Advances in Environmental and Engineering Research (AEER) is an international peer-reviewed Open Access journal published quarterly online by LIDSEN Publishing Inc. This periodical is devoted to publishing high-quality peer-reviewed papers that describe the most significant and cutting-edge research in all areas of environmental science and engineering. Work at any scale, from molecular biology through to ecology, is welcomed.

Main research areas include (but are not limited to):

  • Atmospheric pollutants
  • Air pollution control engineering
  • Climate change
  • Ecological and human risk assessment
  • Environmental management and policy
  • Environmental impact and risk assessment
  • Environmental microbiology
  • Ecosystem services, biodiversity and natural capital
  • Environmental economics
  • Control and monitoring of pollutants
  • Remediation of polluted soils and water
  • Fate and transport of contaminants
  • Water and wastewater treatment engineering
  • Solid waste treatment

Advances in Environmental and Engineering Research publishes a variety of article types (Original Research, Review, Communication, Opinion, Comment, Conference Report, Technical Note, Book Review, etc.). We encourage authors to be succinct; however, authors should present their results in as much detail as necessary. Reviewers are expected to emphasize scientific rigor and reproducibility.

Indexing: 

Publication Speed (median values for papers published in 2023): Submission to First Decision: 6.1 weeks; Submission to Acceptance: 16.1 weeks; Acceptance to Publication: 9 days (1-2 days of FREE language polishing included)

Current Issue: 2024  Archive: 2023 2022 2021 2020
Open Access Original Research

Evaluation of Active Tectonic Features of Nandakini River Basin, Lesser Himalaya, India by Using Morphometric Indices: A GIS Approach

Pawan Kumar Gautam †,*, Anoop Kumar Singh 

Department of Geology, University of Lucknow, Lucknow – 226007, U.P., India

† These authors contributed equally to this work.

Correspondence: Pawan Kumar Gautam

Academic Editor: Leonel J. R. Nunes

Special Issue: Environmental Risk Assessment and Risk Management

Received: July 11, 2022 | Accepted: January 03, 2023 | Published: January 30, 2023

Adv Environ Eng Res 2023, Volume 4, Issue 1, doi:10.21926/aeer.2301014

Recommended citation: Gautam PK, Singh AK. Evaluation of Active Tectonic Features of Nandakini River Basin, Lesser Himalaya, India by Using Morphometric Indices: A GIS Approach. Adv Environ Eng Res 2023; 4(1): 014; doi:10.21926/aeer.2301014.

© 2023 by the authors. This is an open access article distributed under the conditions of the Creative Commons by Attribution License, which permits unrestricted use, distribution, and reproduction in any medium or format, provided the original work is correctly cited.

Abstract

Tectonic geomorphology has been evidenced as an essential tool to define and measure recent tectonic deformation. The Nandakini River originates from Nanda Ghunti glacier at 6886 m in Lesser Himalaya. It covers about ~551 km2 basin area, with a fourth-order stream. It confluences with the Alaknanda River near Nand Prayag (at elevation 852 m), Chamoli district of Uttarakhand. The drainage pattern is predominantly dendritic in the study area. The morphotectonic indices are measured to know the tectonic activity of the drainage basin. Morphometric indices reflecting hypsometric Integral is 1.07, indicating a deeper incision and a slight erosion. The basin elongation ratio is 0.17, suggesting that the area is tectonically active. The drainage basin asymmetric factor (44) suggest tilting. In addition, the drainage basin shape (3.27) indicates the basin is tectonically strong, the transverse topographic symmetry factor (0.31) indicates the asymmetric nature, the valley floor width to valley height ratio is 0.54 shows the deep, narrow valleys, and Active V-shaped incision. Stream longitudinal profile showing the area tectonically influenced. The Sinuosity Index of 1.12 suggests that low sinuous nature. The Stream length-gradient Index is 10.64, indicating high tectonic activity in valleys and basin areas. The transverse profile helps to understand the tilting of the basin. The Surface profile shows irregularity, and the linear trend of the profile shows the tilting of the basin. River carrying capacity with large boulder-size sediment indicates youth stage and tectonically active environment. The morphotectonic indices suggest that the drainage basin was affected by the regional structures and the present tectonics.

Keywords

Lesser Himalaya; Nandakini River basin; morphotectonic indices; GIS

1. Introduction

The major rivers of Northern India originated from the great Himalaya [1]. During the last decades, geoscientists have paid much attention to understanding the tectonic design of the Kumaun-lesser Himalaya in particular and the Himalaya in general [2,3,4,5,6,7,8,9,10,11,12,13]. The Himalaya are amongst the most excellent and identifiable structures of tectonic forces extending 2,900 km along the boundary between India and Tibet. The structural (tectonic) Himalayas can be divided from north to south, the Eastern Himalayas, the Central Himalayas, and the Western Himalayas as follows [14].

The Himalayan region is tectonically and seismically active; therefore, investigations of neotectonic evidence mainly affect geomorphology [1,15,16,17]. The continuous north-western motion of the Indian Plate at a velocity of about 5 cm/year indicates about 2500 km of post-collision shortening has occurred due to Indian and Eurasian plate collision during 50-55 Million years [18,19,20,21,22]. The Himalayan region's rugged topography is reshaped by neotectonic activity, such as thrusts/faults, resulting in geomorphic changes along the banks of all major rivers [11,23]. The present study area falls in the central part of the Himalayan belt in the Garhwal Himalaya. It is important for studying the distinctive features of the Himalayan fold-and-thrust belt [24]. The Garhwal Lesser Himalaya is cut by various thrusts/faults, e.g., MBT, MCT, HFT, Munsiari Thrust (MT), Ramgarh Thrust (RT), Tons Thrust (TT)/Berinag Thrust (BT) and STDS. This middle latitudinal domain, tectonically bounded at the top by the MCT and base by the MBT, comprises mainly Precambrian clastic sediments and meta-sedimentary rocks [25,26]. The MCT zone has the largest width of about 80 km in the Kumaun and Garhwal regions [27]. The Garhwal Lesser Himalayan area is situated on the west side of the Himalayas and surrounded by Main Central Thrust (MCT) in the north, and Main Boundary Thrust (MBT) in South and South West, Srinagar Thrust in the east, west, and center of the network. Most of the higher magnitude earthquakes (Magnitude Scale-5.5) in the Himalaya occurred around the (MCT) due to the thrust-flawed separation zone and the northern Indian plate. Different aspects of seismo-tectonic and seismicity have been studied using local seismological and strong resistance networks [9,28,29,30,31,32]. Gharwal Himalayas is situated near the central part of the Himalayan origins. Therefore, the region is affected by the Himalayan tectonic as compared to the areas close to EW in the direction. Many geomorphological features describe tectonically active regions on the earth to know the recent tectonic processes during 100 Ka [33]. There has been evidence of neotectonics and reactivation of faults and thrusts throughout the North West Himalayas [34]. Tecto-geomorphology analyzes landforms created by tectonic processes or the various applications of geomorphic concepts to solve tectonic problems [35]. Topographic-based geomorphic indices are developed as essential techniques for recognizing areas with rapid deformity of the tectonic [35,36]. It deals with active Cenozoic deformation’s tectonic and geomorphological process interactions [37]. Drainage basins is a dynamic system that may possess records of formation and evolution since its most tectonic-geomorphic processes to occur within its limits [38]. Thus, landscape expansion results from the evolution of drainage basins [39]. Geomorphic features of drainage basins and stream networks can be represented adequately using Digital Elevation Models (DEM) and Geographical Information Systems (GIS) [12,40,41,42]. These tools allow the rapid, accurate, and scientifically provable of fluvial channels and landscape characterization [43,44,45,46,47,48,49,50,51]. The Nandakini basin lies between MBT and MCT at the northern Baijnath klippe, which is regarded as one of the active thrusts in the Indian Himalaya [52,53,54,55,56,57,58]. The geomorphic indices are a dominant tool for appraising the effect of active tectonics [32]. This study aims to evaluate the tectonic zones using morphometric indices. The objective will be to help prepare and manage probable disasters like landslides and select suitable sites for settlements and transportation.

2. Study Area

Nandakini river originates from Nanda Ghunti glacier at 6886 m altitude in the Lesser Himalaya. It is joins in Alaknanda River at 852 m altitude near Nand Prayag, Chamoli district of Uttarakhand. Geographically it lies between 30°10′ to 30°22′ N latitude and 79°18′ to 79°47′ E longitude. The geographical area of the Nandakini basin is ~551 km2 (Figure 1). The study area is characterized by several geomorphic features like long ridges, rugged topography, steep slopes, stepped river terraces, and dense forests.

Click to view original image

Figure 1 Location map of the Study area.

2.1 Climate

The study area comes under temperate climates. The climate here is mostly cold and sub-humid, with frigid winters from December to February and mild hot from March to May. Rainfall in the region is highly variable and depends on the altitude. Monsoon starts in June with westerly winds and receives about 1100 to 1305 mm rainfall [59]. The area receives maximum rainfall between June to September and post-monsoons during October and November. Rain swiftly decreases at the end of September and less in November. Very occasionally, some area experience snowfall in January.

2.2 Geological Settings

The basin is surrounded by hills and deep valleys that effectively disrupt and are associated with a complex geological setting. Most of the basin, the Baijnath klippe, in the west, the klippen represent the eroded remnants of the Ramgarh and Munsiari thrust sheets, which are correlated with the thrust units underlying the Vaikrita thrust. As per structure, the Lesser Himalaya Zone reflects a duplex thrust system consisting of horses of Berinag, Chakrata–Rautgara, and Deoban Formations overlain by the crystalline Ramgarh along the Ramgarh Thrust as roof thrust and the MHT as the floor thrust [24,60]. The MCT, representing a shear zone, brings the HHC to overlain by Lesser Himalaya. [13] states, the HHC, and the Vaikrita Group underlain by the Munsiari and Ramgarh nappes (Table 1, Figure 2).

Table 1 Geology of Study area [13].

Click to view original image

Figure 2 (a) Tectonic map of the Uttarakhand (black rectangle indicates the area of study [61]. (b) Geological map of the Nandakini River Basin [13].

Munsiari thrust along the Alaknanda River, where a window through the Berinag thrust hanging wall has exposed underlying dolomites and black schists of the Deoban and Mandhali formations [61]. The north of NAT and BT (Berinag Thrust) in the Lesser Himalaya and up to the Higher Himalaya show maximum deformation [62]. The bedrock refers to the Central Crystalline [8] composed primarily of medium-grade to high-grade metamorphic a derivative of pelitic, semi-pelitic, and psammitic sediments, which are interlayer sporadically with the meta basics and, to a lesser range (i.e., in the Dhak and Jharkula regions), calcareous rocks. In the south, the Central Crystalline’ is overlain by the Garhwal Group through the MCT [13,61].

2.3 Seismicity

The micro-seismicity recorded instrumentally by the Wadia Institute of Himalayan Geology network shows the clustering of events along a transitional physiographic region between the High and Lesser Himalaya and along the Main Central Thrust zone (MCT) [63]. The widespread occurrence of small and moderate earthquakes along MCT and subsidiary thrusts between the MBT and MCT [64]. Uttarakhand was struck by two Mild Earthquakes, the 1991 Uttarkashi earthquake, and the 1999 Chamoli earthquake. There is a historical account of the 1803 earthquake causing severe damage and destruction primarily in Garhwal and moderate-to-minor effect in the adjoining areas of Kumaun. The 1991 Uttarkashi earthquake occurred over a low-angle reverse fault southern of the Vaikrita Thrust (MCT–II). Nearly to 700 died, and a considerable effect was inflicted on the property. Monitored through a series of aftershocks and micro-seismicity (1979 to 1980 and 1984 to 1986), Show the focus at Bhatwari, to the north, at 10-km depth. The epicenter, below the MCT–II, is at the same level [65,66]. Based on aftershocks distribution, the Uttarkashi Fault, a bedrock-mapped fault, was interpreted as the causative fault [67]. The 1999 Chamoli earthquake, with a magnitude (of 6.4), devastated the Chamoli and neighboring villages, inflicting about 500 casualties. The Chamoli earthquake indicates reverse faulting with plane dipping NNE at a low, 15-20° angle. The epicenter lies within the regional micro-seismicity belt located south of the Vaikrita Thrust in the Main Central Thrust (MCT) zone. Displacements may be controlled by tectonically enhanced erosion balanced by displacement within the hanging wall of the Vaikrita Thrust/MCT [68]. The aftershocks recorded by the WIHG network lie below the MCT at 7-17 km depth over the inferred ramp region [66]. The earthquake catalog suggests that Mw 6.3 (Uttarkashi, 1991) and Mw 6.6 (Chamoli, 1999) are the highest magnitude earthquakes in this part during the last three decades [69].

3. Methodology

The study has been completed based on the following data sets:

  1. Survey of India (SOI) Toposheets (53N/7, 53N/8, 53N/11, 53N/12 and 53N/15) of 1:50,000 scale of the years 1964 and 1967.
  2. SRTM Digital Elevation Model (DEM) data (http://srtm.csi.cgiar.org/srtmdata), along with GIS software, is used to analyze and extraction of morphotectonics and drainage networks.

The basin area has been manually outlined from (SOI) toposheets, and other morphometric metric calculations and maps have been prepared in Arc Map 10 GIS software. The methodology adopted for computing and numerical formulas of Morphometric indices including Drainage network, Bifurcation ratio (Rb), Circularity ratio (Rc), Hypsometric Integral (Hi), Basin elongation ratio (Re); Drainage basin symmetry such as Asymmetric factor (Af), Drainage basin shape (Bs), Transverse topographic symmetry index (TTSF), Valley floor width to valley height index (Vf), Sinuosity Index (Si); Stream length-gradient Index (SLGI) included SLGsegment, SLItotal and SLGanomalous points; Tectono-geomorphologic: Stream longitudinal profile (Ls) and Transverse profile (Tp) are used to interpret the tectonic relationship of the basin (Figure 3). The drainage network was analyzed following the [70] and stream orders following the [71] method.

Click to view original image

Figure 3 Geomorphic indices applied in this study modified after [34,36,72,73,74,75,76] are (a) Hypsometric curve. (b) HI: hypsometric integral. (c) BS: drainage basin shape. (d) AF: asymmetry factor. (e) SLGI: stream length-gradient. (f) VF: valley floor width-to-height ratio. (g) Sinuosity Index (SI). (h) Transverse topographic symmetry factor (TTSF). (i) Basin elongation Ratio (Re).

3.1 Bifurcation Ratio (Bi)

Normally bifurcation ratio ranges from 3.0 to 5.0 for basins in which the geological features (e.g., active faults) do not alter the drainage pattern where climate, subsurface rock type, and topography are more or less homogeneous [71]. The lower values of ‘Rb’ reflect less structural disruptions and less disturbed drainage, whereas the higher values imply high structural complexity and low permeability of the terrain. It is calculated by the formula

\[ R_{b}=\frac{N_{u}}{N_{u}+1} \tag{1} \]

Nu is the total number of stream order 'u', and ‘Nu + 1’ are the Number of segments of the higher-order.

3.2 Circularity Ratio (Rc)

It is calculated by the formula

\[ R_{c}=\frac{4 \pi A}{P^{2}} \tag{2} \]

where A and P is an area and perimeter of the basin respectively.

3.3 Hypsometric Integral (HI)

High HI values indicate that most topography is high relative to the mean representing a youthful topographic stage. Low HI values represent a mature stage of development [77] (Figure 3a, 3b). A hypsometric basin Integral shows the correlation between the basin area and the altitudes of the basin. Hi is calculated by the following formula

\[ \mathrm{Hi}=\frac{\left(\mathrm{H}_{\text {mean }}-\mathrm{H}_{\min }\right)}{\left(\mathrm{H}_{max }-\mathrm{H}_{\min }\right)} \tag{3} \]

where Hmean = mean elevation value, Hmax. = maximum elevation value and Hmin. = minimum elevation.

3.4 Basin Elongation Ratio (Re)

It is used for determining the shapes of the drainage basin. Tectonic effects influence the elongated shape of the basin, and it reflects on a drainage basin [36].

\[ \operatorname{Re}=\left(\frac{2}{\mathrm{Lb}}\right)\left(\frac{\mathrm{A}}{\pi}\right)^{0.5} \tag{4} \]

where Re = Elongation Ratio, A = Area of the Basin (km2), Lb = Basin length.

3.5 Drainage Basin Shape (Bs)

The Bs index describing a drainage basin elongation ratio is defined as

\[ \mathrm{Bs}=\frac{\mathrm{Bl}}{\mathrm{Bw}} \tag{5} \]

where Bs = drainage basin shape ratio, Bl = length of the basin, and Bw = width of the basin.

3.6 Asymmetry Factor (Af)

The asymmetry factor (Af) proposed by [78] is the ratio of area to the right of the basin Ar (facing downstream) to the total area of the basin. At normalized by multiplying the ratio with 100. The asymmetry factor is sensitive to tilting perpendicular to the trunk of the stream [78]. The asymmetry factor is useful in determining relative tectonic upliftment for basin areas.

The asymmetric factor (Af) is defined as

\[ \mathrm{Af}=100\left(\frac{\mathrm{Ar}}{\mathrm{At}}\right) \tag{6} \]

The ‘Af’ value (>50) indicates a tilting to the left bank, and the value (<50) shows a significant tilting of the drainage basin to the right side of the channel either because of active tectonics or lithological influence [72].

3.7 Stream Length-Gradient (SLGI)

It is one of the indexes suggested by the following formula for tectonic force assessment.

\[ \mathrm{SLG}_{\text {segment }}=\left(\frac{\Delta \mathrm{H}}{\Delta \mathrm{L}} \times \mathrm{L}\right) \tag{7} \]

where ‘SLG’ = Stream length gradient, ‘∆H/∆L’ is the Stream slope or gradient of the basin, ‘ΔH’ = height difference between two contours in the drainage basin, ‘ΔL’ = length of the stream in between these two contours and ‘L’ = total length of the channel from the midpoint of the segment to the head of the channel.

\[ \mathrm{SLI}_{\text {total }}=\frac{\Delta \mathrm{H}_{\text {total }}}{\text { in } \mathrm{L}_{\text {total }}} \tag{8} \]

where ‘ΔH’ is a difference in heights between the headwaters and the mouth and ‘L’ is a total length of the channel from the midpoint of the segment to the head of the channel.

\[ \text { SLGI }_{\text {anomalous points }}=\frac{\text { SLG segment }}{\text { SLItotal }} \tag{9} \]

where ‘L’ is a total length of the channel from the midpoint of the Selected extend to the channel's highest point.

3.8 Valley Floor Width-To-Height Ratio (Vf)

In estimating the basin floor width and valley height, we analyzed ‘Vf’ in different sections. Those basins, which have been the proximity to the structures but have U-shaped valleys, have been influenced by the erosion of the main river channel. The low values are confined to regions with less erosion, deep, narrow valleys, and V-shaped, Active Incision. The Valley floor width to valley height ratio is defined as

\[ \mathrm{Vf}=\frac{2 \mathrm{Vfw}}{((\mathrm{Eld}-\mathrm{Esc})+(\mathrm{Erd}-\text { Esc }))} \tag{10} \]

3.9 Sinuosity Index (SI)

The sinuosity index has been defined, and all sections for investigating the channel Sinuosity parameters of the river were measured using the procedure [74]. Based on the channel sinuosity index [74,79] proposed a classification of channel pattern as straight (<1), Low sinuous (1-1.3), High sinuous (1.3-1.5), and Meandering (>1.5) (Table 2). The overall length of the channel and the main channel length are computed using the given equation for each reach and are defined as

\[ \rm{S}_\rm{i}=\frac{\rm{C}_{l}}{\rm{V}_{l}} \tag{11} \]

Table 2 Summary of Morphotectonic parameters in the study area.

The Nandakini River is estimated to be 57 km long, and the river segments were demarcated into 30 sections into a series of equal-length reaches. Each have a length of 1.9 km.

3.10 Transverse Topographic Symmetry Factor (TTSF)

The topographical index or transverse symmetry variable was determined as

\[ \text { TTSF }=\frac{\text { Da }}{\text { Dd }} \tag{12} \]

‘Da’ is the space from the drainage basin's midline to the active belt's midline, and ‘Dd’ is the space from the midline to the basin limit [72].

4. Results and Discussion

4.1 Morphometric Indices

The qualitative morphometric analyses were carried out on the drainage network, the spatial distribution of the drainage basin, and their geometric relationships on topographic profiles by using SRTM DEM, following Strahler’s stream ordering system [71]. Evaluation and derivation of morphometric parameters measured and delineated in the Arc-GIS platform [82]. The basic characteristic of stream networks is the stream segment, and the assignment of stream orders is the first stage in the geomorphic analysis [83]. The drainage basin of the Nandakini River (~551 km2) was classified into fourth order by using 117 streams. The majority of streams (91) were found to be of the 1st order, occupying about 77.78% of the whole watershed, whereas, 2nd, 3rd, and 4th order streams occupy 17.1%, 4.3%, and 0.9%, respectively. The drainage is predominantly dendritic (Table 3, Figure 4).

Table 3 Morphometric parameters as measured in Nandakini Basin. U = Stream Order, Nu = Stream Number, Lu = Stream Length, Lsm = Mean Stream Length, Rl = Stream length Ratio, Rb = Bifurcation Ratio.

Click to view original image

Figure 4 (a) Drainage map of the Basin, (b) Bivariate plots of Regression of logarithm of number of streams versus stream order and (c) Regression of logarithm of stream lengths versus stream order [70].

4.1.1 Bifurcation Ratio (Rb)

In Nandakini Basin, the 1st, 2nd, 3rd, and 4th order are 4.55, 4 and 5 (Table 2, Table 3). Which shows the basin is lithologically and structurally controlled.

4.1.2 Circularity Ratio (Rc)

The circularity ratio is defined as the ratio of the basin area and the circle area with the perimeter as of the basin area [71,77]. [77] has characterized the basin of the circularity ratios that range from 0.40 to 0.50, which shows sub-elongated and highly porous homogenous geologic materials [84,85,86]. The basin's circularity ratio value (0.38), which subbed elongated and in mature stage topography, supported a dendritic pattern of the drainage network, high discharge of runoff, and low permeability condition (Table 2).

4.1.3 Hypsometric Integral (Hi)

[87] proposed a dimensionless parameter hypsometric integral as a tool for measuring landscape evolution. Hypsometric analysis (or area-altitude analysis) studies the distribution of horizontal cross-sectional area of a landmass concerning elevation. Hypsometric analysis has been used to interpret the relative amount of deformation or degradation due to hydrological and erosional processes that have taken place in a watershed during an appreciable part of the geological period. Hypsometric integral proves beneficial in calculating and comparing various sub-basins in a watershed having different area elevation values irrespective of scale. It is related to drainage basins' geomorphic and tectonic development in terms of their structures and processes [35,71,87]. The hypsometric contour curve 6886, 3484, 2800, 2360, 1937, 852, and the Hi value for this basin is calculated at 1.07. This indicates a deeper incision and a slight erosion (Table 2, Table 4).

Table 4 Hypsometric Integrals [80].

Which indicate the Deep incision in the region [35] as (Figure 5).

Click to view original image

Figure 5 Hypsometric integral of the Basin [35,56].

4.1.4 Basin Elongation Ratio (Re)

The ‘Re’ for the Nandakini river basin is 0.17, referring to the tectonic disturbance in the basin area (Table 2).

4.2 Drainage Basin Symmetry

4.2.1 Asymmetry Factor (Af)

Calculating the asymmetry factor for tectonic frequency determination of discharge of the stream to the major watercourse [72]. This index also defines the directions of neo-tectonic. The asymmetry factor explains the tectonic tendency of the channel discharge to the significant watercourse activity and is responsible for elevating and substituting separate blocks versus broad inclination [78]. The asymmetry factor is the higher level of neo-tectonic action tilts [88] (Table 2), resulting in basins (Figure 6).

Click to view original image

Figure 6 Block diagram of basin drainage asymmetry factor [72].

4.2.2 Drainage Basin Shape (Bs)

Basin shape is an important parameter because it influences and controls the geometry of the stream network [89]. It influences tectonic processes. Extended basins frequently define tectonically active areas and sub-circular forms without structural deformation or deformation level [36]. The Drainage basin shape for the basin is 3.27 (Table 2), which indicates the basin is tectonically active. The shape of a drainage basin is controlled by tectonic processes [90].

4.2.3 Transverse Topographic Symmetry Factor (TTSF)

It refers to Trunk River deviation from the basin mid-line. Its values range from (0-1), with a value similar to '0' indicating symmetric basin or ranges and a value greater than or less than '0' suggesting asymmetric basin or tilting [8,72,79]. The values of ‘TTSF’ for all the drainage basins have been calculated at different segments of the basins (Figure 7a). The Average value of TTSF of the drainage basins is 0.31, indicating asymmetric basins influenced by neotectonics (Table 2, Table 5).

Table 5 Segments selected for computation of Transverse Topographic Symmetry Factor (TTSF).

Click to view original image

Figure 7 (a) The Transverse Topographic Symmetry Factor (TTSF) is calculated by T = Da/Dd, where Da = distance from midline of the drainage basin, Dd = distance from basin midline to the basin divide and (b) Valley floor to valley width ratio (Vf).

4.2.4 Valley Floor Width to Valley Height Ratio (Vf)

The Calculating ‘Vf’ index in this analysis, by ten cross-sections on the mainstream and sub-streams in the highlands of 467.72, 570.28, 1032.02, 780.58, 468.35, 627.69, 1120.96, 1027.45, 1162.36 and 652.05 that show the valley floor heights, they are drawn perpendicular to the stream. The necessary calculations have been performed on them. The average ‘Vf’ of the basin is 0.54 (Table 2, Table 6), suggesting low values indicative of less erosion, deep, narrow valleys, and V-shaped (Figure 7b, Figure 8).

Click to view original image

Figure 8 The cross-section for Valley floor width to valley height ratio (Vf).

Table 6 Segments selected for valley floor to valley width ratio (Vf).

4.2.5 Sinuosity Index (Si)

Channel sinuosity is a key parameter in understanding the role of active tectonics. Along with active faults where erosion rates are high due to the pulverized rock material, the river erodes its valley linearly, forming strath terraces. Channel sinuosity values are lower in regions with relatively high tectonic effects. The range of the Si value varies from 0.99 to 1.50. The average sinuosity of 1.12 is called low sinuosity (Figure 9). The channel sinuosity is controlled by channel gradient, sediment load, resistance to lateral erosion of a river, stage of valley development, and the structural characteristics of the area through which it flows [90,91,92,93].

Click to view original image

Figure 9 Calculated the Sinuosity Index of the basin which is divided into straight, Low and Meandering zones.

In this study, three zones of the Sinuosity Index have been observed based on [74], namely as straight, low sinuous zone, and high sinuous zone. The measured, observed sinuosity and linear fit curve for channel following straight (Si-0.99) at 17 km. The low sinuosity zone (Si-1.02 to 1.25) referred to low sinuosity [94]. The sinuosity of the river is 1.25 which also shows the braided nature [95]. Mostly the river has low sinuosity. Sinuosity increases when the outer bend of the bank erodes more rapidly than the inner bend [96]. The alluvial theory of meanders relates this lateral erosion imbalance to hydraulic valves [80,97,98].

Similarly, [99] revealed that remarkable sinuosity changes relate to changes in river discharge and sediment load at the tributary interflow. Even the smallest topographical changes influence the sinuosity of low gradient rivers [100]. The meandering zone (Si-1.50) shows the meandering zone at 40 km from the source. Most meandering streams show substantial changes in sinuosity over time, reflecting not only the geomorphic history but also active tectonics. Stream variations and other influences on valley slopes [42,101]. The straight and low sinuosity zone shows the high peek zone (V) shape valley. Mostly the river has low sinuosity.

4.3 Stream Length Gradient Index (SLGI)

4.3.1 SLGsegment

[81] defined the tectonic activity and nature of the rocks. The SLGsegment for the basin is 1285.73, which is recommended for High tectonic activity (Table 2).

4.3.2 SLItotal

The stream length gradient ratio is studied to analyze tectonic influence, rock resistance, and topography. The roughly equivalent tectonic activity in the basin was assessed [34,102]. The SLItotal for the basin is 120.81.

4.3.3 SLGIanomalous points

The tectonic activity is relatively high, and vice versa [73,81]. Most of the SLGI anomalous points are situated in the locations where the change in slope gradient on the longitudinal profiles or where the rivers take right-angled turns. The SLGI for the Nandakini River basin is 10.64 (Table 2). A high value of the SLG represents the hard bedrock of stream and reflects.

4.4 Tectono-Geomorphologic Investigations

Based on the integrated application of satellite imagery and digital data, a take-out of reliable information on the geomorphologic evolution of the basin was carried out. We attempted to evaluate the effect of tectonic deformation along the Nandakini River valley, indicating the geologic complexity setting.

4.4.1 Longitudinal Profile (Ls)

A curvature that shows the relationship between heights versus distance downstream of a channel. Longitudinal channel profiles may indicate instability due to climatic, rock-type, and tectonic factors [103]. Hence, it is one of the geomorphic indexes used to identify areas with intense tectonic deformation [35,73]. The river is evident in the cyclic evolution of the basins due to local neotectonic control [104]. The longitudinal river profiles contain the uplift and tectonic records [72]. This profile starts at 4955.22 m height at the head, and after reaching the maximum, the profile further shows the decline and ends its journey at 965.23 m height at the mouth. The longitudinal profiles of the streams show a sudden change in slopes, graded profile, a considerable amount of incision, and deposition along their profiles, showing tectonically instability (Figure 10).

Click to view original image

Figure 10 (a) Longitudinal profile of the Nandakini River. (b) Illustrates a concave up profiles and distribution of the knickpoints. From knickpoints (N1–N4), it flows on steep slope in a highly faulted zone and thrusting, whereas (K5–K6) indicate the resistance of hard rock formation. In the final stage, the river shows stiff resistance rock.

4.4.2 Transverse Profile (Tp)

It is prepared using 90 SRTM DEM and section as NTP-1, NTP-2, NTP-3, NTP-4, NTP-5, NTP-6, NTP-7, NTP-8, NTP-9, and NTP-10. These segments are running parallel to each other along E-W in direction. These profiles are taken across the river within the basin. Each profile shows irregularity, tilting, and undulation. The Profile NTP-1, NTP-2, NTP-3, NTP-4, NTP-8, NTP-9 and NTP-10 tilting toward left valley side while Profile NTP-5, NTP-6, NTP-7 tilting right valley side (Figure 11). Its entire journey profile cuts the tributaries and geomorphic features at various locations. The transverse profiles of the river show a cross-section of the channel and valley of a river in the path of the river at some point. A river cross profile varies as it shifts from the head to the mouth due to changes in the energy of the river and the processes the river carryout.

Click to view original image

Figure 11 Transverse profile of the Nandakini River basin. Section within five profiles following NTP1 to NTP10. In which blue colour shows the right bank, yellow colour shown left bank, Pink colour showing the main channel. The surface profile showing irregularity and the linear trend of profile showing the tilting of the basin.

The valley and narrowing channel deep in the upper course because of the large amount of vertical and little lateral erosion. In the upper course, the sides of a river valley are very steep, term “V-shaped valley” for these valleys. Due to an increase in lateral erosion, the valley decreases in width in the middle section. Due to increasing vertical erosion, its depth does not change significantly. The land in the valley to the side of the channel is recognized as the river's floodplain. The mouth of the valley is narrow again due to a sudden decrease in slope. The upper course of the basin tilts toward the right side, while the lower course tilts toward the left side. This tilting of the profile shows the subsurface tectonic activity. It has been found that the valley increases in the middle course, and the valley may become narrow again in the lower course. Such transverse profiles were prepared to understand the structure of the valley and its width from the origin of the mouth and recent tectonic activities.

5. Conclusions

In this research, efforts have been made to discuss the tectonic-geomorphic signatures of the Nandakini basin, which is subject to neotectonic activity. Through qualitative and quantitative morphometric analyses with an emphasis on topography, I demonstrate the role of surface processes and topographic adjustment on landscape development in the study area. The morphometric parameters indicate a transient response of the landscape to disequilibrium initiated by active tectonics. The circularity ratio reveals that the basin is sub-elongated, whereas the asymmetric basin factor indicates the effect of active tectonics or substantial lithological control in the study area. The Nandakini terrain points to the youthful topography. The hypsometric integral data suggest that the basin has experienced unevenness and respond to Deep incision and slight erosion due to the active tectonics in the area. Further, the presence of high-elevation surfaces to the tectonic disturbances in the area. Re, Bs, SLGI is 0.17, 3.27, and 10.64, respectively, reporting that the area is tectonically active. TTSF is 0.31 implies the asymmetric nature. The river profile showing topographical undulation, tilting, and disequilibrium condition suggests that the river is tectonically influenced. The geomorphic features include unpaired fluvial terraces with tilting, rugged topography and boulders size sediment, V-shaped valleys, active river terraces, river meandering, and lake depositions, landslides, and mudflow indicate active tectonism. From knickpoints N1 to N4, the river flows on the high slope in the highly faulted zone showing action by the tectonic activity. Abrupt break in slope, sinuous to meandering course of the river, youthful topography, and elongated shaped basin add to the evidence of neotectonic. The higher rate of rock weathering, particularly of schists in the vicinity of the active supplementary faults, disintegrated rocks and numerous landslides, is additionally responsible for the landscape revolution. The drainage basins are significantly elongated in shape, and sediment carrying capacity is high and tilted. Most river valleys are narrow and deep, indicating active incision due to tectonic activity.

Acknowledgments

I am very thankful to the Head of Department of Geology, University of Lucknow, Lucknow, for providing the working facilities and place to carry out this work. I would like to express my sincere thanks to my Ph.D. supervisor Prof. Dhruv Sen Singh, Department of Geology, University of Lucknow, Lucknow for his guidance and their scientific value, relentless encouragement and help in all possible ways. I gratefully acknowledged the University Grant Commission (UGC), Rajiv Gandhi National Fellowship, in the form of Senior Research Fellowship (RGNF-2017-18-SC-UTT-34263) in a Ph.D. program at the University of Lucknow, Lucknow, Uttar Pradesh, India and Anoop Kumar Singh to SERB NPDF (PDF/2020/000251) for financial support.

Author Contributions

Pawan Kumar Gautam wrote and edited the manuscript. He also prepared the necessary maps. Anoop Kumar Singh made corrections to the manuscript.

Competing Interests

The authors have declared that no competing interests exist.

References

  1. Singh DS, Dubey CA, Kumar D, Vishawakarma B, Singh AK, Tripathi A, et al. Monsoon variability and major climatic events between 25 and 0.05 ka BP using sedimentary parameters in the Gangotri Glacier region, Garhwal Himalaya, India. Quat Int. 2019; 507: 148-155. [CrossRef]
  2. Auden JB. Structure of the Himalayas in Garhwal. Rec Geol Surv India. 1937; 71: 407-433.
  3. Chamyal LS. Stratigraphy of the Lesser Himalayan rocks in Kumaun. Proc Indian Acad Sci.. 1991; 100: 293-306. doi: 10.1007/BF02895989. [CrossRef]
  4. Ahmad A, Saxena A, Siddhanta BK. Structural analysis of eastern kumaun Himalaya and related problems. Heim Geol. 1980; 10: 280-303.
  5. Bhattacharya AR, Agarwal KK. Mylonites from the Kumaun Lesser Himalaya. Neues Jahrb fur Mineral Abh. 1985; 152: 65-77.
  6. Bhattacharya AR, Agarwal KK. Strain analysis and fold shape modification in a crystalline complex of the Lesser Himalaya. Krystallinikum. 1989; 20: 7-25.
  7. Bhattacharya AR. New light on the stratigraphy and structure of the Kumaun Lesser Himalaya, A study in mathematical perspective. In: Structural geology of the Himalaya. New Delhi: Today and Tomorrow’s publishers; 1979. pp. 151-462.
  8. Heim A, Gansser A. Central Himalaya: Geological observations of the Swiss expedition, 1936. Mem Soc Helv Sci Nat. 1939; 73: 1-245.
  9. Misra RC, Sharma RP. Structure of Almora Crystalline, Lesser Kumaun Himalaya. An interpretation. Him Geol. 1972; 2: 330-341.
  10. Saxena SP, Rao PN. Does Almora Nappe exists? Him Geo India. 1975; 5: 169-184.
  11. Singh AK, Kumar D, Dubey CA, Gautam PK, Vishawakarma B, Singh DS. Glacial Morpho-Sedimentology and processes of landscape evolution in Gangotri Glacier Area, Garhwal Himalaya, India. In: Climate Change and Geodynamics in Polar Regions. CRC Press; 2022. [CrossRef]
  12. Valdiya KS. An outlook of the structural setup of the Kumaun Himalaya. Geol Soc India. 1979; 20: 145-157.
  13. Valdiya KS. Geology of the Kumaun Lesser Himalaya. Dehradun: Wadia Institute of Himalayan Geology; 1980. pp. 290-291.
  14. Gansser A. Geology of the Himalaya. New York: Interscience; 1964.
  15. Nakata T. Geomorphic history and crustal movement of the foot-hills of the Himalaya. Sci Rep Tohoku Univ. 1972; 22: 39-177.
  16. Nossin JJ. Outline of the geomorphology of the Doon valley, northern UP, India. Zeitschrift für Geomorphol. 1971; 12: 18-50.
  17. Singh AK, Parkash B, Mohindra R, Thomas JV, Singhvi AK. Quaternary alluvial fan sedimentation in the Dehradun valley piggyback basin, NW Himalaya: Tectonic and palaeoclimatic implications. Basin Res. 2001; 13: 449-471. [CrossRef]
  18. Garzanti E, Baud A, Mascle G. Sedimentary record of the northward fight of India and its collision with Eurasia (Ladakh Himalaya, India). Geodinamica Acta. 1987; 1: 297-312. [CrossRef]
  19. Jade S, Mukul M, Gaur VK, Kumar K, Shrungeshwar TS, Satyal GS, et al. Contemporary deformation in the Kashmir–Himachal, Garhwal and Kumaon Himalaya: Significant insights from 1995-2008 GPS time series. J Geodesy. 2014; 88: 539-557. [CrossRef]
  20. Jade S, Shrungeshwara TS, Kumar K, Choudhury P, Dumka RK, Bhu H. India plate angular velocity and contemporary deformation rates from continuous GPS measurements from 1996 to 2015. Sci Rep. 2017; 7: 11439. [CrossRef]
  21. Patriat P, Achache J. India-Eurasia collision chronology has implications for crustal shortening and driving mechanism of plates. Nature. 1984; 311: 615-621. [CrossRef]
  22. Rowley DB. Age of initiation of collision between India and Asia: A review of stratigraphic data. Earth Planet Sci Lett. 1996; 145: 1-13. [CrossRef]
  23. Joshi LM, Kotlia BS. Tectonic footprints and landscape evaluation along kulur river valley, Kumaun lesser Himalaya, India. J Asian Earth Sci. 2018: 162; 121-136. [CrossRef]
  24. Srivastava P, Mitra G. Thrust geometries and deep structure of the outer and Lesser Himalaya, Kumoan and Garhwal (India): Implications for evolution of the Himalayan fold and thrust belt. Tectonics. 1994: 13; 89-109. [CrossRef]
  25. Brookfield ME. The Himalayan passive margin from Precambrian to Cretaceous times. Sediment Geol. 1993: 84; 1-35. [CrossRef]
  26. Yin A, Harrison TM. Geologic evolution of the Himalayan–Tibetan orogen. Annu Rev Earth Planet Sci. 2000; 23: 211-280. [CrossRef]
  27. Thakur VC. Geology of Western Himalaya. Oxford: Pergamon Press; 1992. pp. 363.
  28. Arjun K, Ashwani K, Gupta SC, Himanshu M, Arup S. Earthquake source parameters and their scaling for the Uttarakhand region of North-West Himalaya. Disaster Adv. 2014; 7: 28-41.
  29. Sarkar I. The role of the 1999 Chamoli earthquake in the formation of ground cracks. J Asian Earth Sci. 2004; 22: 529-538. doi: 10.1016/S1367-9120(03)00093-2. [CrossRef]
  30. Kumar A, Mittal H, Sachdeva R, Kumar A. Indian strong motion instrumentation network. Seismol Res Lett. 2012; 83: 59-66. [CrossRef]
  31. Monalisha N, Sitharam T, Sreevalsa K. A Revisit to Seismic Hazard at Uttarakhand. Int J Geotech Earthq Eng. 2015; 6: 56-73. doi: 10.4018/IJGEE.2015070104. [CrossRef]
  32. Kumar N, Dumka RK, Mohan K, Chopra S. Relative active tectonics evaluation using geomorphic and drainage indices, in Dadra and Nagar Haveli, western India. Geod Geodyn. 2022; 13: 219-229. [CrossRef]
  33. Mayer L. Tectonic geomorphology of escarpments and mountain fronts. In: Active tectonics. Washington: National Academic Press; 1986. pp. 125-135.
  34. Valdiya KS, Kotlia BS. Fluvial geomorphic evidence for Late Quaternary reactivation of a synclinally folded nappe in Kumaun Lesser Himalaya. J Geol Soc India. 2001; 58: 303-317.
  35. Keller EA, Pinter N. Active tectonics: Earthquake, uplift and landscape. 2nd ed. New Jersey: Prentice Hall; 1996.
  36. Bull WB, McFadden LD. Tectonic geomorphology north and south Garlock fault, California. In: Geomorphology in Arid Regions. Binghamton: State University of New York; 1977.
  37. Burbank DW, Anderson RS. Tectonic geomorphology. Oxford: Blackwell Publishing; 2001. p. 274.
  38. Romshoo SA, Bhat SA, Rashid I. Geoinformatics for assessing the morphometric control on hydrological response at watershed scale in the Upper Indus Basin. J Earth Syst Sci. 2012; 121: 659-686. [CrossRef]
  39. Singh V, Tandon SK. The Pinjaur dun (intermontane longitudinal valley) and associated active mountain fronts, NW Himalaya: Tectonic geomorphology and morphotectonic evolution. Geomorphology. 2008; 102: 376-394. [CrossRef]
  40. Hajam RA, Hamid A, Bhat S. Application of morphometric analysis for geo-hydrological studies using geo-spatial technology – a case study of Vishav Drainage Basin. Hydrol Curr Res. 2013; 4: 157.
  41. Waikar ML, Nilawar AP. Morphometric analysis of a drainage basin using geographical information system: A case study. Int J Multidiscip Curr Res. 2014; 2: 179-184.
  42. Wyzga B. River response to channel regulation: Case study of the Raba River, Carpathians, Poland. Earth Surf Process Landf. 1993; 18: 541-556. [CrossRef]
  43. Burbank DW, Leland J, Fielding E, Anderson RS, Brozovic N, Reid MR, et al. Bedrock incision, rock uplift and threshold hillslopes in the northwestern Himalayas. Nature. 1996; 379: 505-510. [CrossRef]
  44. Dietrich WE, Wilson CJ, Montgomery DR, McKean J. Analysis of erosion thresholds, channel networks, and landscape morphology using a digital terrain model. J Geol. 1993; 101: 259-278. [CrossRef]
  45. Moglen GE, Bras RL. The importance of spatially heterogeneous erosivity and the cumulative area distribution within a basin evolution model. Geomorphology. 1995; 12: 173-185. [CrossRef]
  46. Montgomery DR, Foufoula-Georgiou E. Channel network source representation using digital elevation models. Water Resour Res. 1993; 29: 3925-3934. [CrossRef]
  47. Sklar L, Dietrich WE. River longitudinal profiles and bedrock incision models: Stream power and the influence of sediment supply. Geophys Monogr. 1998; 107: 237-260. [CrossRef]
  48. Snyder NP, Whipple KX, Tucker GE, Merritts DJ. Landscape response to tectonic forcing: Digital elevation model analysis of stream profiles in the Mendocino triple junction region, northern California. Geol Soc Am Bull. 2000; 112: 1250-1263. [CrossRef]
  49. Tarboton DG, Bras RL, Rodriguez-Iturbe I. On the extraction of channel networks from digital elevation data. Hydrol Process. 1991; 5: 81-100. [CrossRef]
  50. Tucker GE. Modeling the large-scale interaction of climate, tectonics and topography. State College, PA: The Pennsylvania State University; 1996.
  51. Weissel JK, Seidl MA. Inland propagation of erosional escarpments and river profile evolution across the southeast Australian passive continental margin. Geophys Monogr. 1998; 107: 189-206. [CrossRef]
  52. Joshi LM, Pant PD, Kotlia BS, Kothyari GC, Luirei K, Singh AK. Structural overview and morphotectonic evolution of a strike-slip fault in the zone of North Almora Thrust, Central Kumaun Himalaya, India. J Geol Res. 2016; 2016: 6980943. [CrossRef]
  53. Luirei K, Bhakuni SS, Kothyari GC, Tripathi K, Pant PD. Quaternary extensional and compressional tectonics revealed from Quaternary landforms along Kosi River valley, outer Kumaun Lesser Himalaya, Uttarakhand. Int J Earth Sci. 2016; 105: 965-981. [CrossRef]
  54. Luirei K, Bhakuni SS, Srivastava P, Suresh N. Late Pleistocene–Holocene tectonic activities in the frontal part of NE Himalaya between Siang and Dibang river valleys, Arunachal Pradesh, India. Zeitschrift für Geomorphol. 2012; 56: 477-493. [CrossRef]
  55. Pant PD, Chauhan R, Bhakuni SS. Development of transverse fault along North Almora Thrust, Kumaun Lesser Himalaya, India: A study based on field and magnetic fabrics. J Geol Soc India. 2012; 79: 429-448. [CrossRef]
  56. Pant PD, Kothyari GC, Luirei K. Geomorphological and geological investigations of neotectonics activity of Saryu River Faults (SRF), a part of North Almora Thrust (NAT) in Seraghat-Basoli area in Central Kumaun, Uttaranchal. J Geol Soc India. 2007; 70: 815-823.
  57. Valdiya KS. Himalayan transverse faults and folds and their parallelism with subsurface structures of North Indian plains. Tectonophysics. 1976; 32: 353-386. [CrossRef]
  58. Valdiya KS. Reactivation of terrain-defining boundary thrusts in central sector of the Himalaya: Implication. Curr Sci. 2001; 81: 1418-1431.
  59. Yadav R, Tripathi SK, Pranuthi G, Dubey SK. Trend analysis by Mann-Kendall test for precipitation and temperature for thirteen districts of Uttarakhand. J Agrometeorol. 2014; 16: 164-171. [CrossRef]
  60. Célérier J, Harrison TM, Webb AAG, Yin A. The Kumaun and Garwhal Lesser Himalaya, India: Part 1. Structure and stratigraphy. Geol Soc Am Bull. 2009; 121: 1262-1280. [CrossRef]
  61. Srivastava RN, Ahmed A. Geology and structure of Alaknanda valley. Himal Geol. 1979; 9: 225-254.
  62. Dumka RK, Kotlia BS, Kumar K, Satyal GS. Quantification of crustal strain rate in Kumaun Himalaya (India) using GPS measurements of crustal deformation. Himal Geol. 2014; 35: 146-155.
  63. Paul A, Prasath AR, Singh R. Slip heterogeneities evaluated for earthquakes M>4.0 using waveform modelling in the Garhwal region of Central Seismic Gap in Northwest Himalaya, India. Him Geol. 2015; 36: 153-160.
  64. Dumka RK, Kotlia BS, Kumar K, Satyal GS, Joshi LM. Crustal deformation revealed by GPS in Kumaun Himalaya, India. J Mt Sci. 2014; 11: 41-50. [CrossRef]
  65. Khattri KN, Chander R, Gaur VK, Sarkar I, Kumar S. New seismological results on the tectonics of the Garhwal Himalaya. Proc Indian Acad Sci Earth Planet Sci. 1989; 98: 91-109. [CrossRef]
  66. Thakur VC, Kumar S. Seismotectonics of the 20 October 1991 Uttarkashi earthquake in Garhwal Himalaya, North India. Terra Nova. 1994; 6: 90-94. [CrossRef]
  67. Kayal JR. Earthquake source processes in northeast India: A review. J Himal Geol. 1996; 17: 53-69.
  68. Kothyari GC, Joshi N, Taloor AK, Malik K, Dumka R, Sati SP, et al. Reconstruction of active surface deformation in the Rishi Ganga basin, Central Himalaya using PSInSAR: A feedback towards understanding the 7th February 2021 flash flood. Adv Space Res. 2022; 69: 1894-1914. [CrossRef]
  69. Dumka RK, Kotlia BS, Kothyari GC, Paikrey J, Dimri S. Detection of high and moderate crustal strain zones in Uttarakhand Himalaya, India. Acta Geod et Geophys. 2018; 53: 503-521. [CrossRef]
  70. Horton RE. Erosional developments of streams and their drainage basin: Hydrophysical approach to quantitative morphology. Geol Soc Am Bull. 1945; 56: 275-370. [CrossRef]
  71. Strahler AN. Quantitative geomorphology of drainage basin and channel networks. In: Handbook of applied hydrology. New York: Mc Graw Hill Book Company; 1964.
  72. Cox RT. Analysis of drainage basin symmetry as a rapid technique to identify areas of possible quaternary tilt block tectonics: An example from the Mississippi Embayment. Geol Soc Am Bull. 1994; 106: 571-581. [CrossRef]
  73. Pinter N, Keller EA. Active tectonics: Earthquakes, uplift, and landscape. New Jersey: Prentice Hall; 2002.
  74. Leopold LB, Wolman MG, Miller JP. Fluvial processes in geomorphology. San Francisco: WH Freeman and Company; 1964.
  75. Özkaymak Ç. Tectonic analysis of the Honaz Fault (western Anatolia) using geomorphic indices and the regional implications. Geodin Acta. 2015; 27: 110-129. [CrossRef]
  76. Özsayın E. Relative tectonic activity assessment of the Çameli Basin, Western Anatolia, using geomorphic indices. Geodin Acta. 2016; 28: 241-253. [CrossRef]
  77. Miller VC. A quantitative geomorphic study of drainage basin characteristic in the clinch, Mountain area, Virginia and Tennessee. New York: Columbia University, Department of Geology; 1953.
  78. Pinter N. Applications of tectonic geomorphology for deciphering active deformation in the Pannonian Basin, Hungary. Proceedings of the workshop on “Applications of GPS in Plate Tectonics in Research on Fossil Energy Resources and in Earthquake Hazard Assessment”, occasional papers of the Geological Institute of Hungary. 2005; 204: 25-51.
  79. Hare PW, Gardner TW. Geomorphic indicators of vertical neotectonism along converging plate margins, Nicoya Peninsula, Costa Rica. Tecton Geomorphol. 1985; 4: 75-104.
  80. Strahler AN. Hypsometric (area-altitude) analysis of erosional topography. Geol Soc Am Bull. 1952; 63: 1117-1142 [CrossRef]
  81. Hack JT. Stream–profile analysis and stream–gradient index. J Res US Geol Surv. 1973; 1: 421-429.
  82. Gautam PK, Singh DS. Flood hazard and risk assessment of Deoha River Basin, Central Ganga Plain, India: An GIS approach. Disaster Adv. 2022; 15: 42-51. [CrossRef]
  83. Gautam PK, Singh DS, Singh AK, Kumar D. Geomorphic analysis of Baghain River, Yamuna Basin, and its implication for drainage characteristic and tectonics using remote sensing and GIS techniques. J Geol Soc India. 2022; 98: 1573-1584. [CrossRef]
  84. Pareta K, Pareta U. Quantitative morphometric analysis of a watershed of Yamuna basin, India using ASTER (DEM) data and GIS. Int J Geomat Geosci. 2011; 2: 248-269.
  85. Rai PK, Mohan K, Mishra S, Ahmad A, Mishra VN. A GIS-based approach in drainage morphometric analysis of Kanhar River Basin, India. Appl Water Sci. 2017; 7, 217-232. [CrossRef]
  86. Singh P, Gupta A, Singh M. Hydrological inferences from watershed analysis for water resource management using remote sensing and GIS techniques. Egypt J Remote Sens Space Sci. 2014; 17: 111-121. [CrossRef]
  87. Ikeda S, Parker G, Sawai K. Bend theory of river meanders. Part 1. Linear development. J Fluid Mech. 1981; 112: 363-377. [CrossRef]
  88. Mittal H, Kumar A, Ramhmachhuani R. Indian national strong motion instrumentation network and site characterization of its stations. Int J Geosci. 2012; 3: 1151-1167. [CrossRef]
  89. Kumar D, Singh DS, Prajapati SK, Khan I, Gautam PK, Vishawakarma B. Morphometric parameters and neotectonics of Kalyani River Basin, Ganga Plain: A remote sensing and GIS approach. J Geol Soc India. 2018; 91: 679-686. [CrossRef]
  90. Gautam PK, Singh DS, Kumar D, Singh AK. A GIS-based approach in drainage morphometric analysis of Sai River Basin, Uttar Pradesh, India. J Geol Soc India. 2020; 95: 366-376. [CrossRef]
  91. Gautam PK. Application of Geo-Spatial technique in river shifting analysis of the Ghaghara River: Case study from Bahraich to Ballia, Uttar Pradesh, India. 2019; 8: 362-368. [CrossRef]
  92. Gautam PK. Spatio-temporal analysis of sinuosity of Ghaghara River: A remote sensing and GIS approach. Int J Sci Res. 2019; 8: 1361-1364. [CrossRef]
  93. Prasad N. Some aspects of meandering streams of the Barakar basin and their sinuosity indexes. Perspect Geomorphol. 1982; 4: 93-102.
  94. Singh DS, Tiwari AK, Gautam PK. The Burhi Gandak: Most sinuous river. In: The Indian Rivers. Singapore: Springer; 2018. pp. 209-219. [CrossRef]
  95. Singh DS, Awasthi A. Natural hazards in the Ghaghara River area, Ganga Plain, India. Natural Hazard. 2011; 57: 213-225. [CrossRef]
  96. Kumar GP, Sen SD, Dhirendra K, Kumar SA, Balkrishan V. Assessment of morphometric characteristics and decadal fluctuation of groundwater level of Kali River basin and its link with climate change, West Uttar Pradesh, India Using Geographical Information System. Disaster Adv. 2022; 15: 26-38. [CrossRef]
  97. Edwards BF, Smith DH. River meandering dynamics. Phys Rev E. 2002; 65: 046303. [CrossRef]
  98. Einstein A. Die Ursache der Mäanderbildung der Flußläufe und des sogenannten Baerschen Gesetzes. Naturwissenschaften. 1926; 11: 223-224. [CrossRef]
  99. Timár G. Controls on channel sinuosity changes: A case study of the Tisza River, the Great Hungarian Plain. Quat Sci Rev. 2003; 22: 2199-2207. [CrossRef]
  100. Holbrook J, Schumm SA. Geomorphic and sedimentary response of rivers to tectonic deformation: A brief review and critique of a tool for recognizing subtle epirogenic deformation in modern and ancient settings. Tectonophysics. 1999; 305: 287-306. [CrossRef]
  101. Burnett AW, Schumm SA. Alluvial river response to neotectonic deformation in Louisiana and Mississippi. Science. 1983; 222: 49-50. [CrossRef]
  102. Dar RA, Romshoo SA, Chandra R, Ahmad I. Tectono-geomorphic study of the Karewa Basin of Kashmir Valley. J Asian Earth Sci. 2014; 92: 143-156. [CrossRef]
  103. Hack JT. Studies of longitudinal stream profiles in Virginia and Maryland. US Government Printing Office; 1957. pp. 45-97. [CrossRef]
  104. Sen PK. Geomorphological analysis of drainage basins. West Bengal, India: The University of Burdwan; 1993. pp. 132-133.
Newsletter
Download PDF Download Citation
0 0

TOP