2018, 11, e1921. [Google Scholar] [CrossRef] [Green Version]
  11. Ponette-González, A.G.; Marín-Spiotta, E.; Brauman, K.A.; Farley, K.A.; Weathers, K.C.; Young, K.R. Hydrologic Connectivity in the High-Elevation Tropics: Heterogeneous Responses to Land Change. Bioscience 2014, 64, 92–104. [Google Scholar]
  12. Appenzeller, T. Fire on the Mountain. Science 2019, 365, 1094–1097. [Google Scholar] [CrossRef]
  13. Huete, A.; Didan, K.; Miura, T.; Rodriguez, E.P.; Gao, X.; Ferreira, L.G. Overview of the Radiometric and Biophysical Performance of the MODIS Vegetation Indices. Remote Sens. Environ. 2002, 83, 195–213. [Google Scholar] [CrossRef]
  14. Justice, C.; Townshend, J.; Holben, A.; Tucker, C. Analysis of the Phenology of Global Vegetation Using Meteorological Satellite Data. Int. J. Remote Sens. 1985, 6, 1271–1318. [Google Scholar] [CrossRef] [Green Version]
  15. Tucker, C.J.; Sellers, P.J. Satellite Remote Sensing of Primary Production. Int. J. Remote Sens. 1986, 7, 1395–1416. [Google Scholar] [CrossRef] [Green Version]
  16. Delire, C.; de Noblet-Ducoudré, N.; Sima, A.; Gouirand, I. Vegetation Dynamics Enhancing Long-Term Climate Variability Confirmed by Two Models. J. Clim. 2011, 24, 2238–2257. [Google Scholar] [CrossRef]
  17. Li, Z.; Huffman, T.; McConkey, B.; Townley-Smith, L. Monitoring and Modeling Spatial and Temporal Patterns of Grassland Dynamics Using Time-Series MODIS NDVI with Climate and Stocking Data. Remote Sens. Environ. 2013, 138, 232–244. [Google Scholar] [CrossRef]
  18. Myneni, R.B.; Hall, F.G.; Sellers, P.J.; Marshak, A.L. Interpretation of Spectral Vegetation Indexes. IEEE Trans. Geosci. Remote Sens. 1995, 33, 481–486. [Google Scholar] [CrossRef]
  19. Paruelo, J.M.; Epstein, H.E.; Lauenroth, W.K.; Burke, I.C. ANPP Estimates from NDVI for the Central Grassland Region of the United States. Ecology 1997, 78, 953–958. [Google Scholar] [CrossRef]
  20. Gamon, J.A.; Field, C.B.; Goulden, M.L.; Griffin, K.L.; Hartley, A.E.; Joel, G.; Penuelas, J.; Valentini, R. Relationships between NDVI, Canopy Structure, and Photosynthesis in Three Californian Vegetation Types. Ecol. Appl. 1995, 5, 28–41. [Google Scholar] [CrossRef] [Green Version]
  21. Arboit, M.E.; Maglione, D.S. Categorización de Las Manzanas Urbanas Para La Integración de La Silvicultura Urbana En La Planificación de Las Ciudades. Caso de Estudio: Área Metropolitana de Mendoza, Argentina. III Congr. Int. Isuf-H. Ciudad. Compact. Vs. Ciudad. Difusa 2020, 297–305. [Google Scholar] [CrossRef]
  22. Burgheimer, J.; Wilske, B.; Maseyk, K.; Karnieli, A.; Zaady, E.; Yakir, D.; Kesselmeier, J. Relationships between Normalized Difference Vegetation Index (NDVI) and Carbon Fluxes of Biologic Soil Crusts Assessed by Ground Measurements. J. Arid. Environ. 2006, 64, 651–669. [Google Scholar] [CrossRef]
  23. Kunkel, M.L.; Flores, A.N.; Smith, T.J.; McNamara, J.P.; Benner, S.G. A Simplified Approach for Estimating Soil Carbon and Nitrogen Stocks in Semi-Arid Complex Terrain. Geoderma 2011, 165, 1–11. [Google Scholar] [CrossRef] [Green Version]
  24. Gu, Y.; Brown, J.F.; Verdin, J.P.; Wardlow, B. A Five-Year Analysis of MODIS NDVI and NDWI for Grassland Drought Assessment over the Central Great Plains of the United States. Geophys. Res. Lett. 2007, 34. [Google Scholar] [CrossRef] [Green Version]
  25. Dayeh, M.A.; Desai, M.I.; Dwyer, J.R.; Rassoul, H.K.; Mason, G.M.; Mazur, J.E. Composition and Spectral Properties of the 1 Au Quiet-Time Suprathermal Ion Population during Solar Cycle 23. Astrophys. J. 2009, 693, 1588–1600. [Google Scholar] [CrossRef] [Green Version]
  26. Mewaldt, R.A.; Looper, M.D.; Cohen, C.M.S.; Haggerty, D.K.; Labrador, A.W.; Leske, R.A.; Mason, G.M.; Mazur, J.E.; von Rosenvinge, T.T. Energy Spectra, Composition, and Other Properties of Ground-Level Events during Solar Cycle 23. Space Sci. Rev. 2012, 171, 97–120. [Google Scholar] [CrossRef]
  27. Schweiger, A.K.; Cavender-Bares, J.; Townsend, P.A.; Hobbie, S.E.; Madritch, M.D.; Wang, R.; Tilman, D.; Gamon, J.A. Plant Spectral Diversity Integrates Functional and Phylogenetic Components of Biodiversity and Predicts Ecosystem Function. Nat. Ecol. Evolume 2018, 2, 976–982. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, K.; Franklin, S.E.; Guo, X.; Cattet, M. Remote Sensing of Ecology, Biodiversity and Conservation: A Review from the Perspective of Remote Sensing Specialists. Sensors 2010, 10, 9647–9667. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, R.; Gamon, J.A.; Cavender-Bares, J.; Townsend, P.A.; Zygielbaum, A.I. The Spatial Sensitivity of the Spectral Diversity-Biodiversity Relationship: An Experimental Test in a Prairie Grassland. Ecol. Appl. 2018, 28, 541–556. [Google Scholar] [CrossRef] [Green Version]
  30. Eklundh, L.; Jönsson, P. TIMESAT for Processing Time-Series Data from Satellite Sensors for Land Surface Monitoring. In Multitemporal Remote Sensing; Ban, Y., Ed.; Springer: Cham, Switzerland, 2016; Volume 20. [Google Scholar] [CrossRef]
  31. Moreira, A.; Bremm, C.; Fontana, D.C.; Kuplich, T.M. Seasonal Dynamics of Vegetation Indices as a Criterion for Grouping Grassland Typologies. Sci. Agricola 2019, 76, 24–32. [Google Scholar] [CrossRef]
  32. van Leeuwen, W.J.D.; Hartfield, K.; Miranda, M.; Meza, F.J. Trends and ENSO/AAO Driven Variability in NDVI Derived Productivity and Phenology alongside the Andes Mountains. Remote Sens. 2013, 5, 1177–1203. [Google Scholar] [CrossRef] [Green Version]
  33. Walther, G.R.; Post, E.; Convey, P.; Menzel, A.; Parmesan, C.; Beebee, T.J.C.; Fromentin, J.M.; Hoegh-Guldberg, O.; Bairlein, F. Ecological Responses to Recent Climate Change. Nature 2002, 416, 389–395. [Google Scholar] [CrossRef]
  34. Li, H.; Liu, L.; Liu, X.; Li, X.; Xu, Z. Greening Implication Inferred from Vegetation Dynamics Interacted with Climate Change and Human Activities over the Southeast Qinghai-Tibet Plateau. Remote Sens. 2019, 11, 2421. [Google Scholar] [CrossRef] [Green Version]
  35. IPCC. Climate Change 2007: Impacts, Adaptation and Vulnerability: Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel; Cambridge University Press: Cambridge, UK, 2007. [Google Scholar]
  36. Roerink, G.J.; Menenti, M.; Soepboer, W.; Su, Z. Assessment of Climate Impact on Vegetation Dynamics by Using Remote Sensing. Phys. Chem. Earth 2003, 28, 103–109. [Google Scholar] [CrossRef]
  37. Cleland, E.E.; Chiariello, N.R.; Loarie, S.R.; Mooney, H.A.; Field, C.B. Diverse Responses of Phenology to Global Changes in a Grassland Ecosystem. Proc. Natl. Acad. Sci. USA 2006, 103, 13740–13744. [Google Scholar] [CrossRef] [PubMed]
  38. Myneni, R.B.; Keeling, C.D.; Tucker, C.J.; Asrar, G.; Nemani, R.R. Increased Plant Growth in the Northern High Latitudes from 1981 to 1991. Nature 1997, 386, 698–702. [Google Scholar] [CrossRef]
  39. Schwartz, M.D.; Ahas, R.; Aasa, A. Onset of Spring Starting Earlier across the Northern Hemisphere. Glob. Chang. Biol. 2006, 12, 343–351. [Google Scholar] [CrossRef]
  40. Justice, C.; Vermote, E.; Townshend, J.; Defries, R.; Roy, D.; Hall, D.; Salomonson, V.; Privette, J.; Riggs, G.; Strahler, A.; et al. The Moderate Resolution Imaging Spectroradiometer (MODIS): Land Remote Sensing for Global Change Research. IEEE Trans. Geosci. Remote Sens. 1998, 36, 1228–1249. [Google Scholar] [CrossRef] [Green Version]
  41. Zhang, X.; Friedl, M.A.; Schaaf, C.B. Global Vegetation Phenology from Moderate Resolution Imaging Spectroradiometer (MODIS): Evaluation of Global Patterns and Comparison with in Situ Measurements. J. Geophys. Res. Biogeosci. 2006, 111. [Google Scholar] [CrossRef]
  42. Li, Z.; Li, X.; Wei, D.; Xu, X.; Wang, H. An Assessment of Correlation on MODIS-NDVI and EVI with Natural Vegetation Coverage in Northern Hebei Province, China. Procedia Environ. Sci. 2010, 2, 964–969. [Google Scholar] [CrossRef] [Green Version]
  43. Viña, A.; Henebry, G.M. Spatio-Temporal Change Analysis to Identify Anomalous Variation in the Vegetated Land Surface: ENSO Effects in Tropical South America. Geophys. Res. Lett. 2005, 32. [Google Scholar] [CrossRef] [Green Version]
  44. Poveda, G.; Álvarez, D.M.; Rueda, Ó.A. Hydro-Climatic Variability over the Andes of Colombia Associated with ENSO: A Review of Climatic Processes and Their Impact on One of the Earth’s Most Important Biodiversity Hotspots. Clim. Dyn. 2010, 36, 2233–2249. [Google Scholar] [CrossRef]
  45. Silvia, L.; Alexander, T.; Anna, K.; Polina, K. Assessment of carbon dynamics in Ecuadorian forests through the Mathematical Spatial Model of Global Carbon Cycle and the Normalized Differential Vegetation Index (NDVI). E3S Web Conf. 2019, 96, 02002. [Google Scholar] [CrossRef] [Green Version]
  46. Chamorro Sevilla, H.E.; Erazo, A. Estudio Multiespectral Del Cultivo de Tuna Para Determinar Los Índices NDVI, CWSI y SAVI, a Partir de Imágenes SENTINEL 2A, En El Cantón Guano, Provincia de Chimborazo, Ecuador. Enfoque UTE 2019, 10, 55–66. [Google Scholar] [CrossRef]
  47. Myers, N.; Mittermeler, R.A.; Mittermeler, C.G.; da Fonseca, G.A.B.; Kent, J. Biodiversity Hotspots for Conservation Priorities. Nature 2000, 403, 853–858. [Google Scholar] [CrossRef] [PubMed]
  48. Mena, P.; Hofstede, R. The Ecuadorian Páramos. Botánica Económica de Los Andes Centrales; Universidad Mayor de San Andrés: La Paz, Bolivia, 2006. [Google Scholar]
  49. Madriñán, S.; Cortés, A.J.; Richardson, J.E. Páramo Is the World’s Fastest Evolving and Coolest Biodiversity Hotspot. Front. Genet. 2013, 4, 192. [Google Scholar] [CrossRef] [Green Version]
  50. Hofstede, R. Los Servicios Del Ecosistema Páramo: Una Visión Desde La Evaluación de Ecosistemas Del Milenio; EcoCiencia: Quito, Ecuador, 2008. [Google Scholar]
  51. Morales, J.A.; Estévez, J.V. El Páramo: ¿Ecosistema En Vía De Extinción? Rev. Luna Azul 2006, 22, 39–51. [Google Scholar]
  52. Buytaert, W.; Wyseure, G.; de Bièvre, B.; Deckers, J. The Effect of Land-Use Changes on the Hydrological Behaviour of Histic Andosols in South Ecuador. Hydrol. Process. 2005, 19, 3985–3997. [Google Scholar] [CrossRef]
  53. Grau, R.; Aide, M. Globalization and Land-Use Transitions in Latin America. Ecol. Soc. 2008, 13. [Google Scholar] [CrossRef] [Green Version]
  54. Wang, J.; Rich, P.M.; Price, K.P. Temporal Responses of NDVI to Precipitation and Temperature in the Central Great Plains, USA. Int. J. Remote Sens. 2003, 24, 2345–2364. [Google Scholar] [CrossRef]
  55. Zhang, X.; Friedl, M.A.; Schaaf, C.B.; Strahler, A.H.; Hodges, J.C.F.; Gao, F.; Reed, B.C.; Huete, A. Monitoring Vegetation Phenology Using MODIS. Remote Sens. Environ. 2003, 84, 471–475. [Google Scholar] [CrossRef]
  56. Duan, S.B.; Li, Z.L.; Li, H.; Göttsche, F.M.; Wu, H.; Zhao, W.; Leng, P.; Zhang, X.; Coll, C. Validation of Collection 6 MODIS Land Surface Temperature Product Using in Situ Measurements. Remote Sens. Environ. 2019, 225, 16–29. [Google Scholar] [CrossRef] [Green Version]
  57. Krakauer, N.Y.; Pradhanang, S.M.; Lakhankar, T.; Jha, A.K. Evaluating Satellite Products for Precipitation Estimation in Mountain Regions: A Case Study for Nepal. Remote Sens. 2013, 5, 4107–4123. [Google Scholar] [CrossRef] [Green Version]
  58. Kalisa, W.; Igbawua, T.; Henchiri, M.; Ali, S.; Zhang, S.; Bai, Y.; Zhang, J. Assessment of Climate Impact on Vegetation Dynamics over East Africa from 1982 to 2015. Sci. Rep. 2019, 9, 16865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Kaser, G. Glacier-Climate Interaction at Low Latitudes. J. Glaciol. 2001, 47, 195–204. [Google Scholar] [CrossRef] [Green Version]
  60. Thompson, L.G. Ice Core Evidence for Climate Change in the Tropics: Implications for Our Future. Quat. Sci. Rev. 2000, 19, 19–35. [Google Scholar] [CrossRef]
  61. Jomelli, V.; Cooley, D.; Naveau, P.; Rabatel, A. The Little Ice Age in the Tropical Andes. In Proceedings of the American Geophysical Union: Fall 2003 Conference, San Francisco, USA, 1 December 2003. PP51A-07. [Google Scholar]
  62. Cerón, C.; Palacios, W.; Valencia, R.; Sierra, R. Las Formaciones Naturales de La Amazonía Del Ecuador. In Propuesta Preliminar de un Sistema de Clasificación de Vegetación para el Ecuador Continental; Sierra, R., Ed.; Editorial Rimana: Quito, Ecuador, 1999; pp. 111–119. [Google Scholar]
  63. INEC. Encuesta de Superficie y Producción Agropecuaria Continua; Independent National Electoral Commission: Ilocos Norte, Philippines, 2018. [Google Scholar]
  64. Vallejo, M.C.; Sacher, W. Encyclopedia of Mineral and Energy Policy; Tiess, G., Majumder, T., Cameron, P., Eds.; Springer: Berlin/Heidelberg, Germany, 2017. [Google Scholar] [CrossRef]
  65. Standmuller, T. Cloud Forests in the Humid Tropics: A Bibliographic Review; United Nations University Press: Tokyo, Japan, 1986. [Google Scholar]
  66. Young, B.E.; Young, K.R.; Josse, C. Vulnerability of Tropical Andean Ecosystems to Climate Change. Clim. Chang. Biodivers. Trop. Andes 2011, 1, 12. [Google Scholar]
  67. Ramsay, P.M. The Paramo Vegetation of Ecuador: The Community Ecology, Dynamics and Productivity of Tropical Grasslands in the Andes; Bangor University: Bangor, UK, 1992. [Google Scholar]
  68. Buytaert, W.; Iñiguez, V.; Celleri, R.; de Bièvre, B.; Wyseure, G.; Deckers, J. Analysis of the Water Balance of Small Páramo Catchments in South Ecuador. Environ. Role Wetl. Headwaters 2006, 63, 271–281. [Google Scholar] [CrossRef]
  69. Farley, J.; Aquino, A.; Daniels, A.; Moulaert, A.; Lee, D.; Krause, A. Global Mechanisms for Sustaining and Enhancing PES Schemes. Ecol. Econ. 2010, 69, 2075–2084. [Google Scholar] [CrossRef]
  70. Harling, G. The Vegetation Types of Ecuador: A Brief Survey. In Tropical Botany; Larsen, K., Holm-Nielsen, L.B., Eds.; Academic Press: London, UK, 1979; pp. 165–174. [Google Scholar]
  71. Pourrut, P.; Acosta, J.; Winckell, A.; Rodriguez, J. Los Climas Del Ecuador; Centro Ecuatoriano de Investigación Geográfica: Quito, Ecuador, 1983. [Google Scholar]
  72. Alcaraz-Segura, D.; Baldi, G.; Durante, P.; Garbulsky, M.F. Análisis de La Dinámica Temporal Del NDVI En Áreas Protegidas: Tres Casos de Estudio a Distintas Escalas Espaciales, Temporales y de Gestión. Ecosistemas 2008, 17, 108–117. [Google Scholar]
  73. Tiedemann, J.; Zerda, H.; Grilli, M.; Ravelo, A. Distribución Espacial de Anomalías Del NDVI Derivado Del Sensor VEGETATION SPOT 4/5 y Su Relación Con Las Coberturas Vegetales, Usos de La Tierra y Características Geomorfológicas En La Provincia de Santiago Del Estero, Argentina. Rev. Ambiência 2010, 6, 379–391. [Google Scholar] [CrossRef] [Green Version]
  74. Veettil, B.K.; Pereira, S.F.R.; Wang, S.; Valente, P.T.; Grondona, A.E.B.; Rondón, A.C.B.; Rekowsky, I.C.; de Souza, S.F.; Bianchini, N.; Bremer, U.F.; et al. Un Análisis Comparativo Del Comportamiento Diferencial de Los Glaciares En Los Andes Tropicales Usando Teledetección. Investig. Geográficas 2016, 3–36. [Google Scholar] [CrossRef] [Green Version]
  75. Zoffoli, M.L.; Madanes, N.; Kandus, P. Contribución de Series Temporales de NDVI NOAA/AVHRR al Análisis Funcional En Humedales. In Proceedings of the Anais XIII Simposio Brasileiro de Sensoramiento Remoto, Florianópolis, Brasil, 21–26 Apri1 2007. [Google Scholar]
  76. Manan, P.; Shital, S.; Kalubarme, M.H. Impact of Climate Change and Drought Analysis on Agriculture In Sabarkantha Districtusing Geoinformatics Technology. Glob. J. Eng. Sci. Res. 2019, 6, 133–144. [Google Scholar]
  77. Tucker, C.J.; Pinzon, J.E.; Brown, M.E.; Slayback, D.A.; Pak, E.W.; Mahoney, R.; Vermote, E.F.; el Saleous, N. An Extended AVHRR 8-Km NDVI Dataset Compatible with MODIS and SPOT Vegetation NDVI Data. Int. J. Remote Sens. 2005, 26, 4485–4498. [Google Scholar] [CrossRef]
  78. MAE SIstema de Clasificación de Los Ecosistemas Del Ecuador Continental; Ministerio del Ambiente: Magdalena del Mar, Peru, 2013.
  79. Jiang, X.; Was, L.; Du, Q.; Hu, B.X. Estimation of NDVI Images Using Geostatistical Methods. Earth Sci. Front. 2008, 15, 71–80. [Google Scholar] [CrossRef]
  80. Schafer, R.W. What Is a Savitzky-Golay Filter? IEEE Signal. Process. Mag. 2011, 28, 111–117. [Google Scholar] [CrossRef]
  81. Funk, C.; Peterson, P.; Landsfeld, M.; Pedreros, D.; Verdin, J.; Shukla, S.; Husak, G.; Rowland, J.; Harrison, L.; Hoell, A.; et al. The Climate Hazards Infrared Precipitation with Stations—A New Environmental Record for Monitoring Extremes. Sci. Data 2015, 2, 150066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Wan, Z. A Generalized Split-Window Algorithm for Retrieving Land-Surface Temperature from Space. IEEE Trans. Geosci. Remote Sens. 1996, 34, 892–905. [Google Scholar] [CrossRef] [Green Version]
  83. Stenseth, N.C.; Ottersen, G.; Hurrell, J.W.; Mysterud, A.; Lima, M.; Chan, K.S.; Yoccoz, N.G.; Ådlandsvik, B. Studying Climate Effects on Ecology through the Use of Climate Indices: The North Atlantic Oscillation, El Niño Southern Oscillation and Beyond. Proc. R. Soc. B Biol. Sci. 2003, 270, 2087–2096. [Google Scholar] [CrossRef]
  84. Hill, M.J.; Donald, G.E. Estimating Spatio-Temporal Patterns of Agricultural Productivity in Fragmented Landscapes Using Avhrr NDVI Time Series. Remote Sens. Environ. 2003, 84, 367–384. [Google Scholar] [CrossRef]
  85. de Beurs, K.; Wright, C.; Henebry, G. Dual Scale Trend Analysis for Evaluating Climatic and Anthropogenic Effects on the Vegetated Land Surface in Russia and Kazakhstan. Environ. Res. Lett. 2009, 4, 040512. [Google Scholar] [CrossRef]
  86. de Beurs, K.; Henebry, G. Spatio-Temporal Statistical Methods for Modelling Land Surface Phenology. In Phenological Research: Methods for Environmental and Climate Change Analysis; Springer: Dordrecht, The Netherlands, 2010; pp. 177–208. ISBN 9789048133352. [Google Scholar]
  87. Sen, P.K. Estimates of the Regression Coefficient Based on Kendall’s Tau. J. Am. Stat. Assoc. 1968, 63, 1379–1389. [Google Scholar] [CrossRef]
  88. Benesty, J.; Chen, J.; Huang, Y.; Cohen, I. Noise Reduction in Speech Processing; Springer: Berlin/Heidelberg, Germany, 2009; Volume 2. [Google Scholar]
  89. Haigh, J.; Conover, W.J. Practical Nonparametric Statistics. J. R. Stat. Soc. Ser. A 1981, 144, 370. [Google Scholar] [CrossRef]
  90. Yan, Z.; Wang, S.; Ma, D.; Liu, B.; Lin, H.; Li, S. Meteorological Factors Affecting Pan Evaporation in the Haihe River Basin, China. Water 2019, 11, 317. [Google Scholar] [CrossRef] [Green Version]
  91. Minaya, V.; Corzo, G.; Romero-Saltos, H.; van der Kwast, J.; Lantinga, E.; Galárraga-Sánchez, R.; Mynett, A. Altitudinal Analysis of Carbon Stocks in the Antisana Páramo, Ecuadorian Andes. J. Plant Ecol. 2016, 9, 553–563. [Google Scholar] [CrossRef] [Green Version]
  92. Ramsay, P.M.; Oxley, E.R.B. An Assessment of Aboveground Net Primary Productivity in Andean Grasslands of Central Ecuador. Source Mt. Res. Dev. 2001, 21, 161–167. [Google Scholar] [CrossRef] [Green Version]
  93. Ilbay-Yupa, M.; Lavado-Casimiro, W.; Rau, P.; Zubieta, R.; Castillón, F. Updating Regionalization of Precipitation in Ecuador. Theor. Appl. Clim. 2021, 143, 1513–1528. [Google Scholar] [CrossRef]
  94. Domínguez-Castro, F.; García-Herrera, R.; Vicente-Serrano, S.M. Wet and Dry Extremes in Quito (Ecuador) since the 17th Century. Int. J. Climatol. 2018, 38, 2006–2014. [Google Scholar] [CrossRef]
  95. Emck, P. A Climatology of South Ecuador—With Special Focus on the Major Andean Ridge as Atlantic-Pacific Climate Divide; University of Erlangen–Nuremberg: Erlangen, Germany, 2007; Volume 1. [Google Scholar]
  96. Rollenbeck, R.; Bendix, J. Rainfall Distribution in the Andes of Southern Ecuador Derived from Blending Weather Radar Data and Meteorological Field Observations. Atmos. Res. 2011, 99, 277–289. [Google Scholar] [CrossRef]
  97. Estrella, E.H.; Stoeth, A.; Krakauer, N.Y.; Devineni, N. Quantifying Vegetation Response to Environmental Changes on the Galapagos Islands, Ecuador Using the Normalized Difference Vegetation Index (NDVI). Environ. Res. Commun. 2021, 3, 065003. [Google Scholar] [CrossRef]
  98. Haverd, V.; Smith, B.; Canadell, J.G.; Cuntz, M.; Mikaloff-Fletcher, S.; Farquhar, G.; Woodgate, W.; Briggs, P.R.; Trudinger, C.M. Higher than Expected CO2 Fertilization Inferred from Leaf to Global Observations. Glob. Chang. Biol. 2020, 26, 2390–2402. [Google Scholar] [CrossRef] [Green Version]
  99. Vega-Jácome, F. Respuesta de La Vegetación a Diferentes Escalas Temporales de Sequía En Los Andes Peruanos. Serv. Nac. De Meteorol. E Hidrol. Del Perú. 2019. Available online: https://repositorio.senamhi.gob.pe/handle/20.500.12542/294 (accessed on 28 June 2022).
  100. Mazzarino, M.; Finn, J.T. An NDVI Analysis of Vegetation Trends in an Andean Watershed. Wetl. Ecol. Manag. 2016, 24, 623–640. [Google Scholar] [CrossRef]
  101. Chen, M.; Parton, W.J.; Hartman, M.D.; del Grosso, S.J.; Smith, W.K.; Knapp, A.K.; Lutz, S.; Derner, J.D.; Tucker, C.J.; Ojima, D.S.; et al. Assessing Precipitation, Evapotranspiration, and NDVI as Controls of U.S. Great Plains Plant Production. Ecosphere 2019, 10, e02889. [Google Scholar] [CrossRef] [Green Version]
  102. Álvarez-Dávila, E.A.; Cayuela, L.; González-Caro, S.; Aldana, A.M.; Stevenson, P.R.; Phillips, O.; Lvaro Cogollo, A.; Peñuela, M.C.; von Hildebrand, P.; Jimpenez, E.; et al. Forest Biomass Density across Large Climate Gradients in Northern South America Is Related to Water Availability but Not with Temperature. PLoS ONE 2017, 12, e0171072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Girardin, C.A.J.; Espejob, J.E.S.; Doughty, C.E.; Huasco, W.H.; Metcalfe, D.B.; Durand-Baca, L.; Marthews, T.R.; Aragao, L.E.O.C.; Farfán-Rios, W.; García-Cabrera, K.; et al. Productivity and Carbon Allocation in a Tropical Montane Cloud Forest in the Peruvian Andes. Plant Ecol. Divers. 2014, 7, 107–123. [Google Scholar] [CrossRef]
  104. Spracklen, D.V.; Righelato, R. Tropical Montane Forests Are a Larger than Expected Global Carbon Store. Biogeosciences 2014, 11, 2741–2754. [Google Scholar] [CrossRef] [Green Version]
  105. Laraque, A.; Ronchail, J.; Cochonneau, G.; Pombosa, R.; Guyot, J.L. Heterogeneous Distribution of Rainfall and Discharge Regimes in the Ecuadorian Amazon Basin. J. Hydrometeorol. 2007, 8, 1364–1381. [Google Scholar] [CrossRef] [Green Version]
  106. Vaca-Jiménez, S.; Gerbens-Leenes, P.W.; Nonhebel, S. Water-Electricity Nexus in Ecuador: The Dynamics of the Electricity’s Blue Water Footprint. Sci. Total Environ. 2019, 696, 133959. [Google Scholar] [CrossRef]
  107. Crespo, P.J.; Feyen, J.; Buytaert, W.; Bücker, A.; Breuer, L.; Frede, H.G.; Ramírez, M. Identifying Controls of the Rainfall–Runoff Response of Small Catchments in the Tropical Andes (Ecuador). J. Hydrol. 2011, 407, 164–174. [Google Scholar] [CrossRef]
  108. Reich, P.B.; Luo, Y.; Bradford, J.B.; Poorter, H.; Perry, C.H.; Oleksyn, J. Temperature Drives Global Patterns in Forest Biomass Distribution in Leaves, Stems, and Roots. Proc. Natl. Acad. Sci. USA 2014, 111, 13721–13726. [Google Scholar] [CrossRef]
  109. Rapp, J.M.; Silman, M.R.; Clark, J.S.; Girardin, C.A.J.; Galiano, D.; Tito, R.; Doak, D.F. Intra- and Interspecific Tree Growth across a Long Altitudinal Gradient in the Peruvian Andes. Ecology 2012, 93, 2061–2072. [Google Scholar] [CrossRef]
  110. Bendix, J. Precipitation Dynamics in Ecuador and Northern Peru during the 1991/92 El Nino: A Remote Sensing Perspective. Int. J. Remote Sens. 2000, 21, 533–548. [Google Scholar] [CrossRef]
  111. Bendix, J.; Trachte, K.; Palacios, E.; Rollenbeck, R.; Göttlicher, D.; Nauss, T.; Bendix, A. El Niño Meets La Niña-Anomalous Rainfall Patterns in the “Traditional” El Niño Region of Southern Ecuador. Erdkunde 2011, 65, 151–167. [Google Scholar] [CrossRef]
  112. Cárdenas, G. Teleconexión El Niño-Oscilación Del Sur/Oscilación Del Atlántico Norte y Su Relación Con La Precipitación En Colombia; Universidad de Bogotá Jorge Tadeo Lozano: Bogotá, Colombia, 2022. [Google Scholar]
  113. Francou, B.; Vuille, M.; Favier, V.; Cáceres, B. New Evidence for an ENSO Impact on Low-Latitude Glaciers: Antizana 15, Andes of Ecuador, 0°28′S. J. Geophys. Res. Atmos. 2004, 109, D18106. [Google Scholar] [CrossRef]
  114. Lavad-Casimiro, W.S.; Labat, D.; Ronchail, J.; Espinoza, J.C.; Guyot, J.L. Trends in Rainfall and Temperature in the Peruvian Amazon–Andes Basin over the Last 40 Years (1965–2007). Hydrol. Process. 2013, 27, 2944–2957. [Google Scholar] [CrossRef]
  115. Morán-Tejeda, E.; Bazo, J.; López-Moreno, J.I.; Aguilar, E.; Azorín-Molina, C.; Sanchez-Lorenzo, A.; Martínez, R.; Nieto, J.J.; Mejía, R.; Martín-Hernández, N.; et al. Climate Trends and Variability in Ecuador (1966–2011). Int. J. Climatol. 2016, 36, 3839–3855. [Google Scholar] [CrossRef] [Green Version]
  116. Navarro-Serrano, F.; López-Moreno, J.I.; Domínguez-Castro, F.; Alonso-González, E.; Azorin-Molina, C.; El-Kenawy, A.; Vicente-Serrano, S.M. Maximum and Minimum Air Temperature Lapse Rates in the Andean Region of Ecuador and Peru. Int. J. Climatol. 2020, 40, 6150–6168. [Google Scholar] [CrossRef]
  117. Poveda, G.; Gil, M.M.; Quiceno, N. El Ciclo Anual de La Hidrología de Colombia En Relación Con El ENSO y La NAO. Bull. Inst. Fr. d’Etudes Andin. 1998, 27, 721–731. [Google Scholar] [CrossRef]
  118. Rodríguez-Erazo, N.; Pabón-Caicedo, J.D.; Bernal-Suárez, N.R.; Martínez-Collanes, J. Cambio Climático y Su Relación Con El Uso Del Suelo En Los Andes Colombianos, 1st ed.; Instituto de Investigación de Recursos Biológicos Alexander Von Humboldt: Bogotá, Colombia, 2010; Volume 1. [Google Scholar]
  119. Sulca-Jota, J.C.; Vuille, M.; Roundy, P.; Takahashi, K.; Espinoza, J.C.; Silva Vidal, Y.; Zubieta Barragán, R. Impactos de La Concurrencia de La Oscilación Madden-Julian (MJO) y de El Niño-Oscilación Sur (ENOS) En Las Temperaturas Mínimas de Verano En Los Andes Centrales Del Perú; Instituto Geofísico del Perú: La Molina, Peru, 2018; Volume 5. [Google Scholar]
  120. Arias, P.A.; Garreaud, R.; Poveda, G.; Espinoza, J.C.; Molina-Carpio, J.; Masiokas, M.; Viale, M.; Scaff, L.; van Oevelen, P.J. Hydroclimate of the Andes Part II: Hydroclimate Variability and Sub-Continental Patterns. Front. Earth Sci. 2021, 8, 666. [Google Scholar] [CrossRef]
  121. Poveda, G.; Waylen, P.R.; Pulwarty, R.S. Annual and Inter-Annual Variability of the Present Climate in Northern South America and Southern Mesoamerica. Palaeogeogr. Palaeoclim. Palaeoecol. 2006, 234, 3–27. [Google Scholar] [CrossRef]
  122. Poveda, G.; Mesa, O.J.; Agudelo, P.A.; Álvarez, J.F.; Arias, P.A.; Moreno, H.A.; Salazar, L.F.; Toro, V.G.; Vieira, S.C. Influencia Del ENSO, Oscilación Madden-Julian, Ondas Del Este, Huracanes y Fases de La Luna En El Ciclo Diurno de La Precipitación En Los Andes Tropicales de Colombia. Meteorol. Colomb. 2002, 35, 3–12. [Google Scholar]
  123. Poveda, G.; Mesa, O.J.; Salazar, L.F.; Arias, P.A.; Moreno, H.A.; Vieira, S.C.; Agudelo, P.A.; Toro, V.G.; Alvarez, J.F. The Diurnal Cycle of Precipitation in the Tropical Andes of Colombia. Mon. Weather. Rev. 2005, 133, 228–240. [Google Scholar] [CrossRef]
  124. Vicente-Serrano, S.M.; Aguilar, E.; Martínez, R.; Martín-Hernández, N.; Azorin-Molina, C.; Sanchez-Lorenzo, A.; el Kenawy, A.; Tomás-Burguera, M.; Moran-Tejeda, E.; López-Moreno, J.I.; et al. The Complex Influence of ENSO on Droughts in Ecuador. Clim. Dyn. 2016, 48, 405–427. [Google Scholar] [CrossRef] [Green Version]
  125. Dettinger, M.D.; Battisti, D.S.; Garreaud, R.D.; McCabe, G.J.; Bitz, C.M. Interhemispheric Effects of Interannual and Decadal ENSO-Like Climate Variations on the Americas. Interhemispheric Clim. Link. 2001, 1–16. [Google Scholar] [CrossRef]
  126. Lagos, P.; Silva, Y.; Nickl, E.; Mosuera, K. El Niño Related Precipitation Variability in Perú. Adv. Geosci. 2008, 14, 231–237. [Google Scholar] [CrossRef] [Green Version]
  127. Cáceres, B.; Francou, B.; Favier, V.; Bontron, G.; Maisincho, L.; Tachker, P.; Bucher, R.; Taupin, J.D.; Delachaux, F.; Chazarin, J.P. EI Glaciar 15 Del Antisana. Diez Años de Investigaciones Glaciológicas. Mem. De La Prim. Conf. Int. De Cambio Climático: Impacto En Los Sist. De Alta Montaña 2007, 1, 63–74. [Google Scholar]
  128. Madden, R.A.; Julian, P.R. Observations of the 40-50-Day Tropical Oscillation—A Review. Mon. Weather. Rev. 1994, 122, 814–837. [Google Scholar] [CrossRef]
  129. Recalde-Coronel, G.C.; Zaitchik, B.; Pan, W.K. Madden–Julian Oscillation Influence on Sub-Seasonal Rainfall Variability on the West of South America. Clim. Dyn. 2020, 54, 2167–2185. [Google Scholar] [CrossRef]
Figure 1. Location of the Páramos in Ecuador. (a) Three panels for easy visualization of the study area. These were numbered from north to south. The locations of the hydrological stations (turquoise points) and their watersheds (orange polygons) are displayed, along with the average NDVI (Normalized Difference Vegetation Index) for the period 2001–2018. (b) Location of Ecuador in South America. (c) Location of the Tussock Páramo ecosystem in, the cuts A-A’ and B-B’, and the location of the ecuadorian hydrological basins. S1–S3 sections indicated in (a). (d) Location of the 12 high-elevation hydrological basins that have more than 50% of their surface covered by the tussock Páramo ecosystem.
Figure 1. Location of the Páramos in Ecuador. (a) Three panels for easy visualization of the study area. These were numbered from north to south. The locations of the hydrological stations (turquoise points) and their watersheds (orange polygons) are displayed, along with the average NDVI (Normalized Difference Vegetation Index) for the period 2001–2018. (b) Location of Ecuador in South America. (c) Location of the Tussock Páramo ecosystem in, the cuts A-A’ and B-B’, and the location of the ecuadorian hydrological basins. S1–S3 sections indicated in (a). (d) Location of the 12 high-elevation hydrological basins that have more than 50% of their surface covered by the tussock Páramo ecosystem.
Sustainability 15 11562 g001
Figure 2. (a) Elevation profile along the continental divide. (b) Histogram of elevations derived from 30 m DEM for both the Páramo ecosystem and the tussock Páramo. (c) Elevation profile of cross section A–A’s where the altitudinal extension of the tussock Páramo ecosystem can be observed in both the western and eastern mountain ranges. (d) Elevation profile of cross section B–B’s and altitudinal extension of the tussock Páramo ecosystem in the western and eastern mountain ranges. The location of the cross sections A–A’ and B–B’ can be seen in Figure 1. Green lines shown the upper and lower elevation limits of the paramo ecosystem.
Figure 2. (a) Elevation profile along the continental divide. (b) Histogram of elevations derived from 30 m DEM for both the Páramo ecosystem and the tussock Páramo. (c) Elevation profile of cross section A–A’s where the altitudinal extension of the tussock Páramo ecosystem can be observed in both the western and eastern mountain ranges. (d) Elevation profile of cross section B–B’s and altitudinal extension of the tussock Páramo ecosystem in the western and eastern mountain ranges. The location of the cross sections A–A’ and B–B’ can be seen in Figure 1. Green lines shown the upper and lower elevation limits of the paramo ecosystem.
Sustainability 15 11562 g002
Figure 3. Schematization of the spatial division of the study area with its corresponding percentage of the total area. The central line of the western and eastern mountain range represents their summit line. Zone 1: West of the Northern Western Range; zone 2: East of the Northern Western Range; zone 3: Northern Inter-Andean Valley; zone 4: West of the Northern Royal Range; zone 5: East of the Northern Royal Range; zone 6: West of the Southern Western Range; zone 7: East of the Southern Western Range; zone 8: Southern Inter-Andean Valley; zone 9: West of the Southern Royal Range; zone 10: East of the Southern Royal Range.
Figure 3. Schematization of the spatial division of the study area with its corresponding percentage of the total area. The central line of the western and eastern mountain range represents their summit line. Zone 1: West of the Northern Western Range; zone 2: East of the Northern Western Range; zone 3: Northern Inter-Andean Valley; zone 4: West of the Northern Royal Range; zone 5: East of the Northern Royal Range; zone 6: West of the Southern Western Range; zone 7: East of the Southern Western Range; zone 8: Southern Inter-Andean Valley; zone 9: West of the Southern Royal Range; zone 10: East of the Southern Royal Range.
Sustainability 15 11562 g003
Figure 4. Comparison of the monthly mean of the NDVI anomalies of the study area (black bars) and the 36–month moving mean of the NDVI anomalies (red line) of the study area averaged for the period 2001–2018. El Niño and La Niña events are represented by blue and fuchsia dots, respectively.
Figure 4. Comparison of the monthly mean of the NDVI anomalies of the study area (black bars) and the 36–month moving mean of the NDVI anomalies (red line) of the study area averaged for the period 2001–2018. El Niño and La Niña events are represented by blue and fuchsia dots, respectively.
Sustainability 15 11562 g004
Figure 5. Peaks of quarterly maximums of NDVI in the grasslands. For each pixel, the quarter where the maximum NDVI values usually occur has been selected. The year was divided into the following quarters: DFJ, MAM, JJA, and SON according to Ecuadorian seasonality. Black Boxes S1–S3 indicated the section number.
Figure 5. Peaks of quarterly maximums of NDVI in the grasslands. For each pixel, the quarter where the maximum NDVI values usually occur has been selected. The year was divided into the following quarters: DFJ, MAM, JJA, and SON according to Ecuadorian seasonality. Black Boxes S1–S3 indicated the section number.
Sustainability 15 11562 g005
Figure 6. Monthly discharge registered in 12 high-elevation hydrographic basins that have more than 50% of their surface covered by the Páramo ecosystem. The download was provided by the National Institute of Meteorology and Hydrology (INAMHI).
Figure 6. Monthly discharge registered in 12 high-elevation hydrographic basins that have more than 50% of their surface covered by the Páramo ecosystem. The download was provided by the National Institute of Meteorology and Hydrology (INAMHI).
Sustainability 15 11562 g006
Table 2. NDVI range values and general interpretation according to vegetation density [76].
Table 2. NDVI range values and general interpretation according to vegetation density [76].
NDVI ValueCover Type
−1.0–0.1Sterile area
0.1–0.5Vegetative cover
0.5–0.7Dense vegetation
0.7–1.0Very dense vegetation
Table 3. INAMHI surface water discharge stations at high elevations. The Páramo ecosystem is the dominant land cover type covering more than 50% of the surface within the hydrographic basins.
Table 3. INAMHI surface water discharge stations at high elevations. The Páramo ecosystem is the dominant land cover type covering more than 50% of the surface within the hydrographic basins.
CodeNameLat (°)Long (°)Elevation (m)Area (km2)% of PáramoData
H0064El Ángel en Puente Ayora0.6375−77.95182889124.7756.02003–2013
H0333San Lorenzo en San Lorenzo−1.6901−78.99782438106.8066.92000–2013
H0337Pangor Aj Salto−1.9319−79.00281480280.0550.12000–2013
H0793Cusubamba−1.0644−78.69222962181.1265.92000–2013
H1143Ambato en Mazanahuaico−1.2824−78.76363018450.6757.52005–2013
H0158Pita Aj Salto−0.5710−78.42403550127.3280.82000–2009
H0722Yanahurco Dj Valle−0.6953−78.2825360687.0698.32000–2013
H0787Alao en Hda. Alao−1.8772−78.51173200114.6072.32000–2013
H0788Puela Aj. Chambo−1.5122−78.47472475208.0352.22000–2013
H0789Guargualla Aj. Cebadas−1.8739−78.60522828189.3873.02004–2013
H0790Cebadas Aj. Guamote−1.8872−78.63842840707.3862.32000–2013
H0896Matadero en Sayausi−2.8766−79.07302602299.5182.52000–2013
Table 4. Description of the studied teleconnection indices.
Table 4. Description of the studied teleconnection indices.
NameYears/
Resolution
Definition/Website
Antarctic
Oscillation (AAO)
1979–present
Monthly
Empirical orthogonal function to the 1000-hPa mean height anomaly.
www.cpc.ncep.noaa.gov/products/precip/CWlink/daily_ao_index/aao/ (accessed on 21 February 2020)
ENSO Multivariate Index (MEI)1950–present
Monthly
Principal component of sea pressure level, zonal and meridional components of surface wind, sea surface temperature, surface air temperature, and cloud cover.
psl.noaa.gov/enso/mei/ (accessed on 21 February 2020)
Madden–Julian Oscillation (MJO)1978–present
Daily
A pair of empirical orthogonal functions of the combined fields of averaged 850-hPa zonal wind, 200-hPa zonal wind, and satellite-observed outgoing longwave radiation.
www.cpc.ncep.noaa.gov/products/precip/CWlink/daily_mjo_index/proj_norm_order.ascii (accessed on 30 January 2020)
North Atlantic Oscillation (NAO)1950–present
Monthly
Rotated principal component analysis on 500 mb height anomalies.
www.cpc.ncep.noaa.gov/products/precip/CWlink/pna/nao.shtml (accessed on 23 January 2020)
Pacific Decadal Oscillation (PDO)1854–present
Monthly
Spatial average of the monthly sea surface temperature in the Pacific Ocean north of 20° N. www.ncdc.noaa.gov/teleconnections/pdo/ (accessed on 24 January 2020)
Niño 1 + 21948–present
Monthly
Sea surface temperature in the El Niño 1 + 2 region. psl.noaa.gov/data/correlation/nina1.data (accessed on 22 January 2020)
Niño 31948–present
Monthly
Sea surface temperature in the El Niño 3 region.
psl.noaa.gov/data/correlation/nina3.data (accessed on 22 January 2020)
Niño 41948–present
Monthly
Sea surface temperature in the El Niño 4 region.
psl.noaa.gov/data/correlation/nina4.data (accessed on 22 January 2020)
Niño 3.41948–present
Monthly
Sea surface temperature in the El Niño 3.4 region.
psl.noaa.gov/data/correlation/nina34.data (accessed on 22 January 2020)
Table 5. Elevation, precipitation and land surface temperature summarized for the defined zones within the study area.
Table 5. Elevation, precipitation and land surface temperature summarized for the defined zones within the study area.
No.Study Area %ElevationPrecipitationLand Surface Temperature
MeanSDMeanSDMeanSD
122.8291.9291.9778.5107.217.62.7
28.3202.1202.1619.882.217.93.0
34.7190.9190.9911.0123.619.53.1
426.0208.9208.9946.9134.216.33.5
511.3245.3245.31285.1186.415.74.0
615.8267.9267.9436.164.018.33.5
74.4211.4211.4577.276.717.43.7
82.5230.0230.0799.5108.417.53.2
93.4232.5232.5933.8131.018.33.1
100.8131.6131.6944.6132.519.43.0
Table 6. Average of the mean, median, standard deviation, z-score, Mann–Kendall, Sen Slope and TI-NDVI in each of the areas shown in Figure 3. Trimestral Picks: March-April-May (MAM), June-July-August (JJA).
Table 6. Average of the mean, median, standard deviation, z-score, Mann–Kendall, Sen Slope and TI-NDVI in each of the areas shown in Figure 3. Trimestral Picks: March-April-May (MAM), June-July-August (JJA).
No.MeanMedianSDz-ScoreMann
Kendall
Sen SlopeTI-NDVITrimestral
Picks
10.560.570.071−0.170.120.000163715.8MAM
20.580.580.0740.010.130.000163802.1MAM
30.600.600.0710.260.090.000123926.0MAM
40.590.590.0700.220.130.000163906.7MAM
50.590.600.0780.240.150.000183915.7MAM
60.540.540.069−0.430.160.000183585.6MAM
70.550.550.055−0.370.150.000143610.7MAM
80.570.570.053−0.050.080.000093767.2JJA
90.590.590.0520.130.210.000203857.3MAM
100.590.590.0470.140.250.000253860.1MAM
Table 7. Summary of correlation between NDVI and climatic variables. Pearson’s correlation was applied at various temporal levels (monthly, bimonthly, quarterly, four-monthly, half-yearly and annually). Only are presented correlations with a significant level of 95%.
Table 7. Summary of correlation between NDVI and climatic variables. Pearson’s correlation was applied at various temporal levels (monthly, bimonthly, quarterly, four-monthly, half-yearly and annually). Only are presented correlations with a significant level of 95%.
Temporary
Scale
PrecipitationTemperaturePrec. + Temp.
Mean Corr.% DataMean Corr.% DataMean Corr.% Data
1 month0.14744−0.414500.42853
2 months0.30858−0.429470.49552
3 months0.40973−0.446400.52750
4 months0.45155−0.190130.56751
6 months0.48048−0.544360.62045
12 months0.2808−0.54090.63026
Table 8. Average of the six-monthly Pearson correlation of NDVI—Precipitation, NDVI—Temperature and NDVI—Precipitation + Temperature of the significant cells corresponding to each zone shown in Figure 3. For each zone, the percentage of significant correlations with respect to the cells within them.
Table 8. Average of the six-monthly Pearson correlation of NDVI—Precipitation, NDVI—Temperature and NDVI—Precipitation + Temperature of the significant cells corresponding to each zone shown in Figure 3. For each zone, the percentage of significant correlations with respect to the cells within them.
No.NDVI—PrecipitationNDVI—TemperatureNDVI—Prec. + Temp.
Mean Corr.% DataMean Corr.% DataMean Corr.% Data
10.5273−0.57570.6866
20.4750−0.50280.6043
30.4735−0.48240.5747
40.4025−0.51270.5735
50.4132−0.45190.4932
60.5270−0.58620.6666
70.4450−0.45150.5322
80.41340.0000.443
90.4245−0.4520.553
100.4132−0.4270.6810
Table 9. Pearson’s correlation coefficients (R2) as a result of a linear regression between mean monthly discharge flows and the total mean of NDVI. Linear regression was performed between the total mean of NDVI, the mean of the VI time series for each group.
Table 9. Pearson’s correlation coefficients (R2) as a result of a linear regression between mean monthly discharge flows and the total mean of NDVI. Linear regression was performed between the total mean of NDVI, the mean of the VI time series for each group.
StationsMean NDVI
Western RangeRoyal Range
R2R2
H00640.06-
H03330.45 *-
H03370.55 *-
H0896-0.50 *
H0793-0.43 *
H1143-0.20
p value: * < 0.05;
Table 10. Average of the spatial correlation between NDVI and the Global Teleconnection Indices (GTI). Pearson’s correlation with a significance level of 95% was applied in each cell.
Table 10. Average of the spatial correlation between NDVI and the Global Teleconnection Indices (GTI). Pearson’s correlation with a significance level of 95% was applied in each cell.
No.AAOMEIMJONAOPDO
Mean R% DataMean R% DataMean R% DataMean R% DataMean R% Data
10.2376%0.1926%0.2031%0.3291%0.2069%
20.2374%0.1920%0.2238%0.3282%0.2569%
30.2467%0.1924%0.2339%0.3086%0.2762%
40.2473%0.1921%0.2133%0.2974%0.2350%
50.2574%0.1925%0.2625%0.3172%0.2141%
60.2168%0.2667%0.1810%0.3896%0.3391%
70.2137%0.2369%0.2013%0.3587%0.2973%
80.2426%0.2767%0.1920%0.3077%0.2451%
90.2666%0.1939%0.1817%0.3187%0.2776%
100.2568%0.1933%0.142%0.3091%0.2890%
No.EL NIÑO 1 + 2EL NIÑO 3EL NIÑO 4EL NIÑO 3.4
Mean R% DataMean R% DataMean R% DataMean R% Data
10.2155%0.2054%0.2020%0.1934%
20.2049%0.1941%0.2018%0.1825%
30.2145%0.2042%0.2025%0.1930%
40.2357%0.2049%0.2018%0.1930%
50.2565%0.2251%0.2021%0.2033%
60.2576%0.2785%0.2658%0.2574%
70.2265%0.2375%0.2568%0.2370%
80.2459%0.2971%0.2868%0.2970%
90.2366%0.2158%0.2038%0.2044%
100.2484%0.2055%0.1928%0.1832%
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

Villarreal-Veloz, J.; Zapata-Ríos, X.; Uvidia-Zambrano, K.; Borja-Escobar, C. Spatio-Temporal Description of the NDVI (MODIS) of the Ecuadorian Tussock Grasses and Its Link with the Hydrometeorological Variables and Global Climatic Indices. Sustainability 2023, 15, 11562. https://doi.org/10.3390/su151511562

AMA Style

Villarreal-Veloz J, Zapata-Ríos X, Uvidia-Zambrano K, Borja-Escobar C. Spatio-Temporal Description of the NDVI (MODIS) of the Ecuadorian Tussock Grasses and Its Link with the Hydrometeorological Variables and Global Climatic Indices. Sustainability. 2023; 15(15):11562. https://doi.org/10.3390/su151511562

Chicago/Turabian Style

Villarreal-Veloz, Jhon, Xavier Zapata-Ríos, Karla Uvidia-Zambrano, and Carla Borja-Escobar. 2023. "Spatio-Temporal Description of the NDVI (MODIS) of the Ecuadorian Tussock Grasses and Its Link with the Hydrometeorological Variables and Global Climatic Indices" Sustainability 15, no. 15: 11562. https://doi.org/10.3390/su151511562

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