Next Article in Journal
Quantification of Alpine Metamorphism in the Edolo Diabase, Central Southern Alps
Next Article in Special Issue
The Evolution of the Two Largest Tropical Ice Masses since the 1980s
Previous Article in Journal
Latest Pleistocene and Holocene Floodplain Evolution in Central Europe—Insights from the Upper Unstrut Catchment (NW-Thuringia/Germany)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Topographical Impact on Snow Cover Distribution in the Trans-Himalayan Region of Ladakh, India

1
Department of Geography, South Asia Institute, Heidelberg University, 69115 Heidelberg, Germany
2
Heidelberg Center for the Environment, Heidelberg University, 69120 Heidelberg, Germany
*
Author to whom correspondence should be addressed.
Geosciences 2022, 12(8), 311; https://doi.org/10.3390/geosciences12080311
Submission received: 15 July 2022 / Revised: 17 August 2022 / Accepted: 18 August 2022 / Published: 20 August 2022
(This article belongs to the Special Issue Mountain Glaciers, Permafrost, and Snow)

Abstract

:
This article presents the distribution of seasonal snow cover in the Trans-Himalayan region of Ladakh over the observation period of 2000–2019. Seasonal snow cover area and duration have been monitored and mapped based on the MODIS Normalised Difference Snow Index (NDSI). Using different MODIS cloud removal algorithms, monthly mean cloud-covered areas have been reduced to 3%. Pixel-wise approaches using Mann–Kendall (MK) and Sen’s slope trend tests allow to assess seasonal and annual trends of snow cover days (SCD) and snow cover area (SCA) across seven delineated subregions of Ladakh. Analyses include the impact of topographical parameters (elevation, slope, aspect). Overall, the mean annual SCA amounts to 42%, varying from 15% in August to 71% in February. However, large differences of SCA have been detected between and within subregions. The trend analysis of SCA shows a non-significant, slight increase for summer as well as for the entire year and a decrease for spring and winter seasons. The SCD trend analysis indicates more pixels with a significant increase than a decrease. In total, 12% of all pixels show an increasing trend in summer, 6% over the entire year, 3% in autumn, and 2% in spring and winter, whereas less than 2% of all pixels show a decreasing trend in all seasons. The results are important for regional irrigated agricultural production and freshwater supply in the context of climate change.

1. Introduction

Seasonal snow cover and glaciers are key components of the Himalayan hydrological cycle, as more than 60% of the annual discharge is supplied by meltwater, with considerable variation across the region [1,2,3,4,5]. Especially in the cold-arid Trans-Himalayan region of Ladakh, meltwater supply from the cryosphere is of utmost importance for irrigated agriculture in spring and summer, when water demand is highest [6,7]. Thus, detailed information on snow-covered areas is essential for integrated water resource management and for assessments of regional climate change and associated impacts on meltwater runoff [8,9]. To address these crucial aspects, mapping and monitoring of spatial and temporal snow cover dynamics are urgently needed to support cryosphere-dependent livelihoods [10]. Despite the importance of seasonal snow cover for irrigated land use, the amount and timing of high-altitude snowfall is largely unknown [11]. Only few studies have analysed the distribution, duration, and trends of seasonal snow cover on the Himalayan scale [12,13,14,15]. To quantify snow-covered area (SCA) and snow cover days (SCD) at the regional scale across High-Mountain Asia [2,3,16,17,18,19,20], different snow products of Moderate-Resolution Imaging Spectroradiometer (MODIS) have been used, including the 8-day global-gridded product with a spatial resolution of 0.05° [13] or daily snow products with a spatial resolution of 500 m [12,14,20,21]. However, enormous interregional differences and large variabilities make it difficult to assess snow cover distribution and trends for the entire Himalayan region [12,13]. Therefore, detailed studies are required for an improved understanding of snow cover distribution and related changes. On the regional scale, the snow cover pattern is influenced by elevation and other topographical parameters such as aspect and slope [22,23]. Only few studies consider these topographic controls on snow cover distribution in the Himalayas [3,21,24]. For the Trans-Himalayan region of Ladakh, seasonal snow cover dynamics and their changes are largely unknown as long-term snow observations are only available for the meteorological station in Leh (Figure 1). Using climatological reanalysis data, the spatial distribution of solid and liquid precipitation can be assessed without considering the impact of topography [25].
Against the background of these knowledge gaps, this article provides an analysis of snow cover distribution and duration in Ladakh using daily MODIS snow product data for the period of 2000–2019. The spatio-temporal patterns and trends of snow-covered area (SCA) and snow cover days (SCD) have been investigated in seven delineated subregions, based on watersheds and mountain ranges. Furthermore, the impact of topography (elevation, slope, and aspect) on the SCA has been analysed to detect and quantify trends in individual subregions.
Ladakh is located in the rain shadow of the Karakoram to the north and the Greater Himalaya to the south. Due to the cold-arid conditions along the western edge of the Tibetan plateau, only small high-altitude glaciers exist along the Central Ladakh Range and on the Changthang Plateau [27,28,29,30], whereas large valley glaciers exist in the Karakoram [31,32] and along the Greater Himalayan Range [33]. In the winters, large aufeis fields occur along the tributaries of the Upper Indus Basin (UIB) and in the endorheic basins of Tshomoriri and Pangong as a temporary cryosphere component [34].
The elevation ranges from 2540 m a.s.l. in the lower Indus Valley near Kargil in western Ladakh to 7742 m a.s.l. on the Saltoro Ridge in the north (Figure 1). The mean elevation of Ladakh amounts to 4875 m a.s.l., increasing from the west to the east (Figure 2, Table 1); thus, the subregion Suru has the lowest and Pangong the highest mean elevation. Temperatures vary between more than 30 °C in the lower Indus valley during summer to less than −40 °C during winter in Drass of the Suru region. The region receives low precipitation, which decreases from the west to the east from about 600 mm to less than 100 mm in the Indus valley and in the Changthang region [25,35].

2. Materials and Methods

2.1. Materials

Daily snow cover maps from MOD10A1 (MODIS Terra) and MYD10A1 (MODIS Aqua) with a resolution of 500 m were used to investigate the temporal and spatial distribution of seasonal SCA and SCD between October 2000 and September 2019. For the period from October 2000 to December 2016, binary snow cover data of MODIS Collection 5 were used, which are based on a global Normalised Difference Snow Index (NDSI) threshold of 0.4 [36]. For the period from January 2017 to September 2019, NDSI data from MODIS Collection 6 were used, as binary snow cover data are no longer available. These data were downloaded from the NASA Distributed Active Archive Center (DAAC), hosted by the National Snow and Ice Data Center (NSIDC) (https://nsidc.org; accessed on 22 November 2019). To validate the MODIS NDSI threshold, Landsat-8 OLI data (downloaded from https://earthexplorer.usgs.gov; accessed on 10 January 2020) and oblique terrestrial photographs from two permanently installed cameras mounted in the Leh valley between 3660 and 4900 m a.s.l. were used (Figure 3).
To analyse the impact of topographical parameters (elevation, slope, and aspect) on snow cover distribution, a void-filled Shuttle Radar Topography Mission Digital Elevation Model (SRTM-DEM) (https://earthexplorer.usgs.gov; accessed on 1 July 2019) was resampled to 500 m. To consider the spatial differentiation of climatic and topographical conditions, the study area was separated into seven subregions based on watersheds and mountain ranges: Siachan (Siachen), Shayok (Shyok), Pangong, Indus, Suru, Zangskar (Zanskar), and Tshomoriri (Tso Moriri) (Figure 2). DEMs with high spatial resolution (High-Mountain Asia (HMA) DEM [37] (retrieved from https://nsidc.org; accessed on 15 August 2020) and TanDEM-X 12 m data (retrieved on special request by German Aerospace Center (DLR)) were used to geo-reference terrestrial photographs, calculate viewsheds, and to delineate snow lines at the local scale using the Monoplotting Tool (MPT 2.0, Monoplotting Tool—WSL) as an additional ground truthing method (Figure 3).

2.2. Methods

To map and analyse SCD and SCA, daily cloud-free MODIS snow maps have been generated. Based on these data, monthly, seasonal, and annual composites have been calculated for the analysis of spatial and temporal distribution of SCA and the impact of topography on it. In a second step, trend analyses have been carried out for the SCA and SCD.

2.2.1. NDSI Threshold Value Calibration

The binary data of global MODIS snow cover products were calculated based on the NDSI (Equation (1)) with a threshold of 0.4, for pixels with the visible reflectance (near infrared (NIR): Band 2) greater than 0.11 and Band 4 (green) reflectance greater than 0.10 [36]. In collection 5, the NDSI is calculated for MODIS Terra using green (Band 4) and shortwave infrared 1 (SWIR 1: Band 6), and for MODIS Aqua using green and SWIR 2 (Band 7) due to non-functional detectors. In collection 6, the quantitative image restoration (QIR) algorithm enables the usages of SWIR 1 for both Terra and Aqua with a higher accuracy compared to the usage of SWIR 2 [38]. Thus, for both MODIS Terra and Aqua, the NDSI can be calculated with the same band combinations (Equation (1)):
NDSI = (Green – SWIR 1)/(Green + SWIR 1)
The comparison of MODIS snow products collection 5 and 6 using different NDSI threshold schemes showed no significant differences and had identical accuracies [39]. Several studies, e.g., [12,40,41], applied the global reference NDSI threshold of 0.4 for the Himalaya and the Zangskar region [42], whereas in other studies a much lower threshold of 0.01 or 0.2 was suggested [39,43]. To enable statistical trend analyses, the same global reference NDSI threshold of 0.4, applied in the binary data of collection 5, was also used for collection 6.
To validate the NDSI threshold, two cloud-free Landsat-8 OLI tiles (path 147/row 36–37) from four seasons were used for ground truthing [44,45,46,47]. Pixels were classified as snow-covered if the NDSI was greater than or equal to 0.40 and the NIR was greater than 0.11 [44,48,49,50]. The snow cover data were resampled to 500 m to match the spatial resolution of MODIS.
An accuracy assessment was carried out using Cohen’s kappa coefficient ( κ ) method on a range of different MODIS NDSI threshold value images. The highest κ for each MODIS threshold value image was regarded to feature the appropriate NDSI threshold value, since κ represents the level of agreement between two datasets, considering the possibility of the agreement occurring by chance [51]. The formula is presented in Equation (2):
κ = p o p e 1 p e
where p o is the relative observed agreement among rates and p e is the hypothetical probability of chance agreement.
The validation shows that the NDSI threshold varies between the seasons. In winter and spring, the NDSI predominantly exceeds the reference threshold value of 0.4, whereas in summer and autumn, the threshold is lower. However, the used NDSI threshold of 0.4, which also represents the global NDSI threshold, shows an overall accuracy of 87.20%.
In a second approach, Landsat-8 OLI and MODIS snow cover maps were compared to the monoplotted terrestrial images taken on the same date.
The results fit well with the terrestrial photos. Comparison amongst the coarse MODIS, high Landsat, and very high terrestrial photos shows that the resolution of pixels plays a huge role in detecting shallow and fragmented snow trace in the transition zone. MODIS underestimates the snow-covered area where a pixel is less than 50% snow-covered, and overestimates where a pixel is more than 50% snow-covered [52]. For example, ephemeral shallow and fragment snow pixels in the near range of the Lungmar camera are not registered by MODIS (Figure 4i), but are detected by Landsat (Figure 4h), and the permanent SCA is overestimated by MODIS where a pixel is more than 50% snow-covered (Figure 4l), compared to Landsat (Figure 4k). A slight underestimation of the snow cover can be assumed due to the shallow occurrence and patchy distribution of snow in the study area.

2.2.2. Methods Applied for MODIS Cloud Removal

Trend analyses of seasonal snow cover require high temporal resolution data. As in optical satellite data, clouds obscure the view on the ground, several approaches have been developed to reduce the degree of cloud-covered pixels in daily MODIS snow products [14,46,53,54,55]. Despite the cold-arid conditions of Ladakh, the mean cloud coverage amounts to 49.54% (Terra) and 55.36% (Aqua), varying from 11.80% (Terra) and 17.19% (Aqua) in October 2017 to 87.86% (Terra) and 90.29% (Aqua) in January 2008. Therefore, all daily MODIS snow products were processed to reduce the cloud pixels (Figure 5).
In the first step, pixel values of Terra and Aqua snow products in collection 5 were reclassified into three categories as snow (original snow and lake ice), land (no snow and lake classes), and clouds (all other missing/unusable classes). For collection 6, the NDSI was used to delineate between snow and land pixels. The ratios of land, snow, and cloud cover indicate the proportions of land, snow, and cloud-covered areas, respectively (Equation (3)):
R C = A C A × 100 % = P X P × 100 %
where R C is the ratio of land, snow, or cloud cover, A C is the area of land, snow, or cloud cover, A is the total area of the study area, P X is the number of land, snow, or cloud pixels, and P is the total number of pixels of the entire study area.
The combination of Terra and Aqua data takes advantage of the three-hour interval time between two observations. The cloud movement enables to reduce cloud-covered pixels under the assumption that snow conditions remain essentially constant within three hours. If a pixel is classified as cloud on one image but is reliably classified as land or snow cover on the other, then the cloud-covered pixel is reclassified appropriately [53,55] (Equation (4)):
P ( x , y , t ) = m a x ( P ( x , y , t ) T , P ( x , y , t ) A )
where x is the latitude, y is the longitude, and t is a temporal (day) pixel, P . P T and P A represent pixels from Terra and Aqua, respectively.
In the second step, a temporal combination was conducted which combines the preceding and succeeding day with the actual day of observation. Cloud-covered pixels are replaced either by snow or land pixels under the assumption of unchanged surface conditions within three days [46,55,56,57]. The formula is presented in Equation (5):
P ( x , y , t ) = 1 ,   if   ( P ( x , y , t 1 ) = 1   and   P ( x , y , t + 1 ) = 1 )
where x is the latitude, y is the longitude, and t is a temporal (day) pixel, P . t 1 and t + 1 represent the preceding and succeeding days, respectively, where 1 corresponds to snow cover and 0 for land cover.
The criterion of discarding images with more than 70% cloud pixels was applied before the third step to improve the results [58]. In total, 183 images were discarded over the entire observation period from October 2000 to September 2019. This approach contained an intermediate spatial step delineating the Regional Snow Line Elevation (RSLE) and the Regional Land Line Elevation (RLLE) by using the average (arithmetic mean) elevation derived from the SRTM DEM of all snow or land pixels of a single dataset [46,54,56,58,59,60] (Equations (6) and (7)):
R S L E = 1 n i = 1 n ( P ( x , y , z ) S ) i  
R L L E = 1 n i = 1 n ( P ( x , y , z ) L ) i
where x is the latitude, y is the longitude, and z is the elevation of each snow pixel, P S , or land pixel, P L , in the list of elevations being averaged, and n is the total number of snow or land pixels.
Finally, cloud pixels were reclassified to land if their elevation value was below the mean RLLE, and to snow if their elevation value was above the mean RSLE (Figure 6). The remaining cloud pixels located in the transition zone between the RSLE and RLLE were assigned as mixed pixels (patches of snow and land) and not included in the analysis. The result of this approach is a completely cloud-free spatio-temporal combination of MODIS data. The formula is presented in Equation (8):
P ( x , y , z ) C = { P ( x , y , z ) S     i f   ( P ( x , y , z ) C > R S L E ) P ( x , y , z )   L i f   ( P ( x , y , z ) C > R L L E ) P ( x , y , z ) M   i f   ( P ( x , y , z ) C   R S L E   a n d   P ( x , y , z ) C R L L E )
where x is the latitude, y is the longitude, and z is the elevation value of a cloud ( P C ), snow ( P S ), land ( P L ), or mixed pixel ( P L ).
The first cloud reduction step (Figure 7), based on the combination of Terra and Aqua MODIS products, reduced the monthly mean cloud cover to 41%. The temporal combination approach decreased the cloud cover to 17%. With the final spatio-temporal combination, the remaining cloud pixels located between the RSLE and RLLE amounted to 3% on average (maximum 18%) and were classified as mixed pixels.

2.2.3. Snow Trend and Correlation Analysis

The monthly, seasonal, and annual mean MODIS SCA and SCD were calculated for the entire region of Ladakh and the seven subregions. To identify trends in time series, the Mann–Kendall (MK) [61,62] and Sen’s slope trend tests were applied [63]. The MK test is applied to identify whether there is a trend in the time series with statical significance at the confidence interval of 90% (P < 0.1), 95 % (p < 0.05) and 99.9% (p < 0.01). If a trend exists, the magnitude can be determined by the non-parametric Sen’s slope estimator, which is a robust method against outliers to estimate the slope of the trend. The trends were calculated for months, seasons, and hydrological years from 2000 to 2019. The values of the trend test and magnitude of trends are shown by Kendall’s tau (τ) coefficient and Sen’s slope (S). The seasons were categorized as autumn (September–November), winter (December–February), spring (March–May), and summer (June–August), and as the entire hydrological year (October–September).
A detailed study was conducted to analyse the spatio-temporal pattern and the impact of topographical parameters (elevation, aspect, and slope) on the SCA and its trends using the modified MK test with a two-variance correction approach [64,65]. A Sen’s slope trend test and a linear regression model were calculated to detect and quantify the magnitude of trends across all subregions over 19 hydrological years (significance levels: p-values < 0.1, <0.05, and <0.01).

3. Results

Snow accumulation starts in September and lasts until February, when the maximum SCA of the entire area of Ladakh reaches 71%, with the exceptions of subregions Zangskar and Suru, where the maximum SCA is delayed until March (Figure 8, Table A2). An average of 61% of the study area is snow-covered during winter, varying from 49% in the snow-scarce winter of 2002/03 to 73% in the snowy winter of 2008/09. Three more snow-scarce winters occurred in 2000/01, 2011/12, and 2017/18, with a SCA of about 50%. In the year 2017/18, the SCA appeared to be significantly below average until summer. While the snow-scarce winter of 2002/03 was detectable in all subregions, the reduced snow cover during the winter of 2011/12 could only be identified in the two northern subregions of Siachan and Shayok, as well as in the south-eastern subregion of Zangskar. In winter of 2017/18, the reduction in snow cover could be detected in all subregions except Suru and Zangskar.
In general, the winter SCA is generally characterized by high interannual variability, ranging from 28% to 72% in December and from 39% to 80% in January, while the SCA seems to be more stable in February, varying between 52% and 83%. About 10% of the entire study area is almost completely snow-free during winter (SCD less than 10 days in winter), and these areas are located along the large valley bottoms of the Indus, Nubra, and Shayok rivers, as well as in the Changthang area of eastern Ladakh. Only minor differences can be detected between the SCA in winter and spring, when 46% of the region that is snow-covered in February is still snow-covered in March (Figure 9, top). Snow melt generally starts in April (Figure 8) and continues until August, when only 13% of the region is snow- or ice-covered. These permanently snow-covered areas are mainly located in the subregion of Siachan and along the Greater Himalayan range, the southern boundary of the study area (Figure 9, top). Depending on the SCA, the mean RSLE ascends from 4813 m a.s.l. in February to 5097 m a.s.l. in April, and to a maximum of 5560 m a.s.l. in August (Table A1).
On the regional scale, the maximum winter SCAs with 79% and 75% can be detected in the subregions of Siachan and Suru, respectively, and the minimum in the Indus subregion, amounting to 41%. While in the Siachan subregion only minor interannual variations of the winter SCA can be identified, ranging between 70% and 85%, the largest differences can be identified for the Tshomoriri subregion, ranging between 34% and 78%. In most subregions, the snow melt starts in April, while in the two northern subregions, a delay of one month can be observed. The minimum SCA in summer amounts to less than 15% in all subregions, with the exception of Siachan, where the SCA amounts to 42% and is larger than the mean winter SCA in the subregions of Indus and Pangong.
The SCA increases with elevation, where areas above 6500 m a.s.l. remain almost completely snow-covered throughout the year (Figure 10). In the subregions Siachan and Suru, more than 80% of the elevation zone of 5500–6500 m a.s.l. is snow-covered, with only slight seasonal variations, whereas in all other subregions, a characteristic seasonality of SCA can be observed ranging from 29% in summer to 70% in winter for the example of the Indus subregion. For the elevation zones of 4500–5500 m a.s.l. and below 4500 m a.s.l., this seasonality can be detected in all subregions where the SCA varies between 15% in summer and 64% in winter. Only in the Pangong subregion is this seasonality alleviated as the SCA in winter is less than 40%. Below the elevation zone of 3500–4500 m a.s.l., the winter snow covers less than 20% in the four northern subregions, while it reaches up to 60% in Suru and 40% in Zangskar.
Aspect is another striking feature of snow cover distribution (Figure 11). South- and southwest-facing slopes are less snow-covered than all other slopes, except in the subregion of Siachan, where a shift to southwest- and west-facing slopes can be observed. In the subregions of Suru, Siachan, and Zangskar, the negative deviation is most prominent in autumn. In contrast, north- and northwest-facing slopes show the highest positive deviation in almost all subregions, except for Shayok, where west- and northwest-facing slopes, and for Tshomoriri, where north- and northeast-facing slopes have the largest SCA. In contrast to elevation and aspect, SCA does not show large differences between individual slope classes, except for the Shayok region, where a decrease of SCA with the increasing slope angle can be detected (Figure 12).
The trend analysis of the SCA for the entire study area shows no significant trends. However, a slight increase of the annual and summer SCA (0.08 and 0.2% year−1) can be detected, while in winter and spring, a slight decrease can be identified (Table 2). The separated trend analysis of SCA on the regional scale shows slight, significant, increasing trends in the subregions of Suru (0.52% year−1) and Zangskar (0.36% year−1), while in the subregions of Siachan, Shayok, and Indus, a slight (non-significant) negative trend of SCA can be observed in spring. In general, the SCA trends show a non-significant heterogeneous pattern (Table A2).
In accordance with the SCA trend analysis, the pixel-wise analysis of SCD reveals a longer duration of snow coverage. However, only 6% of all pixels show a significant increasing trend and 2% of all pixels show a significant decreasing trend for the entire study area of Ladakh. Most pixels with a significant increasing trend in summer are located in the subregions of Suru and Zangskar, with 36% and 18% of all pixels, respectively. In general, this positive trend is also reflected by the trend analysis of the RSLE showing a downward shift for the entire region, which is significant for the months of June and July (−5.2 and −5.6 m year−1, respectively).
Most pixels with a decreasing trend are located in the eastern parts of Ladakh and more often in winter than in summer; for example, in the Pangong subregion, 5% show a decreasing trend in winter and only 1% in summer. In contrast, in the subregion of Siachan, most pixels with a negative trend are from summer (4%) (Figure 9, Table 3). This general negative trend of snow coverage in winter is also reflected in the RSLE, which shows an upward shift for the months of December and January.
The trend analysis of the SCA in relation to elevation shows a heterogeneous pattern. Slight decreasing trends can be detected above 5500 m a.s.l. in almost all subregions and are mostly significant above 6500 m a.s.l. In the elevation zone of 4500–5500 m a.s.l., an increasing trend of SCA can be observed in almost all subregions, which is highest in the southern subregions of Suru and Zangskar. In the north-eastern subregion of Pangong, a significant increasing trend of SCA can be identified for elevations below 4500 m a.s.l. in autumn and summer (Figure 11). In all subregions except Zangskar and Tshomoriri, a significant increasing trend can be detected for the elevation zone of 3500–4500 m a.s.l. (Table A3) for the months of June and July.

4. Discussion

The present study is based on daily MODIS snow cover data. Using a standardized global-NDSI threshold of 0.4 for MODIS snow products has shown an overall accuracy rate of about 87% compared to Landsat data, which is in line with other studies on snow cover mapping [20,46,55,56,66,67,68]. However, the dominant occurrence of snow cover in the Trans-Himalayan region of Ladakh consists of often shallow and fragmented snow traces and patches. This characteristic type of snow distribution creates difficulties in monitoring and change detection [66,69]. As small snow patches cannot be mapped with MODIS data, some studies have used Landsat data with a higher spatial but lower temporal resolution [70] or terrestrial cameras for continuous monitoring [22,71,72].
Comparing the SCA in Ladakh with adjoining parts of the South Asian Mountain Belt confirms the transitional position of the Trans-Himalayan region between the Karakoram and the Western Himalayas (Table 4). The Chandra basin in Himachal Pradesh (minimum elevation of 2800 m a.s.l.) is characterized by a snow cover of about 60% in August and of up to 99% in February [24]. In the Astore basin located to the east of the Nanga Parbat massif (minimum elevation of 1210 m a.s.l.), the SCA varies between 7% and up to 95% [73]. The study area of Ladakh is characterized by smaller differences in SCA, amounting to 15% in August and up to 71% in February, and the mean annual SCA of 42% is almost identical to the situation in the Chenab basin [2]. The average annual SCA value of 42% lies between average SCA values for the Karakoram range with 52% and the Western Himalayan range with 30% [12]. In another study [74], the SCA is estimated for the entire Upper Indus Basin (UIB) with a value of 21%, varying from 14% in the Shayok basin (including the upper watershed on the Tibetan Plateau) to 58% in the Shigar basin. On the regional scale, a decreasing SCA from west to east and from south to north is observed, which reflects the general precipitation gradient [11,75] and the influence of westerlies on solid precipitation in winter [76]. While the maximum SCA in the Astore basin is reached in January [73], in most subregions of Ladakh it occurs in February, except of Suru and Zangskar, where the maximum is not reached until March, which is similar to most parts of the neighbouring Gilgit Baltistan (March) [18] and Jhelum (February) [77]. Snow accumulation in Ladakh starts in September, which is in accordance with observations from Gilgit Baltistan [18] and almost one month earlier than in the Karakoram, observed by Dharpure et al. [12].
The significant impact of aspect on SCA can be identified by the large differences between north-facing and south-facing slopes. As north-facing slopes receive less solar radiation, especially during autumn and winter, snow accumulation starts earlier, and snow melt commences later in comparison to south-facing slopes. These characteristic differences are most prominent in subtropical mountains [78].
A comparison of decadal snow cover trend analyses with studies from adjoining Himalayan regions is difficult, mainly due to differences in the length of the observation periods. Dharpure et al. have detected a non-significant, slightly increasing trend of SCA (0.21% year−1) along the entire Himalayan range between 2000 and 2008, which reversed to a significant negative trend since 2008 (−0.42% year−1). However, almost stable conditions have been identified over the entire observation period from 2000 to 2019 (0.00% year−1) [12]. In the same study, non-significant positive trends of SCA have been observed for the Karakoram and Western Himalayas for 2000–2008, and a significant negative trend since 2008 [12]. In contrast, other studies have reported non-significant decreasing trends for the Upper Indus Basin over the periods 2001–2012 [74] and 2000–2008, respectively, which is significant in winter [13]. No clear trend of SCA can be identified for Ladakh, however the detected slight but non-significant increase of the annual and summer SCA can possibly be related to an increase in cloudburst events, as witnessed during several field visits.
Overall, a comparison of these studies underlines the relevance of regional analyses to capture spatio-temporal dynamics of seasonal snow cover. This is particularly important for cold-arid areas, where irrigation of agricultural fields solely depends on early melt water supply in spring [7,79]. In this context, local and subregional studies of snow cover distribution and duration are of utmost importance for livelihoods and development in Ladakh and neighbouring regions.

5. Conclusions

Similar to the adjoining mountain regions, components of the cryosphere are of vital importance for downstream settlements and irrigated agriculture in Trans-Himalayan regions. This study showed the potential of combining terrestrial photography with satellite imagery to obtain more accurate and detailed information on seasonal snow cover distribution and duration across various scales. The results revealed a high variability in seasonal snow cover across the seven different subregions and over the observation period of almost twenty years. For an improved understanding of spatial and temporal dynamics of regional snow cover distribution and duration, terrestrial data and satellite imagery with high spatial and high temporal resolution should be combined for any hydrological and development studies in the context of climate change.

Author Contributions

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

Funding

This research received no external funding. Heidelberg University helped to bridge critical periods.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All relevant data will become publicly available with the publication of the PhD thesis of the first author at Heidelberg University.

Acknowledgments

The authors would like to express their gratitude to Henk Thoma (Leh, Ladakh, India) for the continuous and generous support over many years. The German Aerospace Center (DLR) kindly provided TanDEM-X data and the Swiss Organisation WSL kindly provided the Monoplotting Tool. Constructive comments by two anonymous reviewers helped to improve the article.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Monthly RSLE, Sen’s slope, and MK tests for the entire Ladakh.
Table A1. Monthly RSLE, Sen’s slope, and MK tests for the entire Ladakh.
JanFebMarAprMayJunJulAugSepOctNovDecAut.Win.Spr.Sum.Ann.
Z-Value0.28−0.70−0.91−0.56−1.26−2.10−2.35−1.33−0.21−0.77−0.280.28−0.560.42−1.19−2.10−1.12
p-Value0.780.480.360.580.210.040.020.180.830.440.780.780.580.670.230.040.26
Sen’s S.1.56−2.59−3.35−1.10−2.16−5.20−5.56−1.29−0.73−2.67−2.392.08−2.080.56−2.97−3.53−1.75
Lin. S.−0.85−1.42−2.38−1.24−2.55−5.13−5.29−1.40−1.32−2.8−2.322.95−1.920.23−2.06−3.94−1.98
Mean48324813495850975247539055265560547853785212501153534885510054925208
SD84.775.363.753.154.159.156.330.358.377.4122.294.462.350.449.343.932.5
Trend
Numbers in bold and blue: significance level, p-value < 0.05.
Table A2. A: Mean SCA (%) of all subregions and the magnitude of trends.
Table A2. A: Mean SCA (%) of all subregions and the magnitude of trends.
JanFebMarAprMayJunJulAugSepOctNovDecAut.Win.Spr.Sum.Ann.
Indus
Z-Value−1.051.190.21−0.830.001.861.610.630.00−0.07−0.42−0.21−0.14−0.28−1.070.63−0.07
p-Value0.290.230.830.411.000.060.110.531.000.940.670.830.890.780.280.530.94
Sen’s S.−0.530.610.06−0.160.040.100.70.05−0.04−0.07−0.14−0.15−0.07−0.12−0.070.08−0.02
Lin. S.−0.190.310.10−0.19−0.140.080.230.06−0.030.08−0.04−0.28−0.01−0.05−0.070.120.00
Mean44.7851.8346.4235.0323.0813.615.063.767.7912.6217.9026.7112.8641.1134.847.4724.05
SD12.5111.8710.727.665.494.623.261.435.387.738.8310.836.167.517.042.614.31
Trend
Pangong
Z-Value−0.350.700.28−0.84−0.28−0.501.820.530.53−0.21−0.42−1.05−0.42−0.91−0.260.280.00
p-Value0.730.480.780.400.780.620.070.600.600.830.670.290.670.360.800.781.00
Sen’s S.−0.160.380.19−0.26−0.12−0.040.210.040.16−0.07−0.26−0.63−0.18−0.23−0.010.040.02
Lin. S.−0.14−0.050.09−0.21−0.200.080.280.100.070.05−0.08−0.71−0.03−0.30−0.110.15−0.06
Mean49.2157.7153.6242.6430.0318.738.076.8810.9116.4621.8932.0116.5146.3142.1011.2329.01
SD14.8111.9712.628.036.726.443.902.226.0210.4410.6512.927.248.948.153.444.82
Trend
Shayok
Z-Value−0.910.56−0.63−0.56−0.560.901.400.210.490.700.14−0.420.07−0.35−1.050.350.14
p-Value0.360.580.530.580.580.370.160.830.620.480.890.670.940.730.290.730.89
Sen’s S.−0.330.17−0.21−0.25−0.190.070.270.010.110.450.18−0.320.06−0.10−0.160.090.04
Lin. S.0.030.04−0.20−0.24−0.220.040.420.090.020.340.08−0.290.13−0.08−0.220.180.01
Mean65.3267.7064.6358.5348.3834.5618.1613.4622.1032.4241.4252.4432.1261.8257.1822.0643.26
SD8.397.045.855.675.887.486.342.226.369.3910.8912.596.545.944.984.824.07
Trend
Siachan
Z-Value0.07−0.63−0.91−0.56−0.14−0.071.260.490.210.140.28−0.070.350.07−1.050.63−0.21
p-Value0.940.530.360.580.890.940.210.620.830.890.780.940.730.940.290.530.83
Sen’s S.0.05−0.10−0.26−0.37−0.10−0.030.310.070.050.090.08−0.040.040.01−0.240.13−0.04
Lin. S.0.10−0.09−0.24−0.38−0.29−0.190.270.050.000.100.10−0.050.07−0.01−0.310.04−0.05
Mean81.5581.7678.6374.6267.4857.4546.3541.8551.5358.6364.3473.5758.2578.9673.5848.5564.81
SD4.914.435.237.106.895.805.172.334.917.268.008.294.593.985.533.813.20
Trend
Suru
Z-Value0.42−0.14−0.490.770.491.612.242.030.560.280.840.280.490.350.491.961.12
p-Value0.670.890.620.440.620.110.030.040.580.780.400.780.620.730.620.050.26
Sen’s S.0.16−0.03−0.140.320.150.830.590.210.190.130.780.180.300.210.150.520.24
Lin. S.0.29−0.06−0.210.350.370.840.570.230.210.110.340.220.200.150.170.550.27
Mean79.4479.4281.7972.7161.4938.2217.7411.2321.2935.1251.2266.2936.0975.0572.0022.4051.33
SD11.487.744.958.248.0310.076.902.915.8211.1017.2010.409.146.594.726.374.11
Trend
Tshomoriri
Z-Value−0.350.141.050.000.350.841.890.560.63−0.42−0.49−0.63−0.14−0.070.561.190.35
p-Value0.730.890.291.000.730.400.060.580.530.670.620.530.890.940.580.230.73
Sen’s S.−0.350.260.68−0.070.210.310.250.070.14−0.27−0.23−0.41−0.03−0.090.320.180.13
Lin. S.−0.200.370.590.190.280.370.280.06−0.090.220.17−0.230.12−0.020.350.240.17
Mean61.2775.1373.4561.1334.0514.706.346.2112.0217.6721.6033.7217.1856.7156.219.0834.77
SD21.2117.3418.2716.1113.457.734.042.5111.3918.0216.7719.1514.2513.0914.553.928.42
Trend
Zangskar
Z-Value−0.350.700.630.490.421.542.031.330.630.000.21−0.070.000.210.421.680.77
p-Value0.730.480.530.620.670.120.040.180.531.000.830.941.000.830.670.090.44
Sen’s S.−0.310.200.170.210.220.630.370.090.120.120.10−0.030.020.150.130.280.26
Lin. S.0.160.380.240.190.190.560.350.180.110.590.240.010.280.180.210.360.27
Mean72.9380.8982.4472.0549.8926.3711.618.3415.6524.6237.4652.5026.1068.7768.1315.4444.56
SD12.258.878.6110.578.658.814.582.587.2014.7714.9718.669.728.538.534.935.90
Trend
Numbers in bold and green: significance level, p-value < 0.1. Numbers in bold and blue: significance level, p-value < 0.05.
Table A3. Trend analysis of SCA in relation to the elevation zone (%) of all subregions.
Table A3. Trend analysis of SCA in relation to the elevation zone (%) of all subregions.
Elevation ZoneRegionJanFebMarAprMayJunJulAugSepOctNovDec
<3500 mSiachan
Shayok*
Indus
Suru
Zangskar*
3500–4500 mSiachan
Shayok
Pangong*****
Indus
Suru
Zangskar
Tshomoriri
4500–5500 mSiachan
Shayok**
Pangong*
Indus
Suru
Zangskar
Tshomoriri*
5500–6500 mSiachan**
Shayok**
Pangong**
Indus*
Suru***
Zangskar
Tshomoriri
>6500 mSiachan*****
Shayok**
Indus
Suru****
Zangskar
Tshomoriri*
△ Upward Trend, ▽ Downward Trend, − No Trend; Trend Significance: None Low (p < 0.1) Moderate (p < 0.05) High (p < 0.01), * Modified Mann-Kendall Test.

References

  1. Armstrong, R.L.; Rittger, K.; Brodzik, M.J.; Racoviteanu, A.; Barrett, A.P.; Khalsa, S.-J.S.; Raup, B.; Hill, A.F.; Khan, A.L.; Wilson, A.M.; et al. Runoff from Glacier Ice and Seasonal Snow in High Asia: Separating Melt Water Sources in River Flow. Reg. Environ. Chang. 2019, 19, 1249–1261. [Google Scholar] [CrossRef] [Green Version]
  2. Jain, S.K.; Goswami, A.; Saraf, A. Role of Elevation and Aspect in Snow Distribution in Western Himalaya. Water Resour. Manag. 2009, 23, 71–83. [Google Scholar] [CrossRef]
  3. Misra, A.; Kumar, A.; Bhambri, R.; Haritashya, U.K.; Verma, A.; Dobhal, D.P.; Gupta, A.K.; Gupta, G.; Upadhyay, R. Topographic and Climatic Influence on Seasonal Snow Cover: Implications for the Hydrology of Ungauged Himalayan Basins, India. J. Hydrol. 2020, 585, 124716. [Google Scholar] [CrossRef]
  4. Mukherji, A.; Molden, D.; Nepal, S.; Rasul, G.; Wagnon, P. Himalayan Waters at the Crossroads: Issues and Challenges. Int. J. Water Resour. Dev. 2015, 31, 151–160. [Google Scholar] [CrossRef] [Green Version]
  5. Mukherji, A.; Sinisalo, A.; Nüsser, M.; Garrard, R.; Eriksson, M. Contributions of the Cryosphere to Mountain Communities in the Hindu Kush Himalaya: A Review. Reg Env. Chang. 2019, 19, 1311–1326. [Google Scholar] [CrossRef] [Green Version]
  6. Nüsser, M.; Dame, J.; Parveen, S.; Kraus, B.; Baghel, R.; Schmidt, S. Cryosphere-Fed Irrigation Networks in the Northwestern Himalaya: Precarious Livelihoods and Adaptation Strategies Under the Impact of Climate Change. Mt. Res. Dev. 2019, 39, R1–R11. [Google Scholar] [CrossRef] [Green Version]
  7. Nüsser, M.; Schmidt, S.; Dame, J. Irrigation and Development in the Upper Indus Basin: Characteristics and Recent Changes of a Socio-Hydrological System in Central Ladakh, India. Mt. Res. Dev. 2012, 32, 51–61. [Google Scholar] [CrossRef]
  8. Hock, R.; Rasul, G.; Adler, C.; Cáceres, B.; Gruber, S.; Hirabayashi, Y.; Jackson, M.; Kääb, A.; Kang, S.; Kutuzov, S.; et al. High Mountain Areas. In IPCC Special Report on the Ocean and Cryosphere in a Changing Climate; Cambridge University Press: Cambridge, UK, 2019; pp. 131–202. [Google Scholar]
  9. Barnett, T.P.; Adam, J.C.; Lettenmaier, D.P. Potential Impacts of a Warming Climate on Water Availability in Snow-Dominated Regions. Nature 2005, 438, 303–309. [Google Scholar] [CrossRef]
  10. Nüsser, M.; Dame, J.; Kraus, B.; Baghel, R.; Schmidt, S. Socio-Hydrology of “Artificial Glaciers” in Ladakh, India: Assessing Adaptive Strategies in a Changing Cryosphere. Reg. Environ. Chang. 2019, 19, 1327–1337. [Google Scholar] [CrossRef] [Green Version]
  11. Winiger, M.; Gumpert, M.; Yamout, H. Karakorum–Hindukush–Western Himalaya: Assessing High-Altitude Water Resources. Hydrol. Process. 2005, 19, 2329–2338. [Google Scholar] [CrossRef]
  12. Dharpure, J.K.; Goswami, A.; Patel, A.; Kulkarni, A.V.; Snehmani. Assessment of Snow Cover Variability and Its Sensitivity to Hydrometeorological Factors in the Karakoram and Himalayan Region. Hydrol. Sci. J. 2021, 66, 2198–2215. [Google Scholar] [CrossRef]
  13. Immerzeel, W.W.; Droogers, P.; de Jong, S.M.; Bierkens, M.F.P. Large-Scale Monitoring of Snow Cover and Runoff Simulation in Himalayan River Basins Using Remote Sensing. Remote Sens. Environ. 2009, 113, 40–49. [Google Scholar] [CrossRef]
  14. Muhammad, S.; Thapa, A. Daily Terra–Aqua MODIS Cloud-Free Snow and Randolph Glacier Inventory 6.0 Combined Product (M* D10A1GL06) for High-Mountain Asia between 2002 and 2019. Earth Syst. Sci. Data 2021, 13, 767–776. [Google Scholar] [CrossRef]
  15. Salzmann, N.; Huggel, C.; Rohrer, M.; Stoffel, M. Data and Knowledge Gaps in Glacier, Snow and Related Runoff Research—A Climate Change Adaptation Perspective. J. Hydrol. 2014, 518, 225–234. [Google Scholar] [CrossRef] [Green Version]
  16. Gusain, H.; Mishra, V.; Bhutiyani, M. Winter Temperature and Snowfall Trends in the Cryospheric Region of North-West Himalaya. Mausam 2014, 65, 425–432. [Google Scholar] [CrossRef]
  17. Ahmad, I.; Ahmad, Z.; Munir, S.; Shah, S.R.A.; Shabbir, Y. Geo-Spatial Dynamics of Snowcover and Hydro Meteorological Parameters of Astore Basin, UIB, HKH Region, Pakistan. Arab. J. Geosci. 2018, 11, 419. [Google Scholar] [CrossRef]
  18. Bilal, H.; Chamhuri, S.; Mokhtar, M.B.; Kanniah, K.D. Recent Snow Cover Variation in the Upper Indus Basin of Gilgit Baltistan, Hindukush Karakoram Himalaya. J. Mt. Sci. 2019, 16, 296–308. [Google Scholar] [CrossRef]
  19. Dai, L.; Che, T.; Ding, Y.; Hao, X. Evaluation of Snow Cover and Snow Depth on the Qinghai-Tibetan Plateau Derived from Passive Microwave Remote Sensing. Cryosphere 2017, 11, 1933–1948. [Google Scholar] [CrossRef] [Green Version]
  20. Saydi, M.; Ding, J. Impacts of Topographic Factors on Regional Snow Cover Characteristics. Water Sci. Eng. 2020, 13, 171–180. [Google Scholar] [CrossRef]
  21. Sood, V.; Singh, S.; Taloor, A.K.; Prashar, S.; Kaur, R. Monitoring and Mapping of Snow Cover Variability Using Topographically Derived NDSI Model over North Indian Himalayas during the Period 2008–19. Appl. Comput. Geosci. 2020, 8, 100040. [Google Scholar] [CrossRef]
  22. Schmidt, S.; Weber, B.; Winiger, M. Analyses of Seasonal Snow Disappearance in an Alpine Valley from Micro- to Meso-Scale (Loetschental, Switzerland). Hydrol. Process. 2009, 23, 1041–1051. [Google Scholar] [CrossRef]
  23. McKay, G.A.; Gray, D.M. The Distribution of Snowcover. In Handbook of Snow. Principles, Processes, Managment & Use; Gray, D.M., Male, D.H., Eds.; Pergamon Press: Oxford, UK, 1981; pp. 153–190. [Google Scholar]
  24. Sahu, R.; Gupta, R.D. Snow Cover Area Analysis and Its Relation with Climate Variability in Chandra Basin, Western Himalaya, during 2001–2017 Using MODIS and ERA5 Data. Environ. Monit. Assess. 2020, 192, 489. [Google Scholar] [CrossRef] [PubMed]
  25. Dahri, Z.H.; Ludwig, F.; Moors, E.; Ahmad, B.; Khan, A.; Kabat, P. An Appraisal of Precipitation Distribution in the High-Altitude Catchments of the Indus Basin. Sci. Total Environ. 2016, 548–549, 289–306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Norman, R.; Thoma, H. Rang-Skat. A Ladakhi-English Dictionary; Melong Publications: Leh, India, 2019. [Google Scholar]
  27. Schmidt, S.; Nüsser, M. Changes of High Altitude Glaciers in the Trans-Himalaya of Ladakh over the Past Five Decades (1969–2016). Geosciences 2017, 7, 27. [Google Scholar] [CrossRef] [Green Version]
  28. Schmidt, S.; Nüsser, M. Changes of High Altitude Glaciers from 1969 to 2010 in the Trans-Himalayan Kang Yatze Massif, Ladakh, Northwest India. Arct. Antarct. Alp. Res. 2012, 44, 107–121. [Google Scholar] [CrossRef]
  29. Soheb, M.; Ramanathan, A.; Bhardwaj, A.; Sam, L. Spatiotemporal Quantification of Key Environmental Changes in Stok and Kang Yatze Regions of Ladakh Himalaya, India. Geocarto Int. 2022, 1–22. [Google Scholar] [CrossRef]
  30. Soheb, M.; Ramanathan, A.; Angchuk, T.; Mandal, A.; Kumar, N.; Lotus, S. Mass-Balance Observation, Reconstruction and Sensitivity of Stok Glacier, Ladakh Region, India, between 1978 and 2019. J. Glaciol. 2020, 66, 627–642. [Google Scholar] [CrossRef]
  31. Bhambri, R.; Bolch, T.; Kawishwar, P.; Dobhal, D.; Srivastava, D.; Pratap, B. Heterogeneity in Glacier Response in the Upper Shyok Valley, Northeast Karakoram. Cryosphere 2013, 7, 1385–1398. [Google Scholar] [CrossRef] [Green Version]
  32. Bhutiyani, M. Mass-Balance Studies on Siachen Glacier in the Nubra Valley, Karakoram Himalaya, India. J. Glaciol. 1999, 45, 112–118. [Google Scholar] [CrossRef]
  33. Kamp, U.; Byrne, M.; Bolch, T. Glacier Fluctuations between 1975 and 2008 in the Greater Himalaya Range of Zanskar, Southern Ladakh. J. Mt. Sci. 2011, 8, 374–389. [Google Scholar] [CrossRef]
  34. Brombierstäudl, D.; Schmidt, S.; Nüsser, M. Distribution and Relevance of Aufeis (Icing) in the Upper Indus Basin. Sci. Total Environ. 2021, 780, 146604. [Google Scholar] [CrossRef] [PubMed]
  35. Pfister, O. Birds and Mammals of Ladakh; Oxford University Press: Oxford, UK, 2004. [Google Scholar]
  36. Riggs, G.A.; Hall, D.K.; Salomonson, V.V. MODIS Snow Products User Guide to Collection 5. Digit. Media 2006, 80, 1–80. [Google Scholar]
  37. Shean, D. High Mountain Asia 8-Meter DEM Mosaics Derived from Optical Imagery; Version 1; NASA National Snow and Ice Data Center Distributed Active Archive Center: Boulder, CO, USA, 2017. [Google Scholar]
  38. Riggs, G.A.; Hall, D.K.; Román, M.O. MODIS Snow Products Collection 6 User Guide; National Snow and Ice Data Center: Boulder, CO, USA, 2015; 66p. [Google Scholar]
  39. Zhang, H.; Zhang, F.; Zhang, G.; Che, T.; Yan, W.; Ye, M.; Ma, N. Ground-Based Evaluation of MODIS Snow Cover Product V6 across China: Implications for the Selection of NDSI Threshold. Sci. Total Environ. 2019, 651, 2712–2726. [Google Scholar] [CrossRef] [PubMed]
  40. Jain, S.K.; Goswami, A.; Saraf, A.K. Accuracy Assessment of MODIS, NOAA and IRS Data in Snow Cover Mapping under Himalayan Conditions. Int. J. Remote Sens. 2008, 29, 5863–5878. [Google Scholar] [CrossRef]
  41. Singh, S.K.; Rathore, B.P.; Bahuguna, I.M.; Ajai. Snow Cover Variability in the Himalayan-Tibetan Region: Snow cover in Himalayan-Tibetan region. Int. J. Climatol. 2014, 34, 446–452. [Google Scholar] [CrossRef]
  42. Kulkarni, A.V.; Rathore, B.P.; Singh, S.K.; Ajai. Distribution of seasonal snow cover in central and western Himalaya. Ann. Glaciol. 2010, 51, 123–128. [Google Scholar] [CrossRef] [Green Version]
  43. Zhang, Y.; Yan, S.; Lu, Y. Snow Cover Monitoring Using MODIS Data in Liaoning Province, Northeastern China. Remote Sens. 2010, 2, 777–793. [Google Scholar] [CrossRef] [Green Version]
  44. Hall, D.K.; Riggs, G.A.; Salomonson, V.V.; Barton, J.; Casey, K.; Chien, J.; DiGirolamo, N.; Klein, A.; Powell, H.; Tait, A. Algorithm Theoretical Basis Document (ATBD) for the MODIS Snow and Sea Ice-Mapping Algorithms; NASA GSFC: Greenbelt, MD, USA, 2001; 45p. [Google Scholar]
  45. Salomonson, V.V.; Appel, I. Estimating Fractional Snow Cover from MODIS Using the Normalized Difference Snow Index. Remote Sens. Environ. 2004, 89, 351–360. [Google Scholar] [CrossRef]
  46. Tran, H.; Nguyen, P.; Ombadi, M.; Hsu, K.; Sorooshian, S.; Qing, X. A Cloud-Free MODIS Snow Cover Dataset for the Contiguous United States from 2000 to 2017. Sci. Data 2019, 6, 180300. [Google Scholar] [CrossRef]
  47. Zhang, Y.; Cao, T.; Kan, X.; Wang, J.; Tian, W. Spatial and Temporal Variation Analysis of Snow Cover Using MODIS over Qinghai-Tibetan Plateau during 2003–2014. J. Indian Soc. Remote Sens. 2017, 45, 887–897. [Google Scholar] [CrossRef]
  48. Chokmani, K.; Dever, K.; Bernier, M.; Gauthier, Y.; Paquet, L.-M. Adaptation of the SNOWMAP Algorithm for Snow Mapping over Eastern Canada Using Landsat-TM Imagery. Hydrol. Sci. J. 2010, 55, 649–660. [Google Scholar] [CrossRef]
  49. Hall, D.K.; Riggs, G.A.; Salomonson, V.V.; DiGirolamo, N.E.; Bayr, K.J. MODIS Snow-Cover Products. Remote Sens. Environ. 2002, 83, 181–194. [Google Scholar] [CrossRef] [Green Version]
  50. Hall, D.K.; Riggs, G.A.; Salomonson, V.V. Development of Methods for Mapping Global Snow Cover Using Moderate Resolution Imaging Spectroradiometer Data. Remote Sens. Environ. 1995, 54, 127–140. [Google Scholar] [CrossRef]
  51. Cohen, J. A Coefficient of Agreement for Nominal Scales. Educ. Psychol. Meas. 1960, 20, 37–46. [Google Scholar] [CrossRef]
  52. Hall, D.K.; Riggs, G.A. Accuracy Assessment of the MODIS Snow Products. Hydrol. Process. 2007, 21, 1534–1547. [Google Scholar] [CrossRef]
  53. López-Burgos, V.; Gupta, H.V.; Clark, M. Reducing Cloud Obscuration of MODIS Snow Cover Area Products by Combining Spatio-Temporal Techniques with a Probability of Snow Approach. Hydrol. Earth Syst. Sci. 2013, 17, 1809–1823. [Google Scholar] [CrossRef] [Green Version]
  54. Parajka, J.; Pepe, M.; Rampini, A.; Rossi, S.; Blöschl, G. A Regional Snow-Line Method for Estimating Snow Cover from MODIS during Cloud Cover. J. Hydrol. 2010, 381, 203–212. [Google Scholar] [CrossRef]
  55. Parajka, J.; Blöschl, G. Spatio-Temporal Combination of MODIS Images–Potential for Snow Cover Mapping. Water Resour. Res. 2008, 44, W03406. [Google Scholar] [CrossRef]
  56. Gafurov, A.; Bárdossy, A. Cloud Removal Methodology from MODIS Snow Cover Product. Hydrol. Earth Syst. Sci. 2009, 13, 1361–1373. [Google Scholar] [CrossRef] [Green Version]
  57. Gao, Y.; Xie, H.; Yao, T.; Xue, C. Integrated Assessment on Multi-Temporal and Multi-Sensor Combinations for Reducing Cloud Obscuration of MODIS Snow Cover Products of the Pacific Northwest USA. Remote Sens. Environ. 2010, 114, 1662–1675. [Google Scholar] [CrossRef]
  58. Paudel, K.P.; Andersen, P. Monitoring Snow Cover Variability in an Agropastoral Area in the Trans Himalayan Region of Nepal Using MODIS Data with Improved Cloud Removal Methodology. Remote Sens. Environ. 2011, 115, 1234–1246. [Google Scholar] [CrossRef]
  59. Hauksson, B.Þ. Estimation of Regional Snow-Line Elevations in Norway. Testing a Method for Estimating RSLE and Melt Slope Analysis in Norway. Master’s Thesis, University of Oslo, Oslo, Norway, 2017. [Google Scholar]
  60. Krajčí, P.; Holko, L.; Perdigão, R.A.; Parajka, J. Estimation of Regional Snowline Elevation (RSLE) from MODIS Images for Seasonally Snow Covered Mountain Basins. J. Hydrol. 2014, 519, 1769–1778. [Google Scholar] [CrossRef]
  61. Kendall, M. Rank Correlation Methods; Charles Griffin: London, UK, 1975. [Google Scholar]
  62. Mann, H.B. Nonparametric Tests against Trend. Econom. J. Econom. Soc. 1945, 13, 245–259. [Google Scholar] [CrossRef]
  63. Sen, P.K. Estimates of the Regression Coefficient Based on Kendall’s Tau. J. Am. Stat. Assoc. 1968, 63, 1379–1389. [Google Scholar] [CrossRef]
  64. Hamed, K.H.; Rao, A.R. A Modified Mann-Kendall Trend Test for Autocorrelated Data. J. Hydrol. 1998, 204, 182–196. [Google Scholar] [CrossRef]
  65. Yue, S.; Wang, C. The Mann-Kendall Test Modified by Effective Sample Size to Detect Trend in Serially Correlated Hydrological Series. Water Resour. Manag. 2004, 18, 201–218. [Google Scholar] [CrossRef]
  66. Gascoin, S.; Hagolle, O.; Huc, M.; Jarlan, L.; Dejoux, J.-F.; Szczypta, C.; Marti, R.; Sánchez, R. A Snow Cover Climatology for the Pyrenees from MODIS Snow Products. Hydrol. Earth Syst. Sci. 2015, 19, 2337–2351. [Google Scholar] [CrossRef] [Green Version]
  67. Klein, A.G.; Barnett, A.C. Validation of Daily MODIS Snow Cover Maps of the Upper Rio Grande River Basin for the 2000–2001 Snow Year. Remote Sens. Environ. 2003, 86, 162–176. [Google Scholar] [CrossRef]
  68. Liang, T.G.; Huang, X.D.; Wu, C.X.; Liu, X.Y.; Li, W.L.; Guo, Z.G.; Ren, J.Z. An Application of MODIS Data to Snow Cover Monitoring in a Pastoral Area: A Case Study in Northern Xinjiang, China. Remote Sens. Environ. 2008, 112, 1514–1526. [Google Scholar] [CrossRef]
  69. Zhang, Y.; Kan, X.; Ren, W.; Cao, T.; Tian, W.; Wang, J. Snow Cover Monitoring in Qinghai-Tibetan Plateau Based on Chinese Fengyun-3/VIRR Data. J. Indian Soc. Remote Sens. 2017, 45, 271–283. [Google Scholar] [CrossRef]
  70. Girona-Mata, M.; Miles, E.S.; Ragettli, S.; Pellicciotti, F. High-Resolution Snowline Delineation from Landsat Imagery to Infer Snow Cover Controls in a Himalayan Catchment. Water Resour. Res. 2019, 55, 6754–6772. [Google Scholar] [CrossRef] [Green Version]
  71. Aschenwald, J.; Leichter, K.; Tasser, E.; Tappeiner, U. Spatio-Temporal Landscape Analysis in Mountainous Terrain by Means of Small Format Photography: A Methodological Approach. IEEE Trans. Geosci. Remote Sens. 2001, 39, 885–893. [Google Scholar] [CrossRef]
  72. Corripio, J. Snow Surface Albedo Estimation Using Terrestrial Photography. Int. J. Remote Sens. 2004, 25, 5705–5729. [Google Scholar] [CrossRef]
  73. Tahir, A.A.; Chevallier, P.; Arnaud, Y.; Ashraf, M.; Bhatti, M.T. Snow Cover Trend and Hydrological Characteristics of the Astore River Basin (Western Himalayas) and Its Comparison to the Hunza Basin (Karakoram Region). Sci. Total Environ. 2015, 505, 748–761. [Google Scholar] [CrossRef] [PubMed]
  74. Hasson, S.; Lucarini, V.; Khan, M.R.; Petitta, M.; Bolch, T.; Gioli, G. Early 21st Century Snow Cover State over the Western River Basins of the Indus River System. Hydrol. Earth Syst. Sci. 2014, 18, 4077–4100. [Google Scholar] [CrossRef] [Green Version]
  75. Miehe, G.; Winiger, M.; Böhner, J.; Yili, Z. The Climatic Diagram Map of High Asia. Erdkunde 2001, 55, 94–98. [Google Scholar] [CrossRef]
  76. Dimri, A.P.; Chevuturi, A. Model Sensitivity Analysis Study for Western Disturbances over the Himalayas. Meteorol. Atmos. Phys. 2014, 123, 155–180. [Google Scholar] [CrossRef]
  77. Gurung, D.R.; Maharjan, S.B.; Shrestha, A.B.; Singh Shrestha, M.; Bajracharya, S.R.; Murthy, M.S.R. Climate and Topographic Controls on Snow Cover Dynamics in the Hindu Kush Himalaya. Int. J. Climatol. 2017, 37, 3873–3882. [Google Scholar] [CrossRef] [Green Version]
  78. Richter, M. Klimatologische und pflanzenmorphologische Vertikalgradienten in Hochgebirgen. Erdkunde 1996, 50, 205–237. [Google Scholar] [CrossRef]
  79. Nüsser, M.; Baghel, R. The Emergence of the Cryoscape: Contested Narratives of Himalayan Glacier Dynamics and Climate Change. In Environmental and Climate Change in South and Southeast Asia; Schuler, B., Ed.; Brill: Leiden, The Netherlands, 2014; pp. 138–156. [Google Scholar]
Figure 1. The Trans-Himalayan region of Ladakh, sandwiched between the Greater Himalayas, the Karakoram, and the Tibetan plateau. Place names follow the spelling in Ladakhi language [26], e.g., Zangskar (Zanskar), Tshomoriri (Tso Moriri), and Siachan (Siachen).
Figure 1. The Trans-Himalayan region of Ladakh, sandwiched between the Greater Himalayas, the Karakoram, and the Tibetan plateau. Place names follow the spelling in Ladakhi language [26], e.g., Zangskar (Zanskar), Tshomoriri (Tso Moriri), and Siachan (Siachen).
Geosciences 12 00311 g001
Figure 2. Topographical parameters (spatial resolution of 500 m) of the study area (75°32′ E, 35°67′ N to 79°46′ E, 32°34′ N): elevation (left), slope (middle), and aspect (right) (Table 1).
Figure 2. Topographical parameters (spatial resolution of 500 m) of the study area (75°32′ E, 35°67′ N to 79°46′ E, 32°34′ N): elevation (left), slope (middle), and aspect (right) (Table 1).
Geosciences 12 00311 g002
Figure 3. Locations and viewsheds of four terrestrial cameras. Gyamtsa (red) and Lungmar (orange) were used for MODIS validation in the present study.
Figure 3. Locations and viewsheds of four terrestrial cameras. Gyamtsa (red) and Lungmar (orange) were used for MODIS validation in the present study.
Geosciences 12 00311 g003
Figure 4. Terrestrial photograph taken from Gyamtsa on 23 January 2017 (a), overlaid with snow polylines (red lines) and control points (red crosses) derived from Landsat-8 OLI image (b) and MODIS (c). Gyamtsa camera view shed in red colour (d). Landsat-8 OLI (e) and MODIS snow shown in cyan (f). Terrestrial photograph taken from Lungmar on 21 January 2016 (g), overlaid with snow polylines (red lines) derived from Landsat-8 OLI (h) and MODIS (i). Lungmar view shed in orange colour (j). Landsat-8 OLI (k) and MODIS snow (l).
Figure 4. Terrestrial photograph taken from Gyamtsa on 23 January 2017 (a), overlaid with snow polylines (red lines) and control points (red crosses) derived from Landsat-8 OLI image (b) and MODIS (c). Gyamtsa camera view shed in red colour (d). Landsat-8 OLI (e) and MODIS snow shown in cyan (f). Terrestrial photograph taken from Lungmar on 21 January 2016 (g), overlaid with snow polylines (red lines) derived from Landsat-8 OLI (h) and MODIS (i). Lungmar view shed in orange colour (j). Landsat-8 OLI (k) and MODIS snow (l).
Geosciences 12 00311 g004
Figure 5. Flow chart of MODIS snow data processing validated with Landsat-8 imagery and orthorectified photos.
Figure 5. Flow chart of MODIS snow data processing validated with Landsat-8 imagery and orthorectified photos.
Geosciences 12 00311 g005
Figure 6. Number of land (green), snow (cyan), or cloud-covered (grey) pixels in relation to elevation (left) on a free temporal combination MODIS raster image on 16 January 2017. (Right): Spatio-temporal combination of MODIS image of 16 January 2017 after cloud reduction step 3. Cloud-covered pixels below RLLE (4433 m a.s.l.; black dashed line) are classified as land, between RSLE and RLLE (4919 m a.s.l.; red dashed line) as mixed pixels (yellow), and above RSLE as snow.
Figure 6. Number of land (green), snow (cyan), or cloud-covered (grey) pixels in relation to elevation (left) on a free temporal combination MODIS raster image on 16 January 2017. (Right): Spatio-temporal combination of MODIS image of 16 January 2017 after cloud reduction step 3. Cloud-covered pixels below RLLE (4433 m a.s.l.; black dashed line) are classified as land, between RSLE and RLLE (4919 m a.s.l.; red dashed line) as mixed pixels (yellow), and above RSLE as snow.
Geosciences 12 00311 g006
Figure 7. Monthly mean percentage of the cloud-covered pixels of MODIS daily snow product data featuring the reduction of cloud coverage (black line) and overall mean percentages (blue line) in three cloud reduction steps.
Figure 7. Monthly mean percentage of the cloud-covered pixels of MODIS daily snow product data featuring the reduction of cloud coverage (black line) and overall mean percentages (blue line) in three cloud reduction steps.
Geosciences 12 00311 g007
Figure 8. Monthly average percentages (black line) and overall monthly mean (blue line) of SCA for the entire study area of Ladakh over the period of 2000–2019.
Figure 8. Monthly average percentages (black line) and overall monthly mean (blue line) of SCA for the entire study area of Ladakh over the period of 2000–2019.
Geosciences 12 00311 g008
Figure 9. Pixel-wise analysis of snow cover days (SCD) for the entire study area of Ladakh over the period of 2000–2019. Mean seasonal and annual SCD (top), seasonal and annual trends of SCD (a) autumn: September–November, (b) winter: December–February, (c) spring: March–May, (d) summer: June–August, and (e) annual (hydrological year: October–September) (centre). SCD trends with significance level p < 0.1, yellow colour symbolizes no significant trend, red negative, and blue colour positive trends (bottom).
Figure 9. Pixel-wise analysis of snow cover days (SCD) for the entire study area of Ladakh over the period of 2000–2019. Mean seasonal and annual SCD (top), seasonal and annual trends of SCD (a) autumn: September–November, (b) winter: December–February, (c) spring: March–May, (d) summer: June–August, and (e) annual (hydrological year: October–September) (centre). SCD trends with significance level p < 0.1, yellow colour symbolizes no significant trend, red negative, and blue colour positive trends (bottom).
Geosciences 12 00311 g009
Figure 10. Seasonal and annual MODIS SCA and trends in relation to elevation zones of all subregions of Ladakh for the observation period of 2000–2019. p-value significance levels are shown in red (p < 0.01), blue (p < 0.05), and green (p < 0.1) triangles (see also Table A2).
Figure 10. Seasonal and annual MODIS SCA and trends in relation to elevation zones of all subregions of Ladakh for the observation period of 2000–2019. p-value significance levels are shown in red (p < 0.01), blue (p < 0.05), and green (p < 0.1) triangles (see also Table A2).
Geosciences 12 00311 g010
Figure 11. Seasonal MODIS SCA in relation to aspect across all subregions of the study area over the observation period of 2000–2019.
Figure 11. Seasonal MODIS SCA in relation to aspect across all subregions of the study area over the observation period of 2000–2019.
Geosciences 12 00311 g011
Figure 12. Seasonal and annual MODIS SCA and trends in relation to slope across all subregions over the observation period of 2000–2019.
Figure 12. Seasonal and annual MODIS SCA and trends in relation to slope across all subregions over the observation period of 2000–2019.
Geosciences 12 00311 g012
Table 1. Spatial distribution of elevation, slope, and aspect in the seven subregions of Ladakh.
Table 1. Spatial distribution of elevation, slope, and aspect in the seven subregions of Ladakh.
Siachan
(Siachen)
Shayok
(Shyok)
Pangong
Indus
Suru
Zangskar
(Zanskar)
Tshomoriri
(Tso Moriri)
km2%km2%km2%km2%km2%km2%km2%
Elevation (m a.s.l.)
<350026335125--10676740111571--
3500–45008458180817546215476322946423365231324
4500–5500489847563853138354829748312245987267244173
5500–65004333412622256142422931314821298974923
>6500115110--10400010
Slope (°)
<10321631220221109043561833111816312221163349
10–20334632397538106042761644269239607541137741
20–30259525341932370153565212468364575313039
30–401176119299231333260098866100
>401191541--20371340--
Aspect
N12321212851223192076129591415931132010
NE156015147314445172566158651321771542713
E134813155615369142217138651320481444513
SE124712118811295121942117851116391132810
S14101311121122491890117991215331037111
SW149014126012333132374148201219711349015
W113411138513299122240138771319931446314
NW10341013161223691834119451417391238111
Table 2. Annual and seasonal Mann–Kendall test statistics of the mean SCA (%) in the study area of Ladakh from 2000 to 2019.
Table 2. Annual and seasonal Mann–Kendall test statistics of the mean SCA (%) in the study area of Ladakh from 2000 to 2019.
Linear SlopeSen’s SlopeZ-Valuep-ValueTrendMeanSD
Annual0.090.080.560.5841.854.32
Autumn0.12−0.0101.0028.936.50
Winter0.03−0.10−0.210.8361.176.49
Spring−0.01−0.07−0.280.7857.536.33
Summer0.220.200.980.3319.903.95
SD: Standard deviation, : positive trend, : negative trend.
Table 3. Increasing and decreasing trend of SCD in the study area of Ladakh from 2000 to 2019.
Table 3. Increasing and decreasing trend of SCD in the study area of Ladakh from 2000 to 2019.
Increasing Trend (Sen’s Slope > 0)
Allp < 0.01p < 0.05p < 0.1
km2%km2%km2%km2%
Autumn23,538.5035.83126.000.19710.001.081571.002.39
Winter21,587.7532.86103.750.16559.500.851280.751.95
Spring25,163.2538.31205.750.311011.001.541993.003.03
Summer24,815.0037.78725.751.104269.006.507853.7511.96
Annual33,409.5050.86346.500.531934.752.954149.256.32
Decreasing Trend (Sen’s Slope < 0)
Allp < 0.01p < 0.05p < 0.1
km2%km2%km2%km2%
Autumn9907.7515.0811.750.0248.750.07136.500.21
Winter20,396.0031.0536.500.06399.750.611027.251.56
Spring12,440.5018.9447.750.07365.000.56811.251.23
Summer3777.255.75152.250.23447.000.68731.501.11
Annual11,983.2518.24230.500.35743.251.131220.001.86
p: Significance level.
Table 4. Selected studies focusing on snow cover distribution and trends in the high mountain regions of South Asia.
Table 4. Selected studies focusing on snow cover distribution and trends in the high mountain regions of South Asia.
RegionObservation PeriodSCA and TrendsTopicsReference
UIB
(Gilgit Baltistan)
2000–2017Max: 86% in February/March, Min: 36% in August
Slight non-significant increasing trend
Elevation zones[18]
Chandra basin
(Himachal Pradesh)
2001–2017Max: 99% in February, Min: 60% in AugustElevation, slope, aspect[24]
Chenab basin
Satluj basin
Ravi basin
Beas basin
(Himachal Pradesh)
2003–200442% mean annual
23% mean annual
33% mean annual
38% mean annual
Elevation, aspect[2]
UIB2001–2012Annual SCA: non-significant, slightly decreasing trend for UIB and all subbasins, except GilgitElevation, aspect, NAO[74]
Jhelum and Kabul basins2001–2012Annual SCA: non-significant, slightly increasing trend for Jhelum and Kabul basinsElevation, aspect, NAO[74]
Astore basin
(Gilgit Baltistan)
2000–2012Max: 95% in January, Min: 7% in August
stable (tends to slight increase) trend
Elevation zones[73]
Karakoram (KK)
Western Himalaya (WH)
Central Himalaya (CH)
Eastern Himalaya (EH)
Karakoram-Himalaya (KH)
2000–201952% mean annual
31% mean annual
13% mean annual
10% mean annual
26% mean annual
Non-significant, increasing trend for KK, WH, CH
Non-significant, decreasing trend for EH, KH
Meteorological variables[12]
Karakoram
Western Himalaya
Central Himalaya
Eastern Himalaya
Karakoram-Himalaya
2000–200852% mean annual
30% mean annual
12% mean annual
11% mean annual
26% mean annual
Non-significant, increasing trend for all subregions and the entire KH
Meteorological variables[12]
Karakoram
Western Himalaya
Central Himalaya
Eastern Himalaya
Karakoram-Himalaya
2008–201852% mean annual
31% mean annual
12% mean annual
10% mean annual
26% mean annual
decreasing trend for all subregions and the entire KH, which is significant for WH and KH
Meteorological variables[12]
UIB2000–2008negative trend for winter snow cover; no trend for all other seasons and the entire yearElevation zones, runoff[13]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Passang, S.; Schmidt, S.; Nüsser, M. Topographical Impact on Snow Cover Distribution in the Trans-Himalayan Region of Ladakh, India. Geosciences 2022, 12, 311. https://doi.org/10.3390/geosciences12080311

AMA Style

Passang S, Schmidt S, Nüsser M. Topographical Impact on Snow Cover Distribution in the Trans-Himalayan Region of Ladakh, India. Geosciences. 2022; 12(8):311. https://doi.org/10.3390/geosciences12080311

Chicago/Turabian Style

Passang, Stanzin, Susanne Schmidt, and Marcus Nüsser. 2022. "Topographical Impact on Snow Cover Distribution in the Trans-Himalayan Region of Ladakh, India" Geosciences 12, no. 8: 311. https://doi.org/10.3390/geosciences12080311

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