You are on page 1of 19

Journal of Hydrology 612 (2022) 128172

Contents lists available at ScienceDirect

Journal of Hydrology
journal homepage: www.elsevier.com/locate/jhydrol

Research papers

Tracing thermal and non-thermal water circulations in shear zones of


Eastern Ghats Mobile Belt zone, Eastern India- inferences on sustainability
of geothermal resources
Tirumalesh Keesari a, b, *, Sitangshu Chatterjee a, b, Mukund Kumar c, Hemant Mohokar a,
Uday Kumar Sinha a, Annadasankar Roy a, b, Diksha Pant a, b, Suraj D. Patbhaje d
a
Isotope Hydrology Section, Isotope and Radiation Application Division, Bhabha Atomic Research Centre, Trombay, Mumbai, India
b
Homi Bhabha National Institute, Bhabha Atomic Research Centre, Mumbai, India
c
Geological Survey of India, Nagpur, India
d
Geological Survey of India, Jaipur, India

A R T I C L E I N F O A B S T R A C T

Keywords: The development of the geothermal resources in India has not progressed mainly due to inadequate under­
Thermal waters standing of the deeper thermal regime and uncertainty over the sustainability of heat source. In this study,
Mahanadi graben isotopes, hydrochemical techniques and integrated geothermometric models have been applied to a low-medium
Eastern Ghats Mobile Belt
enthalpy geothermal system belonging to shear zones of Mahanadi graben, Eastern Ghats Mobile Belt (Eastern
Geothermal energy
Environmental isotopes
India) to gain better insights. Isotope results indicate that thermal waters are meteoric in nature and absence of
Radiocarbon dating δ18O-shift suggests that the subsurface temperatures are below 200 ◦ C. Distinct isotopic and hydrochemical
signatures of thermal waters clearly suggest their origin from different geothermal reservoirs. Thermal waters at
Atri/Tarabalo sites are derived from the southern part of the graben while at Athmalik, from northern part. Away
from graben in southern direction (Taptapani site), the thermal waters are derived from rain occurring at
elevated regions. Seasonal variations in chemical and isotope data are minimum in thermal waters. The esti­
mated reservoir temperatures range between 92 and 124 ◦ C with reservoir depths 1.6 – 2.4 km. Tritium data
suggests that thermal waters are old and have 12–63% mixing with non-thermal water. The modelled radio­
carbon age of thermal waters (13.3 – 20.7 Ka) suggests paleo-recharge of the reservoirs during Last Glacial
Maximum (LGM), characterized by arid climate and low rainfall. A conceptual isotope model illustrating the
genesis and circulation patterns of thermal and non-thermal waters is presented. The study highlights the po­
tential of isotope-based studies in expanding the geothermal energy utilization in India.

1. Introduction researchers and the thermal models of the present-day EGMB crust
estimated a Moho temperature of ~550 ◦ C (Dasgupta, 1993; Bhowmik
The Proterozoic Eastern Ghats Mobile Belt (EGMB) bordering the et al., 1995). The concentrations of radioactive elements and the heat
Archaean Dharwar, Bastar and Singhbhum cratonic blocks is an production are reported to be highest amongst the granulite belts of
important Precambrian geological entity of Peninsular India known as a India (Kumar et al., 2007) and therefore EGMB is classified under Zone-
blueprint for models of reconstruction of Columbia, Rodinia and II based on the heat flow values (Fig. 1b). Several surface manifestations,
Gondwana (Dasgupta et al., 2013; Valdiya, 2016) (Fig. 1a). EGMB is mainly low to medium enthalpy thermal waters with temperatures
considered to be a collage of several juxtaposed terrains separated by ranging from 42 to 59 ◦ C feature in EGMB (GSI, 1987; GSI, 1991; Thussu,
major shear zones and is characterized by stretching lineation and in­ 2002). Considering the heat flow values, many researchers have sug­
ternal homogeneity in terms of litho-package and tectonic history gested for exploration of thermal waters for electricity production using
(Chetty, 2001; Valdiya, 2016). An ultra-high temperature meta­ binary cycle power plants or through enhanced geothermal systems
morphism at mid to lower crustal depths has been reported by many (Chandrasekharam and Chandrasekhar, 2010).

* Corresponding author at: Isotope Hydrology Section, Isotope and Radiation Application Division, Bhabha Atomic Research Centre, Trombay, Mumbai, India.
E-mail addresses: tiruh2o@gmail.com, tirumal@barc.gov.in (T. Keesari).

https://doi.org/10.1016/j.jhydrol.2022.128172
Received 11 April 2022; Received in revised form 30 June 2022; Accepted 2 July 2022
Available online 11 July 2022
0022-1694/© 2022 Elsevier B.V. All rights reserved.
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

Studies were conducted on the thermal waters of India covering all isotopes of water (2H, 18O & 3H) and its dissolved species (13C, 14C, 15N
the ten geothermal provinces, which mainly reported the thermo- and 34S etc.) were also applied in some cases to gain better insights into
tectonic aspects, quality of thermal water, reservoir temperatures and genesis, mixing processes, geochemical evolution and residence time
geochemical processes (GSI, 1987; GSI, 1991; Minissale et al., 2000; distribution of thermal fluids (Sammel and Craig, 1981; Navada and
Thussu, 2002; Craig et al., 2013; Zimik et al., 2017). Environmental Rao, 1991; Chatterjee et al., 2016, 2017; Keesari et al., 2019). As per the

Fig. 1. a) Major Palaeoproterozoic sutures in Peninsular India classified as the Eastern India, Central India and Western India Suture zones. The Southern Granulite
Terrain was mostly assembled during late Neoproterozoic–Cambrian during the final assembly of the Gondwana supercontinent (modified from Santhosh, 2012), b)
Heat flow map of EGMB, India (cropped from Roy and Mareschal, 2011), c) Geology map with lineaments and geothermal sites (redrawn from Thussu, 2002), a-a′
transect used for generating topographic elevation map from Cartosat 1 (https://bhuvan-app3.nrsc.gov.in, access date and time: 22/11/2020, 8:28 pm) shown in
conceptual diagram (Fig. 12), d) Eastern Ghat Mobile Belt of Proterozoic age surrounded by Singhbhum, Bastar and Dharwar cratons. Mahanadi and Godavari
grabens divide the EGMB into 3 segments and the studied thermal sites fall in central segment (modified from Ramakrishnan, 1987).

2
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

current database, the thermal waters age ranges from 40 years in Hi­ respectively. Many faults and joints exist in Khondalite and Granitic
malayan province to 18.6 Ka in Godavari Valley (Chatterjee et al., 2018, Group of rocks, among which the lineaments trending NE-SW (i.e.,
Chatterjee et al., 2019) and the estimated temperature ranges from Taptapani to Tarabalo) and NW-SE (i.e. Athmalik to Atri) play a critical
<50 ◦ C (Himachal region, NW Himalaya) to 250 ◦ C (Puga Valley, role in thermal water circulation in EGMB (Fig. 1c). The generalized
Ladakh) (Craig et al., 2013; Tiwari et al., 2016). litho-units and digital elevation model of the study area are shown in
Thermal waters of EGMB did not attract much attention from re­ Fig. 2a and b.
searchers and authorities due to their low enthalpy and uncertainty over The Atri site falls in a lateritic terrain within the EGB Group of rocks
their sustainability. Studies on these thermal waters were mainly and has Charnockite and Granite rock exposures in the northwestern and
focused on the water quality and hydrochemical aspects (Thussu, 2002; Khondalite and Charnockite in the southern, southeastern and south­
Zimik et al., 2017 and references therein). But a few studies were also western parts. Gondwana Supergroup rocks are exposed in the north­
conducted to understand the source and recharge of the thermal waters eastern part. In general, the foliation shows moderately higher dips i.e.
using isotope techniques (Navada et al., 1995). The chemical geo­ 50◦ − 70◦ towards north. Four sets of joints are observed trending N-S, E-
thermometry estimated a wide range of reservoir temperatures from 92 W, NW-SE and NE – SW, and the thermal sprout is present at the
to 277 ◦ C (Navada et al., 1995; Mahala, 2019) while radiocarbon ages intersection of NE-SW and NW-SE trending joints (Fig. 2c). Tarabalo site
were found to vary from 5 − 18 Ka (Keesari et al., 2019). A few studies is located within alluvium of the granitic terrain surrounded by isolated
attempted isotope characterization of the non-thermal waters in this hill masses of Charnockite and Khondalite in the northeastern and
region (Farooq, 2018) but a detailed geochemical analysis of the thermal southeastern parts respectively while the exposures of Charnockite and
groundwaters and their intermixing with the non-thermal waters is still granite are present in the southern, western and northwestern parts
missing. Poor understanding of the origin and subsurface temperature of respectively. The foliation within the granite and Charnockite trends E-
thermal water, its recharge history and sustainability act as major W dipping 40◦ − 60◦ towards north while the foliation in the Khondalite
impediment to the development of strategies for economic exploitation trends NW-SE to ENE-WSW dipping 50◦ − 65◦ towards NE to NW. The
of these thermal waters. foliations within the granite are not prominent suggesting latter
We constrained the resident time and temperature of the thermal emplacement. The joints are found to be E-W and N-S trending in the
waters (age and heat content of subsurface thermal reservoirs), there by southern and WNW-ESE and NE-SW trending in the western and
trying to understand the upwelling of the juvenile/geothermal waters northwestern parts respectively. The multiple thermal sprouts are found
that is a novel observation in cratonic, crystalline aquifer with non- at the intersection of NW-SE and NE-SW lineaments (Fig. 2d). The
specific magmatic sources. The use of radioisotopes, chemical geo­ Athmalik site is located on NNE-SSW trending fracture and is very close
thermometers, mixing models and integrated multi-component geo­ to the intersection of NE-SW and NW-SE trending major fracture
thermometry provided unconventional tools to understand the hydro (Fig. 2e). Two sets of foliations are observed dipping NW and NE with
(geo)dynamic processes of thermal waters that can help to establish the high dip. The thermal sprout at Taptapani site fall in the contact zone of
future Enhanced Geothermal Systems (EGS) of this region. Khondalite and Charnockite rocks (Fig. 2f). Two sets of deep to shallow
By using a combination of stable and radioisotope tracers, hydro­ vertical fractures trending NE-SW and NW-SE are noticed. Massive grey
chemistry, geochemical modeling and geothermometry, we developed a granite and micaceous rock exposures with foliation having 57◦ dip
conceptual model for the geothermal reservoir and the associated fluid towards SE direction are also observed in this site.
circulations and pathways in the Mahanadi graben, EGMB. The objec­ The study area has tropical climate characterized by high tempera­
tives of this study are to; i) estimate the subsurface reservoir tempera­ ture, high humidity and medium-to-high rainfall. The annual rainfall
ture, ii) understand the hydrochemical evolution of thermal and non- varies from 1008 to 1988 mm with an average of 1481 mm. The water
thermal waters and their interactions, iii) estimate mean transit times levels range from 2 to 10 m below ground level during pre-monsoon
of thermal water as well as trace the flow paths and iv) develop a model (April 2016) and 2 to 5 m during post-monsoon (November 2016).
for the thermal and non-thermal water circulations in EGMB, Indian The depths of the sampled wells range from 80 to 180 m and the
Shield. groundwater flow is controlled by the topographic slope (Kundu et al.,
2002). The minimum aquifer thickness was found in the high topo­
2. Regional and structural components graphic area and it gradually increases towards the hot springs and at­
tains a maximum thickness of about 66 m near the Tarabalo hot spring
The EGMB falls in a tectonically stable, non-orogenic field of Indian (Farooq et al., 2018).
shield. This was formed along the eastern coast of peninsular India due
to Neoproterozoic collisions between the Indian cratons and east 3. Materials and methods
Antarctica (Dasgupta et al., 2013; Valdiya, 2016). These collisional
events and the subsequent successive deformations created a number of 3.1. Sampling
faults and shear zones, paving way for complex groundwater circulation
patterns and geothermal activity (Ramakrishnan, 1987; Valdiya, 2016). Water samples were collected from both thermal (sprouts) and non-
The Mahanadi and Godavari rift systems divide the EGMB into three thermal sources (dug wells, tube wells and bore wells) at four sites
segments, viz., northern, central and southern (Fig. 1d). The EGB is during pre-monsoon (39), i.e. May 2016 and post-monsoon (46), i.e.,
broadly divided into the three groups, viz., Khondalite Group, Char­ September 2016. A total of 19 thermal waters and 27 non-thermal water
nockite Group and Migmatite Group. Moving from east to west, the samples were collected. Thermal water discharges during post-monsoon
sequence of these groups is Eastern Khondalites, Central Migmatites, period were measured using float and stopwatch method. The sample
Western Khondalites & Charnockites and Westernmost Transition Zone locations for individual sites are shown in Fig. 2c–f. The field parameters
(Ramakrishnan, 1987). The geological succession of the study area is like pH, electrical conductivity (EC), temperature, and dissolved oxygen
shown in Table (S1). Petrological studies infer that the EGB Group of (DO) were measured on site using multi-parameter kit (Hanna Make).
rocks comprises garnet-sillomanite schist and gneisses (Khondalite The temperature range of this equipment is 0.0 to 70.0 ◦ C with a
rocks), hypersthene bearing granulite and gneisses (Charnockite rocks), sensitivity of ± 0.15 ◦ C. Gran method was used to determine alkalinity.
biotite bearing granite gneisses, pink granites and porphyritic granite. Before sample collection, wells were flushed/pumped until a stable
The EGB Group of rocks is directly overlain by the rocks of the Gond­ temperature is obtained. Multiple sets of samples were collected at each
wana Supergroup, which in turn are overlain by alluvium. The Khon­ site, which include a 100 mL of untreated water for anions, 100 mL of
dalite rocks display well developed foliation trending NW-SE to NNW- 0.45 µm filtered water acidified with concentrated HNO3 for cation and
SSE and dipping at high angles 45◦ -75◦ towards NE and north trace elements, 50 mL of unfiltered and untreated water for

3
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

Fig. 2. a) The generalized lithozones and lineaments falling in the study area (drawn from Ramakrishnan et al., 1987, Mahala, 2019), b) Digital elevation model of
the study area with thermal water sites identified, Sampling locations of thermal and non-thermal waters from c) Atri, d) Tarabalo, e) Athmalik and f) Taptapani sites.
Lineaments along with the elevation of the site and its surroundings are also shown. Numbers indicate sample ID’s listed in Table S2.

environmental stable isotopes and 500 mL of unfiltered and untreated calibration purpose. The chemical accuracy was determined by
water for environmental tritium. The details of collected thermal and computing error in charge balance (Eq. (1)), which was found to be
non-thermal water samples are provided in Table S2. within acceptable limits (±5%). The chemical data of pre- and post-
monsoon samples is presented in (Tables S3 and S4).

3.2. Chemical measurements meq(cations) − meq(anions)


CBE(%) = × 100 (1)
meq(cations) + meq(anions)
The anions (Cl , F ,
− −
NO–3
& SO2−
) and cations (Na , K , Mg &
4
+ + 2+

Ca2+) were quantified using Ion Chromatograph (Dionex DX-500)


coupled to a conductivity detector (ED 40). A total of 38 samples were 3.3. Stable isotope measurements
collected for chemical analysis (21 post-monsoon and 17 pre-monsoon).
The analytical columns used for anions and cations separation were A continuous flow Isotope Ratio Mass Spectrometer (IRMS, Isoprime
AS11 and CS12A respectively. Elution was performed by 20 mN H2SO4 100) located at Bhabha Atomic Research Centre (Mumbai, India) was
and 5–15 mN NaOH for cations and anions respectively. Suppressed used for isotope (2H and 18O) measurements. One milli-liter of water
conductivity detection was employed to quantify the ion concentrations sample was equilibrated with H2 and CO2 gases respectively for δ2H and
and the typical detection limits for major ions were 0.01–0.5 mg/L. The δ18O measurements. The operating conditions like, temperature, cata­
trace elements (Rb, Li, B, Cs and Si) were measured using ICP-OES and lyst and equilibrium time for δ2H and δ18O analysis are adopted from
the detection limits were in the range of 1–5 µg/L. Sigma-Aldrich 1000 Keesari et al. (2016). After equilibration, the gas was introduced into the
ppm standard solutions were used to prepare the working standards for mass analyser for isotope ratio measurements. For δ13C measurement,

4
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

dissolved inorganic carbon (DIC) was precipitated in the form of BaCO3 cylinders filled with CO2 gas were stored for 15–30 days to allow radon
in one-liter bottle by adding NaOH (to increase the pH above 10) and decay. Finally, the CO2 was absorbed in the mixture of carbosorb (11.5
carbonate free saturated BaCl2 solution. This precipitate was dried and mL) and permafluor (11.5 mL, scintillator) and counted in LSC (Quan­
transferred into a vial on site. δ13C analysis was performed on the CO2 tulus model 1220) for 20 cycles of 50 min (total 1000 min). The raw 14C
gas generated by combusting BaCO3 precipitate in an elemental analyser activity data was calibrated against SRM 4990C (Oxalic acid) and the
14
(EA) coupled with the mass spectrometer (IRMS, Isoprime 100). C data is reported as pMC (percent Modern Carbon). Nine samples
The results are reported in δ-notation and expressed in units of parts were collected for radiocarbon measurements (Table 1).
per thousand (denoted as ‰). The δ values are calculated using the Eq.
(2) (Coplen, 1996):
( ) 3.5. Estimation of reservoir temperature
Rx
δ(‰) = − 1 × 1000 (2)
Rs Since 1960s, chemical geothermometers have been widely used
2 1
worldwide to predict the subsurface temperature of the geothermal
where R denotes the ratio of heavy to light isotope (e.g. H/ H or reservoir. Commonly employed chemical geothermometers, also known
18
O/16O or 13C/12C) and Rx and Rs are the ratios in the sample and as “classical” geothermometers include silica (Fournier, 1977), Na-K
standard respectively. The measured values are normalized on VSMOW/ (Fournier, 1979, Arnorsson, 1983, Giggenbach et al., 1988), Na-K-Ca
SLAP scale. The precision of measurement for δ2H is ± 0.5 ‰ (2σ) and (Fournier and Truesdell, 1973) and K-Mg geothermometer (Giggen­
for δ18O is ± 0.1 ‰ (2σ). A total of 75 water samples were collected for bach, 1988). The implicit assumption in applying chemical geo­
isotope measurements (pre-monsoon 29 and post-monsoon 46). The δ2H thermometer is that there exists a chemical equilibrium between the
and δ18O isotope data is given in Table S5. δ13C values are reported interacting thermal water and aquifer mineral. For example, if the
against Vienna Pee Dee Belemnite (VPDB) and the precision of mea­ quartz geothermometer is used to estimate the reservoir temperature, it
surement is ± 0.30 ‰ (2σ). Laboratory CaCO3 salt calibrated against the is assumed that thermal water has attained equilibrium with Quartz
VPDB standard was used for daily calibration. The measurement preci­ mineral present in the host rock. Similarly, Na-K geothermometer
sion is estimated by repeated analysis of laboratory standards while the (Fournier, 1979, Arnorsson, 1983) is based on the equilibrium between
accuracy is estimated from the deviation of isotopic ratio of standard albite and K-feldspar minerals. The equations used for estimating
from the assigned value. reservoir temperature using chemical geothermometric methods are
provided in Table S6. In silica enthalpy mixing model, enthalpy derived
3.4. Radioisotope measurements from steam tables is used as an axis rather than temperature. This is
because the combined heat content (enthalpy) of the two waters is
For tritium (3H) measurement, 250 mL of water sample was distilled conserved on mixing whereas the combined temperature is not (Four­
and electrolyzed in a cool bath (1 – 4 ◦ C) for tritium enrichment. The nier and Truesdell, 1970). In this mixing model, the silica concentrations
enriched sample was mixed with scintillator mixture (8:12 mL) and of the examined samples were plotted against their corresponding onsite
counted for beta activity using an ultra-low background liquid scintil­ enthalpies. The enthalpy values were determined using the steam tables
lation counter (LSC, Quantulus model 1220) (Nair et al., 1990). The 3H of Keenan et al. (1969). For the application of this model two end
value is expressed in tritium unit (TU), where one TU has 3H/1H ratio member fluids were considered: viz., thermal (highest temperature
equals to 10− 18 and corresponds to an activity of 0.118 Bq/kg of water. water) and non-thermal (nearby cold water). The subsurface enthalpy of
The IAEA zero tritium blank and NIST 4926e standard were used for the geothermal reservoir can be estimated by extrapolating a line joining
calibrating raw data. The limit of detection of enriched sample for 500 the data points of local groundwater with that of hot springs which in­
min counting time is 0.4 TU (2σ) (Morgenstern and Taylor, 2009). The tersects the quartz solubility curve. The intersection point represents the
counting efficiency and the calibration factor of the counter are about silica content of the parent geothermal water before mixing. Next, a
25% and 70 TU/cpm respectively. A total 24 samples (11 pre-monsoon vertical line is drawn from the intersection point of the quartz solubility
and 13 post monsoon) were collected for this study and the data is given curve to the enthalpy axis and the intersection point in the enthalpy axis
in Table S5. For quality assurance, spiked samples and distilled water denotes the enthalpy of the parent geothermal water before mixing. In
samples are added in each batch of samples so that consistency of integrated multicomponent solute geothermometry (IMG) method, the
measurements is monitored. The measurements are periodically cross­ representative reservoir temperature is estimated based on the satura­
checked through IAEA inter-comparison exercise. tion state analysis of suitable mineral assemblages (Spycher et al.,
For radiocarbon (14C) measurement, the DIC was precipitated from 2014). The saturation indices (SI) of potential reservoir minerals over a
60 L of water and the procedure for BaCO3 precipitation is same as given range of temperatures are calculated and plotted as a function of tem­
for 13C isotope. In the laboratory, phosphoric acid was added to the perature. Reservoir temperature is estimated from the clustering of log
BaCO3 precipitate to evolve CO2 in a vacuum line, which was then (Q/K) curves near to the zero. A stand-alone computer program (GeoT)
passed through the cold trap (mixture of acetone and liquid nitrogen of is used to solve the mass balance/mass action equations using Newton-
temperature – − 70 ◦ C) to remove the water moisture from the gas. The Raphson iterative method (refer Spycher et al., 2014 for details). The

Table 1
Carbon isotopes data and modeled radiocarbon ages of thermal and non-thermal waters.
Sample ID δ13C (‰) 14
C (pMC) Error (1σ pMC) Apparent age T1 Negative Error Positive Error T2 Remarks
(years) (years) (years)

1 − 21 20.1 1.1 13,189 14,482 − 443 467 13,096 Very old


2 − 20 50.1 1.5 5594 7647 − 244 251 5066 old
3 − 19.3 87.1 1.8 1026 2820 − 169 173 179 modern
6 − 17.24 76.2 1.6 2128 3040 − 172 175 265 modern
8 − 16.22 55.7 1.5 4723 4363 − 220 226 2305 old
11 − 18.75 19.5 1.1 13,415 13,795 − 455 481 12,309 very old
15 − 14.71 98.6 2.1 3 − 326 − 174 178 − 3309 modern
19 − 19.88 10.1 1.1 18,919 19,712 − 863 963 18,337 very old
25 − 20.16 10.2 1.1 18,771 19,807 − 848 945 18,315 Very old

Note: T1 and T2 represent age estimates computed from IAEA model and equation given by Wang et al. (2019) respectively.

5
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

thermodynamic database SOLTHERM.H06 is used to compute the ion

average
activity product (Q), the thermodynamic equilibrium constant (K) and SI

107.3

161.3

162.9
32.1

11.7

87.3

80.3
23.3

34.7

76.4
values. GeoT also computes different statistical parameters i.e. median

793
8.0
0.4

3.7

4.7

3.5
(RMED), standard deviation (SDEV), mean (MEAN) and root mean
square error (RMSE) of SI values as well as the temperatures at which

maximum
these statistical parameters attain minimum value (TRMED, TMEAN, TSDEV

1256
34.2

19.0

94.0

50.0

66.7

14.7
Non-thermal n = 8
and TRMSE). For a perfectly clustered system, TRMED, TMEAN, TSDEV and

189

290
198

227
123
8.8
0.4

8.0

8.0

4.3
TRMSE should be identical and this temperature reflects the computed
reservoir temperature. The main limitations of this approach are; need

minimum
for exhaustive mineralogical data, dependency on steam loss and non-

<0.1
29.9

32.0

83.0
89.0
12.0

98.7
29.7
thermal mixing and sensitivity towards the choice of the thermody­

198
6.7
0.3

1.0
5.0
2.0

2.0
2.3
2.6
namic database (Pang and Reed, 1998).

average
4. Results

179.2

196.2
1147
48.8

13.2

93.2
97.3

62.8

73.7

21.7
219
126
8.0
0.3

9.5

3.0

3.7
The thermal discharge is highest at Atri site (233 lpm) followed by

maximum
Taptapani (82 lpm) and Athmalik sites (36 lpm). The thermal manifes­
tations at Tarabalo site are at multiple points and are under marshy

1312
58.4

12.0
16.0

99.0

78.0

78.3

22.1
218

110
244

232
128
8.3
0.5

4.0

4.1
condition, hence the discharge could not be quantified accurately. The

Thermal n = 13
previous reports estimated thermal discharges in the range of 33–420

Minimum
lpm and found overall discharge rates constant throughout the year

Athmalik
(Thussu, 2002; Mahala, 2019).

1084
38.8

32.0

85.0
80.0
12.0

69.0

21.2
206
124
7.6
0.2

1.0
5.0
2.0

2.0
2.3
4.1. Hydrochemical characteristics

average

970.3

<0.1
29.5

59.0

47.0
34.0
51.0

38.0

36.6
The temperature of the thermal water ranges from 38.8 to 58.4 ◦ C in

419
8.2
0.3

1.0

6.0
1.4
1.3
4.5
pre-monsoon and 35.8 to 58.9 ◦ C in post-monsoon seasons. The non-
thermal waters show temperature variations from 29.0 to 34.2 ◦ C in

maximum
pre-monsoon and 26.6 to 38.5 ◦ C in post-monsoon (Fig. S1). Irrespective

1592
29.7
of location, the thermal waters are colorless and have gas bubbling with Non-thermal n = 6

8.6
0.3
sulfur odor while the non-thermal waters are colorless and odorless.
Navada et al. (1995) reported similar thermal water temperatures dur­
minimum

4 , F , – mg/L; Rb, Li, B, Cs - µg/L.


ing 1995, suggesting that the temperature of thermal water did not
29.0

change for the past 25 years.


645
7.6
0.3

+
During pre-monsoon, the non-thermal waters show a wider variation
in electrical conductivity (198–1592 µS/cm) than thermal waters
average

52.3

61.3
85.0

23.7
12.6

42.4
65.0
8.5
0.4

4.7
2.7
2.3

1737.5

1.4
793.7

143.7

144.7
(788–1312 µS/cm). A similar trend is observed in the case of post-
monsoon samples. The summary of the hydrochemical data is pro­
vided in Table 2 and 3.
maximum

Dominance of ions in thermal waters during post-monsoon period is


in the order of;

3434
56.9

65.0
99.0

24.0
13.2

44.4
68.8
797

145

147
8.6
0.5

5.0
3.0
3.0

1.5
Na+ (150–220 mg/L) > Ca2+ (4–18 mg/L) > Mg2+ (0.5–11 mg/L).
Summary table for thermal and non-thermal samples during pre-monsoon season.


Thermal n = 5

Anions follow the order,

Note: Temp- oC; EC- µS/cm; DO, Na+, K+, Ca2+, Mg2+, SiO2, HCO–3, Cl− , SO2−
Cl− (140–245 mg/L) > HCO–3 (54–99 mg/L) > SO2−
minimum

4 (36–101 mg/L).
Tarabalo

Only thermal water at Taptapani site show a conspicuous reduction


43.6

57.0
70.0

23.0
12.2
41.0
40.3
61.2
788

143

143
8.5
0.3

4.0
2.0
2.0

1.3
in Cl− (10 mg/L) and SO2− 4 (5 mg/L) concentrations compared to other
thermal waters.
average

The non-thermal waters show the cation order similar to thermal


8.0
0.3

2.7

6.3
4.6
8.2
663.8

255.7
30.7

67.7

38.0
17.3
53.7

58.3

14.2
56.1

waters,

Na+ (18–180 mg/L) > Ca2+ (6–70 mg/L) > Mg2+ (1–35 mg/L).
But the anions show a different order compared to the thermal
maximum

waters,
352.0

HCO–3 (102–446 mg/L) > Cl− (8–138 mg/L) > SO2−


32.0

61.0
30.0
60.0

13.0
12.3
21.5
31.4
90.0
Non-thermal n = 6

4 (0–24 mg/L).
997

149

120
8.9
0.4

4.0

The major ion trends in thermal and non-thermal waters are shown
in Schoeller diagrams (Fig. S2a and b). No significant difference in the
minimum

major ion concentration ranges is observed among non-thermal waters


<0.1
29.3

10.0

50.0

24.0

34.5

of different sites. The Piper’s classification shows that the thermal and
333

198
7.3
0.3

2.0
6.0
3.0

0.0
0.4
1.3
4.0

non-thermal waters have different facies suggesting different sources or


processes governing geochemistry (Fig. 3a). The hydrochemical facies of
Thermal n = 1

the samples indicate that thermal waters barring Taptapani site, show
dominance of Na+ and Cl− ions (cluster-i) in both seasons. However,
253.0

four clusters are observed for non-thermal waters, cluster-ii (dominated


1192
57.6

59.0
40.0

26.0

63.0
67.3
53.2
Atri

180
8.6
0.5

8.0
8.0
3.0

9.2

4.1

by Ca2+ and HCO–3) indicates recharging nature of water while cluster-iii


(dominated by Na+ and HCO–3) indicates influence of water-rock inter­
Table 2

HCO–3
Temp

Mg2+

action mainly silicate weathering and ion exchange reactions as dis­


Ca2+

SiO2

SO2−
Na+

+
4

Cs+
DO

Li+
pH

K+
EC

Rb
Cl−

F−

cussed in section 5.2. A few non-thermal samples also fall in cluster-i

6
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

indicating the possibility of mixing with thermal sources, which is

minimum maximum average


supported by warmer temperatures in those samples (34.9–38.5 ◦ C).

272.3

150.3
38.7

54.3
Samples falling in cluster-iv with relatively higher EC values (~750 µS/

6.7
3.3

2.3
1.9
29

20

14
6

6
cm) suggest dissolution of salts from the formation. The enrichment of
Cl− (average: 169 mg/L) and Na+ (average: 170 mg/L) in thermal waters
Thermal n = Non-thermal n = 6
over non-thermal waters (average Cl− : 29.26 mg/L; average Na+: 65.6

38.5
382

236
7.3

4.4
70
15
25

67

21
5

4
mg/L) may be attributed to greater leaching of minerals present in rocks
due to elevated temperatures (Tables S3 and S4). The common sources
for Na+ and Cl− could be minerals like biotite, hornblende and feldspar
present in the study area (Mahala, 2019; Maitra et al., 2020). Among
trace elements, fluoride (2.25–13.2 mg/L) concentration is found to be
Taptapani

higher while boron (53.2–128 µg/L), lithium (40.3–232 µg/L) and


rubidium (41–3434 µg/L) are lower in concentrations. Like, major ions,
42.7
369

225
7.4
2.1

4.9
72

18

62

10
the trace elements concentrations are also higher in thermal waters
1

5
compared to non-thermal waters (Table S4), which underlines the in­
minimum maximum average minimum maximum average minimum maximum average minimum maximum average minimum maximum average

694.1

274.7
31.3

81.7

30.3

17.3
fluence of greater water – rock reactions favoured by high temperature.
7.3
2.5

2.6

3.3

7.4
97

11
4.2. Stable isotopes (2H, 18
O and 13
C)
Non-thermal n = 8

1311
34.9

130

132
362
8.5
4.1

9.2
70

12
23

The δ18O and δ2H values of the thermal waters vary from − 6.30 to
5

− 2.80 ‰ and − 36.8 to − 14.02 ‰ respectively during post-monsoon and


− 4.42 to − 2.96 ‰ and − 26.05 to − 15.34 ‰ respectively during pre-
29.5
174

188
6.9
1.1

0.7

4.3
55

78

10
14

monsoon (Table S5). The thermal water from Taptapani site shows
6
1

most depleted isotopic signature (δ18O: − 6.3 ‰) as compared to other


218.3
1090

95.3

sites (Table S5). Like thermal water, the non-thermal waters also show
200

121
0.6

8.3

1.5
47

17

92
8

relatively depleted isotope values at Taptapani site (average δ18O and


δ2H, − 6.4 ‰ and − 40.9 ‰). In general, a wide scatter in isotope data is
observed in non-thermal waters, δ18O: − 5.28 to 0.98 ‰, δ2H: − 31.94 to
1130
58.9
Thermal n = 13

220

126

224
101
8.6
4.2

5.1
18

99

− 3.58 ‰ in pre-monsoon and δ18O: − 6.82 to 0.87 ‰, δ2H: − 45.4 to


9

− 0.97 ‰ during post-monsoon season.


Athmalik

The isotope plots of the thermal and non-thermal waters are shown
35.1
975

190

117

214
7.6

4.7
16

86

85

in Fig. 4a, b. The Global Meteoric Water Line (GMWL: δ2H = 8 × δ18O +
0

10) and Local Meteoric Water Line (LMWL) of the study area (δ2H =
1031.8

7.92 × δ18O + 5.74, after Kumar et al., 2010) are also shown in the plots.
29.4

446
7.5
0.8

0.7

1.7
61

57
35
62

32
11

The possibility of presence of any magmatic component is ruled out


considering the wide contrast between the thermal water isotope values
Non-thermal n = 6

and magmatic sources isotope signatures, which are enriched in δ18O


1915
30.3

7.9

(+6 to +9 ‰) and depleted in δ2H (− 40 to − 80 ‰, Giggenbach, 1992).


2

Low EC (369–1130 µS/cm), low Cl− contents (10–245 mg/L) and


4 , F – mg/L.

absence of enriched stable isotope (δ18O & δ2H) values preclude the
28.9
571
7.2

possibility of marine contribution to the studied thermal waters. The


0
Summary table for thermal and non-thermal samples during post-monsoon season.

geothermal waters fall along the GMWL and LMWL clearly demon­

714.2

162.5

strating their meteoric origin (Fig. 4a) and the slope close to 8 indicates
50.2

81.3
87.5

36.8
142
8.8
2.4

3.8
5.8
0.8

Note: Temp- oC; EC- µS/cm; DO, Na+, K+, Ca2+, Mg2+, SiO2, HCO–3, Cl− , SO2−
16

the recharge without significant evaporation (Clark and Fritz, 1997;


Kumar et al., 2010). No δ18O-shift in the δ2H vs. δ18O plot is observed in
thermal waters (Fig. 4a), which further indicates that the reservoir
57.1

16.2
721

170

144
8.9
4.3
Thermal n = 5

86
97

38

temperature is not sufficient enough (i.e. >150 ◦ C) to shift δ18O value


5
7
1

through extensive water – rock interaction. In geothermal systems with


Tarabalo

high water to rock ratio the fresh rock surface available for isotopic
44.3

15.6
706

150

140
8.6

0.5

exchange would be very small resulting low to negligible 18O-shift


77
70

36
0

2
4

(Nicholson, 1993).
682.2

261.3

There are multiple sources for DIC in subsurface waters like atmo­
31.1

14.3
68.7

61.7
11.3
7.2
2.8

1.7

5.5
78

43

spheric CO2, oxidation of dissolved organic matter, carbonate or silicate


mineral weathering, which can be identified through examining the
Thermal n = Non-thermal n = 6

δ13C-DIC (Clark and Fritz, 1997; Das et al., 2005). The average δ13C of
1050
32.2

180

403
138
8.2
6.6

69
26
80

24
15

the atmospheric CO2 is − 8.1 ‰ (Cerling et al., 1991). C3 vegetation is


3

characterized by more depleted δ13C composition (-25 ‰) than for C4


plants (-13 ‰). The magmatic CO2 shows δ13C values from − 7.5 to − 5.5
30.4
317

161
6.6
0.8

0.9
0.4

‰ VPDB (Exley et al., 1986). The marine carbonates have enriched δ13C
20

63

19
1
6
2

values (0 ‰) while the organic sources are characterized by depleted


δ13C values − 30‰ (Evans et al., 2008). Geogenic CO2 present in deeply
circulating geothermal systems typically has high pCO2 and moderately
1019
57.7

11.8
Atri

180

245
8.5

enriched δ13C value (-6 ‰, Marty and Jambon, 1987). In the present
11
11
98
54

42
1

study, the δ13C values range from − 21 to − 16.2 ‰ for thermal waters
Table 3

HCO–3
Temp

Mg2+

and − 20 to − 14.7 ‰ for non-thermal waters (Table 1). These ranges


Ca2+

SiO2

SO2−
Na+

4
DO
pH

K+
EC

Cl−

F−

overlap with the δ13C – DIC derived from silicate weathering reactions

7
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

Fig. 3. a) Piper’s plot of major ions of thermal and non-thermal waters, b) facies variations in thermal springs of different sites illustrating the minor differences due
to host rock composition, c) geochemical evolution of non-thermal waters and d) Ca2+/sum cations versus temperature plot indicating the shallow non-thermal
water mixing.

(Das et al., 2005). The contributions from marine carbonate, mantle tritium are also reported from other geothermal systems in India (Nav­
degassing and organic matter oxidation seems to be negligible or absent. ada et al., 1995; Chatterjee et al., 2018). In the present study, the tritium
These aspects are further discussed in section 5.4. content of non-thermal waters from all sites ranges from 0.9 to 3.4 TU
indicating modern recharge (Table S5). The 14C values range from 10 to
56 pMC while non-thermal waters show values from 50 to 99 pMC
4.3. Radioisotopes (3H and 14
C)
(Table 1). A close examination of the data suggests that 14C values of the
thermal waters with highest temperature are 20 pMC or lower. Similar
Tritium content in thermal waters provides estimation of residence
low 14C values have been reported from Godavari Valley geothermal
time and shallow groundwater mixing proportions (Sammel and Craig,
systems (Chatterjee et al., 2019). The radiocarbon dating of thermal
1981; Chatterjee et al., 2018) Samples with tritium >1.0 TU indicate
waters is further discussed in section 5.4.
recent recharge or mixing with modern water whereas low tritium (<1.0
TU) indicates recharge prior to 1951 (Clark and Fritz, 1997). Tritium
5. Discussion
data of thermal waters during both seasons (pre-monsoon: 0.4 – 0.9 TU
and post-monsoon 0.4–1.5 TU) is typically 1.5 TU or less (Table S5). This
5.1. Subsurface temperature estimation
suggests that modern recharge to thermal waters is negligible and the
residence time of thermal water could be >50 years. Low tritium (<1
5.1.1. Chemical geothermometry:
TU) thermal waters with high temperatures up to 97 ◦ C are also
A wide scatter is observed in the estimated reservoir temperatures at
observed in other geothermal fields of India (Sarolkar, 1996; Chatterjee
different sites of the study area (Table 4). Quartz geothermometer (no
et al., 2016, 2017). The thermal water at Taptapani site has relatively
steam loss) estimated subsurface temperatures from 112 to 136 ◦ C and a
higher tritium, 1.5 TU, suggesting modern recharge (<50 years) or
similar range of temperatures (111–132 ◦ C) is observed from adiabatic
mixing with modern waters during ascent. Thermal waters with modern

8
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

Fig. 4. Plots of δ2H vs. δ18O of a) thermal waters (GMWL- Global Meteoric Water Line, LWML – Local Meteoric Water Line (Kumar et al., 2010), Rainwater data taken
from Keesari et al. (2016) and b) non-thermal waters, EL – evaporation line, the mean isotope values (with 1σ deviation) of thermal waters of respective site
are mentioned.

Table 4
Estimates of chemical geothermometry, silica enthalpy mixing and integrated multicomponent solute models and reservoir depth.
Sample ID T Q1 Q2 C Na-K (1) Na-K (2) Na-K (3) Na-K (4) Na-K (5) K-Mg Na-K-Ca Na-K-Ca Avg std SiM IGT RD
Mg corr

1 (Atri) 58 136 132 109 156 175 141 144 125 76 148 * 134 3 145 118 2.3
11 (Tarabalo) 57 123 121 95 127 147 104 116 92 65 131 * 122 1 127 112 2.1
12 (Tarabalo) 56 112 111 83 141 160 122 129 108 70 139 * 112 1 ** ** **
19 (Athmalik) 59 133 129 106 164 182 151 152 133 87 151 105 131 3 163 124 2.4
25 (Athmalik) 56 133 129 106 174 192 164 162 145 92 159 109 131 3 ** ** **
28 (Athmalik) 57 132 128 104 165 183 152 153 134 87 118 101 130 3 ** ** **
40 (Taptapani) 43 112 111 83 91 112 62 81 54 * * * 112 1 119 92 1.6

Note: T – thermal water temperature (◦ C), Q1-Quartz conductive, Q2-Quart adiabatic, C- Chalcedony, * estimates lower than thermal water temperature, **- not
determined. The silica and cation geothermometer equations are given in Table S6. Avg- mean of silica geothermometry estimates, std – standard deviation, SiM- Silica
enthalpy mixing model, IGT- Integrated geothermometry model, RD –Reservoir depth in km.

quartz geothermometer (maximum steam loss), indicating no appre­ mixing fraction to thermal water is estimated to be 0.12 to 0.63
ciable steam separation from the thermal water system. Chalcedony (Table S5). The end member pairs for Atri, Tarabalo and Athmalik and
geothermometer estimated temperatures (83–109 ◦ C) lower than Quartz Taptapani sites are ID-1 & 3, ID-11 & 13, ID-19 & 31 and ID-40 & 42
geothermometer values. The Na-K-Mg ternary diagram aids in inferring respectively. The average silica values (pre- and post-monsoon data) are
whether thermal water is immature or in partial equilibrium or attained used in the plot. Points A, B, C and D represent the enthalpy and the
full equilibrium with the rocks at a given temperature (Giggenbach, silica content of the parent geothermal water before mixing for Tapta­
1988). The sample data falls near Mg-corner in the ‘Immature waters’ pani, Tarabalo, Atri and Athmalik sites respectively (Fig. 6). The sub­
region, suggesting that the estimates from Na-K geothermometers may surface temperatures computed using this method (119 to 163 ◦ C) are
be erroneous (Fig. 5a). This is further witnessed in the wide range of found to be slightly higher than the estimates obtained from chemical
subsurface temperatures (54–192 ◦ C) estimated from the Na-K geo­ geothermometers (Table 4).
thermometers. The Na-K-Ca and magnesium corrected Na-K-Ca geo­
thermometers estimate reservoir temperatures from 118− 159 ◦ C and 5.1.3. Integrated multicomponent solute geothermometry
101–109 ◦ C respectively (Table 4). The plot of 10 K/(10 K + Na) vs. 10 Both the classical geothermometry and silica enthalpy mixing
Mg/(10 Mg + Ca) is presented in Fig. 5b. The sample data falls above the models though provided a similar range of reservoir temperatures, the
line inferring that thermal waters are not completely equilibrated with estimates for individual reservoirs are not consistent. This is a common
alkali feldspars. The subsurface temperature estimated from K-Mg geo­ issue for the low to medium enthalpy geothermal systems owing to non-
thermometer ranges from 65 to 92 ◦ C and is lower than the Quartz attainment (or partial attainment) of equilibria. This limitation is
geothermometer estimates. This lowering of the estimated temperatures addressed by applying IMG method. In this study; quartz, magnesite,
suggests contribution of near surface rock-water interaction or non- albit-lo, enstatite, hercynite, grossular, sanidine, pyrite, osumilite are
thermal water mixing. From the above discussion, temperature esti­ taken as ten main clustering minerals based on the geological informa­
mates of Quartz geothermometry are found to be reliable and repre­ tion of the study area (Dasgupta, 1993; Mahala, 2019; Maitra et al.,
sentative. The average reservoir temperatures for Atri, Tarabalo, 2020). GeoT is applied over the four thermal water samples (Samples ID-
Athmalik and Taptapani sites are found to be 134 ± 3 ◦ C, 117 ± 1 ◦ C, 1, 11, 19 and 40) and deep fluid temperature is computed from the ten
131 ± 3 ◦ C and 112 ± 1 ◦ C respectively (Table 4). best clustering minerals. For ID-1 (Atri thermal water), the chemical
composition of water suggests equilibrium with minerals such as quartz,
5.1.2. Mixing model calcite, magnesite, grossular, hercynite, sanidine, osumilite etc. (Fig. 7a)
Positive correlations between the conservative elements and pres­ and estimated equilibrium temperature (TRMED) is found to be 118 ◦ C
ence of environmental tritium in thermal water confirm the mixing (Fig. 7b). In case of ID-11 (Tarabalo thermal water), the clustering is
between thermal and non-thermal waters (Table S5). The non-thermal observed with minerals such as quartz, albit-lo, calcite, magnesite, dolo-

9
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

range from 92 to 124 ◦ C and fall on the lower side of the values obtained
from silica enthalpy mixing model and chemical geothermometry
(Table 4) but the spread in the estimates is much smaller than the pre­
viously reported values, 92 to 277 ◦ C (Navada et al., 1995). These es­
timates indicate a lower enthalpy thermal system in Mahanadi graben
compared to SONATA mobile systems, conforming to the heat flow
values of Indian shield (Kumar et al., 2007; Roy and Mareschal, 2011).

5.2. Geochemical processes

Along the flow groundwater accumulates salts leading to higher EC;


therefore, the geochemical reactions are examined in tandem with EC
values in Piper’s trilinear plot (Fig. 3b and c). In the case of thermal
waters, a clear distinction is noticed in the chemical facies. Athmalik
thermal water shows Na-Cl-SO4 facies while Atri and Tarabalo sites
show Na-Cl-HCO3 facies. Similar nature of subsurface geology at these
sites could be the reason for the observed similarity in water facies.
Thermal waters represented by Na-Cl type with high salinity (~5000 µS/
cm) are common in the West coast geothermal systems in India due to
mixing with marine waters (Chandrasekhar et al., 2018), however in this
study, though the hydrochemical facies is Na-Cl type, the salinity range
is lower (EC ~ 1000 µS/cm) in thermal waters. A similar pattern of low
EC and Na-Cl facies of thermal waters is observed in the Hazaribagh and
Dumka (Jharkhand) and Bakreswar and Tantloi (West Bengal)
geothermal regions (Singh et al., 2020).
The second dominant anion, after Cl− , in the thermal waters of Atri/
Tarabalo and Athmalik sites is HCO–3 and SO2− 4 respectively. The HCO3

concentration of thermal waters is similar in these three sites (54–99


mg/L), however the SO2− 4 concentration is lower (<40 mg/L) in Atri/
Tarabalo sites but relatively higher (80–100 mg/L) in Athmalik site
(Tables S3 and S4). This clearly suggests that the thermal waters are
circulating through different rock types in these sites. At Athmalik site,
Fig. 5. a) Na–K–Mg Giggenbach plot determining the water rock equilibrium the thermal waters issue along the northern fringe while at Atri and
conditions and b) Plot of 10 K/(10 K + Na) vs. 10 Mg/(10 Mg + Ca) (after Tarabalo sites the thermal waters issue along the southern fringe of the
Giggenbach, 1988). Mahanadi graben (Fig. 2a). Northern part of the EGMB is dominated by
Iron Ore Supergroup consisting of pyritic minerals (Mahala, 2019).
Oxidation of pyrite releases high SO2− 4 and H , which reduces the pH of
+
2−
the water. High SO4 concentration and reduced alkaline nature of
thermal water at Athmalik site (pH: 8.0 ± 0.3) compared to other sites
(pH: 8.7 ± 0.2) further substantiate this observation.
In contrast, thermal water at Taptapani site is Na-Ca-HCO3 type
water. Similar type of thermal water was reported in Taptapani and
Salbardi areas of SONATA belt and Rajapur area of West coast
geothermal systems (Minissale et al., 2000). Relatively higher concen­
tration of Ca2+ and HCO–3 in Taptapani site can be attributed to
weathering of mixed plagioclase minerals. Lower temperature (42.7 ◦ C),
lower EC (369 µS/cm) and neutral pH (7.38) compared to rest of the
thermal waters as well as its location away from the Mahanadi graben,
as shown in Fig. 1c, d and 2a, indicate that the geothermal reservoir is
different for this site. This observation is further supported by its distinct
isotopic as well as geothermometry values, discussed in subsequent
sections. Thermal waters do not show any temporal variations in
chemical composition whereas non-thermal waters do (Fig. 3a–c).
The source rocks contributing to the dissolved species in thermal
waters are evaluated by ionic ratios. Low HCO–3/SiO2 ratio (<5) and
Fig. 6. Silica-enthalpy diagram for estimating the reservoir temperatures. intermediate (Na++K+-Cl− )/(Na++K++Ca2+-Cl− ) ratio (0.2 to 0.8)
shown in Table S7, infer silicate weathering (Hounslow, 1995). The
dis, osumilite, sanidine, enstatite etc. (Fig. 7c) and the estimated equi­ plots of Na-normalized molar ratios (Mg2+/Na+ vs. Ca2+/Na+ and
librium temperature is (TRMED) 112 ◦ C (Fig. 7d). In Athmalik area (ID- HCO–3/Na+ vs. Ca2+/Na+) also suggest the contribution of evaporite and
19), quartz, albit-lo, magnesite, grossular, osumilite, sanidine etc. are silicate minerals towards dissolved ions in thermal waters (Fig. 8a and
found to be the main clustering minerals (Fig. 7e) and the estimated b). None of the samples fall in the carbonate mineral zone indicating
reservoir temperature (TRMED) turns out to be 124 ◦ C (Fig. 7f). In Tap­ less/no influence of carbonate minerals on the hydrochemistry of ther­
tapani site (ID-40), quartz, magnesite, osumilite, sanidine, calcite, mal and non-thermal waters. Presence of high SiO2 and high F− further
dolomite, etc. are found to be the main clustering minerals (Fig. 7g) and supports the contribution of minerals from granitic and hornblende
the estimated reservoir temperature (TRMED) is found to be 92 ◦ C gneissic rocks. These observations conform to the nature of rocks present
(Fig. 7h). It is observed that the estimates of reservoirs temperature in EGMB (Mahala, 2019; Maitra et al., 2020). The sample data fall in the

10
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

Fig. 7. Computed saturation indices (SI) as well as statistical analysis of SI as a function of temperature using multicomponent geothermometry (GeoT) for a, b) Atri
(ID-1), c, d) Tarabalo (ID-11), e, f) Athmalik (ID-19) and g, h) Taptapani (ID-40). The convergence point is taken as representative temperature of the ther­
mal reservoir.

chloride region of the Cl-SO4-HCO3 ternary diagram in the case of Atri, 200 to 650 µS/cm. Some samples with higher EC (~750 µS/cm) are
Tarabalo and Athmalik sites indicating that the thermal waters are not characterised by Na-Cl-HCO3 type, similar to that of diluted thermal
steam heated local groundwater and are mature in nature (Fig. 8c). water, suggesting mixing between thermal and non-thermal waters. It is
Thermal water from the Taptapani region falls in the bicarbonate region observed that the less percent of Ca2+ the thermal water contains the
representing mixed water or immature waters. higher temperature it has (Fig. 3d). The facies variations can result from
The non-thermal waters show low EC and the hydrochemical facies various geochemical reactions such as weathering of minerals, ion ex­
are either Ca-Mg-HCO3 type (cluster i) or Na-HCO3 type (cluster iii) change or evaporite dissolution. From Fig. 9a and b, it can be concluded
evolving to Na-Ca-HCO3 or Ca-Na-HCO3 type (cluster ii) as shown that the silicate weathering exerts a major control on the chemistry of
Fig. 3c. These facies variations agree with the increasing EC values from non-thermal waters, which is supported by presence of moderately high

11
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

Fig. 8. Plots of Na-normalized molar ratios of a) Mg2+ vs. Ca2+ and b) HCO–3 vs. Ca2+ demonstrating the relative contribution of evaporite, silicate and carbonate
minerals, c) Anion variation diagram (Cl-SO4-HCO3 ternary diagram) classifying thermal waters into different water fields and d) plot of Na+-Cl− vs. Mg2++Ca2+-
SO2−
4 –HCO3 illustrating ion exchange process in non-thermal waters.

silica concentration (37–132 mg/L). The ion plot of Na+-Cl− vs. different reservoirs or different flow paths (Wright et al., 1991), while
Mg2++Ca2+-SO2− 4 –HCO3 evaluates the ion exchange process occurring

the near uniform ratios observed in the case of SiO2/Cl− (wt. ratio
in thermal and non-thermal waters (Fig. 8d). The non-thermal waters 0.4–0.6) and (Na++K+)/(Cl− +F− ) (wt. ratio 0.72 – 1.1) reflect signifi­
fall along the line with slope − 1 indicating Na+ in water is being cant interaction with country rocks (Table S7). These observations agree
replaced by Ca2+ from clay. But, thermal water data fall as a cluster in with the inferences obtained from SI variations (Table S8). The ionic
the plot indicating less significance of ion exchange process. ratios are quite different in the case of Taptapani thermal water sug­
In order to estimate the extent of water – mineral equilibrium, gesting completely a different source reservoir and/or flow path for
saturation indices (SI) were estimated using the formula (Eq. (3)); thermal water. Presence of high B and Cl− concentrations in thermal
waters suggests deep circulation paths and extended interaction with the
IAP
SI = − log (3) rocks (Arnorsson, 2000). The Cl− /B ratio is used to trace the ground­
Ksp
water flow paths and assess the mixing proportions of different water
Where IAP – ion activity product, Ksp – solubility product. The SI = 0, masses in geothermal systems (Aggarwal et al., 2000). Atri, Tarabalo
>0 and <0 indicates saturation, supersaturation and unsaturation and Athmalik sites exhibit different Cl− /B ratios, 4750, 2237 and 1810
respectively. WATEQ4F program was used to compute SI of common respectively. But the thermal waters located at a particular site showed
minerals. The ranges of saturation indices of thermal and non-thermal similar Cl− /B ratios (Table S7). These non-uniform Cl− /B ratios at
waters with respect to common minerals are represented in Table S8. different sites demonstrate that these conservative constituents are
Saturated to supersaturated condition is exhibited for silicate minerals originated from different reservoirs (Fournier and Truesdell, 1970).
while carbonate minerals demonstrate unsaturated to saturated condi­ The interrelationships among Li, Rb and Cs also help in evaluating
tion. Sulphate and fluoride minerals demonstrate unsaturation in both the secondary processes. Due to the large differences in the respective
thermal and non-thermal waters. The degree of saturation seems to be concentrations of these elements, a scaling factor is applied to spread the
higher for thermal waters than non-thermal waters, which agrees with data evenly in the Li-Rb-Cs ternary diagram. It is observed that the
the enrichment of major ions as well as trace elements. However, the SI thermal water data of Atri, Tarabalo and Athmalik sites fall away from
variations are similar for both thermal and non-thermal waters, which the crustal rock’s composition indicating the presence of different sec­
can be attributed to their interaction with common rock types. This ondary processes (Fig. 9a). Thermal waters in Athmalik site show lowest
could also imply that the thermal water circulations are not deep enough Li+/Cs+ ratio (wt. ratio 9.7 – 15.4) whereas Tarabalo thermal water
to encounter different kind of rock types in EGMB. shows highest Li+/Cs+ ratio (wt. ratio ~30), indicating incorporation of
The characteristic ionic ratios of thermal waters are used to infer the Cs+ in clay minerals at Tarabalo site (Table S7). In the Cl-B-Li ternary
diagram, the sample data falls near the chloride corner, which suggests
nature of reservoir and the associated flow paths. The Na+/Ca2+ ratio
(wt. ratio, 1.0 – 40) is different for different thermal waters indicating that these thermal waters are part of a very old hydrothermal system

12
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

0.001), indicating the means belong to different population (α = 0.05), i.


e. the thermal waters are fed by different reservoirs. Presence of multiple
subsurface thermal reservoirs are reported by Chatterjee et al. (2017) in
peninsular India, viz., the Manuguru geothermal area (Telangana).
Compared with the local rainwater (δ2H: –32 ‰ and δ18O: − 5.5 ‰,
Keesari et al., 2016) the isotope values are enriched for Athmalik, Tar­
abalo and Atri thermal waters, possibly indicating the recharge during
arid climate (discussed in section 5.4). Depleted stable isotope values
(δ2H: − 36.8 ± 0.8 ‰ and δ18O: − 6.3 ± 0.2 ‰) and moderate temper­
ature of the thermal water at Taptapani site indicate a greater mixing of
thermal water with non-thermal water, which can also be substantiated
geochemically. The chemical facies of the Taptapani thermal water is
found to be Na-Ca-HCO3 type whereas the other thermal waters are Na-
Cl type. The preponderance of bicarbonate type of thermal water in
Taptapani area indicates a greater extent of mixing with non-thermal
waters, which is also corroborated from high tritium value of the Tap­
tapani thermal water (1.5 TU) compared to other thermal waters (<1
TU). The approximate degree of mixing is estimated from tritium con­
centrations assuming the deep-seated thermal water has zero tritium
content (Sammel and Craig, 1981). The percentage of non-thermal
water contribution was found to be highest for Taptapani site (63%)
followed by Atri/Tarabalo (~20%) and Athmalik site (~10%). These
estimated mixing proportions justify the lower temperature and
depleted isotopic composition of the thermal water at Taptapani site and
higher temperature and enriched isotopic composition of the thermal
waters in other sites. The EC correlations with δ2H and δ18O also indi­
cate different groups for thermal waters collected from different sites
(Fig. S3), which rules out the possibility of a common reservoir as the
source for the thermal waters in all the four studied sites. This inference
further strengthens the observations obtained from the hydrochemical
classification (Fig. 3a and b) and isotope trends (Fig. 4a).
The deuterium excess (Dx = δ2 H − 8 × δ18 O) provides a reliable
Fig. 9. Ternary diagrams of a) Li-Rb-Cs and b) Cl-B-Li for themal waters in measure of the physical conditions (surface temperature, humidity) of
different sites. the source region responsible for vapor transport to the site of precipi­
tation (Dansgaard, 1964). The Dx of oceanic origin is normally + 10 ‰
(Fig. 9b). The relatively low B content in the thermal springs (<0.13 mg/ in temperate climate (Clark and Fritz, 1997). The mean Dx values of
L) reflects the high degree of maturity in the geothermal system thermal water during pre- and post-monsoon periods are found to be
(Table S4). During the early heating up stages, volatile components such similar in this case, i.e., 8.6 ± 1.3 ‰ and 9.0 ± 2.3 ‰ respectively, while
as B are likely to be expelled from the system. Thus, fluids from the thermal water at Taptapani site has slightly higher Dx value (+13.6 ‰).
‘older’ hydrothermal systems normally show lower B concentration. The Dx values of the thermal waters suggest their oceanic origin with
some modifications during the infiltration. The modifications include
5.3. Isotope systematics contribution of recirculated moisture, evaporative enrichment or mixing
with different moisture sources. The non-thermal waters also fall along
The stable isotopic composition of the thermal waters indicates the GMWL and LWML indicating precipitation recharge to shallow
different clusters, as shown by shaded regions in Fig. 4a. The isotope aquifer system (Fig. 4b). A few samples mainly from Tarabalo site fall
data of these thermal waters measured during 1995 by Navada et al. below LMWL on a line with slope 3.49 indicating evaporation effect.
(1995) also fall in the respective groups indicating that the reservoir Higher evaporation is observed in samples during pre-monsoon
isotope characteristics have not changed during the last 25 years. The compared to post-monsoon (Fig. 4b), which could be due to evapora­
seasonal variation in isotope data is very small indicating less influence tive enrichment. This seasonal variation in isotope data of non-thermal
of current hydrological cycle on the reservoir source. The isotopic values waters suggests contribution of local precipitation to the shallow aquifer
of thermal waters are similar for Atri and Tarabalo sites with mean δ2H system. Unlike thermal waters, there is no site-wise clustering in non-
and δ18O values − 24.01 ± 2.2 ‰ and − 4.1 ± 0.3 ‰ respectively. At thermal waters excepting for Taptapani site (Fig. 4b). This observation
Athmalik site, the mean values of δ2H and δ18O are − 17.1 ± 1.05 ‰ and suggests that the thermal and non-thermal water circulations in EGMB
− 3.2 ± 0.17 ‰ respectively. Isotope data of thermal waters from Atri follow different flow paths. Furthermore, the non-thermal waters show a
and Tarabalo sites fall in a single cluster indicating possibility of a wider scatter in the EC correlations with δ2H and δ18O (Fig. S3) and do
common reservoir, however, thermal waters of Athmalik and Taptapani not show a particular grouping, like in the case of thermal waters. This
have different and distinct stable isotopic signatures inferring the pos­ clearly suggests the influence of soil-water interaction in the top
sibility of different source reservoirs (Fig. 4a). Considering the similarity weathered formation and contribution of precipitation to non-thermal
in basement rocks of EGMB, the discrepancy among the temperature waters. Depleted isotope composition of non-thermal waters of Tapta­
values, ionic ratios and isotopic compositions rules out the possibility of pani site (mean δ18O: − 6.4 ‰) suggests their origin from rainfall
a common reservoir feeding these thermal waters. In other words, this occurring at higher elevation. The DEM of this area indicates that the
confirms the presence of different reservoirs or/and flow paths for altitude of Taptapani site is ~450 m amsl (above mean sea level) while
thermal sprouts at these sites. In order to verify if the means of the other three sites have altitude in the range of 50–75 m amsl (Fig. 2b).
isotope data of these groups, viz., Atri & Tarabalo (n1 = 12) and Ath­ Therefore, it can be concluded that shallow aquifer system of Taptapani
malik (n2 = 10) are statistically significant, ANOVA single factor test site is recharged by precipitation occurring at higher elevations.
was conducted. It is found that the Fstat (60.6) > Fcritical (4.35) (p <

13
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

5.4. Age dating of thermal waters concentration of these thermal waters. However, DIC derived in the
sulphate reduction process normally has δ13C in the range of − 5 to − 10
Very low environmental tritium content observed in thermal waters ‰ (Hoefs, 1997), which again different from the range of δ13C –DIC
of Atri, Tarabalo and Athmalik sites clearly suggests that ages are very values observed in this case, therefore the influence of sulphate reduc­
old (>50 years) while at Taptapani, the thermal water is modern water. tion on the dating of thermal waters can be ignored.
Most of the non-thermal waters have tritium content >1.5 TU indicating The silicate mineral weathering reactions supply DIC to waters
modern recharge. A few non-thermal waters show low tritium (~1 TU), through dissolution of soil and atmospheric CO2. Recharging waters
which can be attributed to thermal water mixing. Radiocarbon dating dissolve root zone CO2 to form carbonic acid (log pCO2 − 2.6) facilitating
was carried out to estimate the age of thermal waters. The major chal­ silicate mineral weathering. In the case of closed condition, the
lenge in estimating accurate radiocarbon age of waters is assigning the consumed CO2 is not replenished by the soil gas and that results in
initial radiocarbon value for DIC. Several models are in use, which can lowering of log pCO2 values, as observed in this case (− 2.3 to − 3.9,
account for the different isotope and geochemical processes modifying Table S8). The log pCO2 also plays a crucial role in controlling the
the initial signatures of 14C of DIC (Clark and Fritz, 1997; Blaser et al., saturation state of carbonate minerals (Charlet et al., 1990) and indicate
2010 and references therein). The source of CO2 is important in the microbial oxidation of organic matter (Benedetti et al., 1996). The
geothermal waters since different processes provide different initial open/closed condition of thermal water with respect to CO2 is further
radiocarbon values for the DIC. The thermo-metamorphic decarbon­ examined from δ13C-DIC vs. HCO–3 plot (Fig. 10a). Atmospheric CO2 has
ation of carbonates and upper mantle degassing contribute to CO2, a typical δ13C value − 8 ‰ (Cerling et al., 1991) therefore the corre­
which has a 14C value of 0 pMC (Evans et al., 2008). Both these sources sponding δ13C-DIC would be +1.6 ‰ shown by line 1 in Fig. 10a. The
contribute to radiocarbon dilution and overestimation of thermal water closed system dissolution of carbonate rocks without any fractionation is
ages. However, the large difference between the δ13C –DIC values of the shown by line 2 (δ13C: 0 ‰). The non-equilibrium dissolution of car­
studied thermal waters (− 20 to − 16 ‰ VPDB) and the expected values bonates (1:1 mixture of soil CO2 with δ13C: − 25 ‰ and carbonates with
of decarbonation and degassing processes (− 2 to +2 ‰ and − 8 to − 3 ‰ δ13C: 0 ‰) is shown by line 3. The equilibration of DIC with soil CO2
respectively) indicate that radiocarbon dilution is not significant. This is (δ13C: − 25 ‰) yields − 17.5 ‰ value as shown by line 4. From the graph,
further supported by presence of low CO2 concentration (<0.2 %) in gas it can be observed that the sample data fall along the line 4 indicating
emissions from these sites (Mahala, 2019). The oxidation of organic equilibrium of DIC with soil CO2. Both thermal and non-thermal waters
material in the presence of dissolved sulphate within the aquifer also show a similar source of DIC. The radiocarbon content of the DIC also
complicates 14C dating due to addition of 14C free carbonate to the DIC modifies similar to δ13C-DIC. Under open system condition, the DIC
reservoir (Aravena et al., 1995). There is no data available on the H2S exchanges continuously with the 14C active soil gas; therefore, the initial

Fig. 10. Plots of a) δ13C-DIC vs. HCO–3, line (1) represents DIC derived from atmospheric CO2, line (2) – closed system dissolution of calcite and line (3) – non-
equilibrium dissolution of carbonates, line (4) – DIC under equilibration with soil CO2, b) log pCO2 vs. HCO–3 displaying the open/close condition of thermal and
non-thermal waters with respect to CO2, dashed line represents log pCO2 of recharging water, c) δ13C vs. 14C, d) pH vs. 14C illustrating the influence of deep carbon
sources. The arrows shown in (b), (c) and (d) are eye-guide lines.

14
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

activity of DIC (14CDIC) remains at 100 pMC. But, under closed equi­ modelled values provided confidence to the reliability of the ages
librium condition, the 14C-DIC decays with time without being replen­ determined. The residence time of thermal waters at Atri (T1: 14.0 – 14.9
ished. An increasing trend in log pCO2 is noticed in thermal waters from Ka) and Tarabalo (T1: 13.3–14.2 Ka) is found to be similar while thermal
− 3.9 to − 2.3, which reaches to the value similar to non-thermal water waters at Athmalik are relatively older (T1: 18.8 – 20.7 Ka). Thermal and
(Fig. 1b) indicating that the DIC is derived dominantly from soil CO2 and non-thermal mixed sources show ages in the range of 2.8 – 7.6 Ka. Non-
the decrease in log pCO2 clearly suggests a closed CO2 condition. Cor­ thermal water is shown by negative radiocarbon age indicating modern
relation between δ13C and 14C is found to be positive implying the recharge. The estimated ages of the thermal waters suggest that these
contribution of carbon sources with dead radiocarbon and depleted δ13C waters are recharged during Last Glacial Maximum (LGM), a relatively
(Fig. 10c). The dissolution of carbonates with time (mineralized organic arid climatic regime. This interpretation is supported by the enriched
matter) is further ascertained by the negative relationship between 14C δ18O values of the thermal waters (δ18O: − 3.2 and − 4.1 ‰ for Athmalik
and pH (Fig. 10d). and Atri/Tarabalo respectively) compared to the present-day meteoric
Radiocarbon correction using Pearson model was not applied in this waters (average rainwater δ18O: − 5.5 ‰, Keesari et al., 2016). We
study as the model demands recharge conditions of the aquifer to obtain interpret the enriched δ18O value of thermal waters to represent
accurate ages. Fontes and Garnier’s (F & G) model though does not recharge during the decreased summer monsoon rainfall, which is
require input parameters and removes the recharge conditions uncer­ characterized by enriched mean annual δ18O of the rainfall. High
tainty, it is applicable to only carbonate systems and therefore was not evaporative flux reflects a significant reduction in the soil moisture
considered. In closed system, like that typically exhibited by deep which facilitates greater fractionation and leads enriched δ18O compo­
groundwater, the IAEA model is apt (Clark and Fritz, 1997), as it takes in sition in recharging waters, as was observed in this case. The inter-
to account the temperature dependent isotope fractionation among tropical aridity in EGMB (Delang-Puri sector, Odisha) during 12 – 17
carbonate species (CO2-HCO–3–CO2− 3 ), which is highly effective for Ka has been reported by Kulkarni et al (1998), wherein the authors also
thermal waters. The relatively high pH (mean: 8.3) and lower log pCO2 noted that the isotopic composition of groundwater is enriched
(<-2.6) observed in thermal waters would not support an open CO2 compared to modern precipitation hinting at the possible recharge
evolutionary pathway, since the continuous supply of CO2 would render during arid climate.
acidic pH. The global sea level data indicate that the sea level during the period
The general formula for corrected 14C groundwater age is given as 20 Ka was 130 m below the current level (Lambeck et al., 2014). This
(equation (4)); could have displaced the shoreline to further east from the current
location. A similar phenomenon was reported in Bengal basin
a14 Cmeasured
Age = − 8267 × ln (4) (Bangladesh) where the shoreline during the LGM (20 Ka) was displaced
q × a14 Cinitial
100–150 km to the south (Ravenscroft et al., 2018). Considering the age
The correction factor q is calculated using the following equation (5): of the thermal waters and the sea level during that time, the sea water
component as the source of thermal water can be squarely ruled out. The
δ13 CDIC − δ13 Ccarb monsoonal circulation was greatly reduced during LGM and the resul­
q= (5)
δ Csoil − εgb − δ13 Ccarb tant precipitation was also isotopically enriched (δ18O: ~ − 3.5 ‰),
13

which matches well with the isotopic composition of thermal waters.


where εgb = − 9483
T + 23.89, which considers isotope exchange among The deep groundwater from Chittagong coastal aquifer (Bangladesh)
2–
carbon species (CO2 gas ↔ HCO–aq
3 ↔ CO3 solid). showed a radiocarbon age of 17 Ka age and an enriched stable isotope
where, δ13CDIC = measured 13C in groundwater, δ13Csoil = δ13C of the composition (Ravenscroft et al., 2018). The paleo-recharge history of
soil CO2, δ13Ccarb = δ13C of the calcite, Ɛgb = additive fractionation thermal waters is clearly displayed in the positive correlations among
factor between CO2 gas and HCO–3 and T is the temperature in kelvin. Temperature, δ18O and water age (Fig. 11). The trends observed in this
In order to account for the contribution of deep inorganic carbon figure clearly suggest that the observed enrichment in thermal waters is
sources and subsequent radiocarbon dilution, we have used the equation due to their recharge during arid climate. The deep groundwaters from
(6) proposed by Wang et al. (2019); Middle Ganga Plains are found to be 3.5 to 4.7 Ka old while in Deltaic
δ13 CDIC − δ13 Cdeadcarb Plains the typical ages range from 1 to 10 Ka (Lapworth et al., 2018) and
q= (6) 3–9 Ka (Hoque and Burgess, 2012). A comparison of the deep ground­
δ13 Crech − δ13 Cdeadcarb
water ages from Ganga Basin with the thermal waters of EGMB suggests
δ13Crech represents the recharge area; δ13Cdead carb represents the deep
source inorganic CO2. The δ13C value of deep source inorganic CO2 of
gas reservoirs and geothermal water is usually around − 1.5 ‰ (Wang
et al., 2019), which is selected as the value of δ13Cdead carb to correct the
14
C age of deep source “dead carbon”. The apparent ages were also
provided for comparison, which is estimated by considering 98.6 pMC
(highest 14C measured in the study area) as initial 14C value. The cor­
rected ages of thermal waters are shown in Table 1.
The results obtained from the 14C correction models slightly differ
from the apparent age estimates (uncorrected), indicating the impact of
temperature dependent fractionation on 14C value and impact of mixing
with deep inorganic carbon source (Table 1). The conformity between
the corrected and uncorrected radiocarbon age dating is largely due to
the absence of dissolution of the dead carbon from the aquifer matrix,
which is exemplified from the δ13C - DIC (-18.75 to − 21 ‰) values of the
thermal water. The δ13C values of DIC clearly reflect the silicate
weathering with soil CO2 derived from C3 type of plants that has an
average value of about − 19.1 ‰ (Das et al., 2005). The phenomenon of
silicate weathering is also substantiated from the subsurface geological
formation of the study area. However, the three modeled age estimates Fig. 11. Temperature – δ18O – water age correlations throwing light on the
are found to have similar range. The agreement among different recharge history of thermal waters.

15
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

longer circulation path of the thermal waters compared to the deep non- Singh et al., 2020). A conceptual hydrogeological-hydrochemical-
thermal groundwater systems. However, the thermal waters from the isotope model for thermal water genesis and circulations in Mahandi
adjacent Godavari river basin show a similar age, between 9 and 18.6 Ka graben, EGMB is presented in Fig. 12. From the above calculations, it can
(Chatterjee et al., 2019) which essentially infers that the evolution his­ be inferred that the geothermal reservoir in EGMB lies in the depth range
tory of both the geothermal systems is same. of 2.1 ± 0.35 km.

5.6. Future scope


5.5. Thermal water genesis, circulations and reservoir depth
Limberger et al. (2018) reported that the geothermal electrical pro­
There are many speculations on the source of heat in shear zones of duction is mostly confined to the areas containing high geothermal
EGMB since very little data is available. Field studies supplemented by gradients and sufficient reservoirs, which are very limited. Hence
satellite data also indicate that the thermal springs lie along lineaments development of technology focusing exploitation of low-enthalpy
(Fig. 1c, 2b), which infer close relationship between tectonics and geothermal systems would pave way for reduction in fossil burning
thermal activity. There is no evidence of any recent magmatic activity and leading to cleaner environment. Till now about 4000 thermal
around these thermal springs and hence the likely source of heat is the springs have been identified in Indian subcontinent and except few high
rocks of EGMB. The radioactive elements concentrations and the heat temperature thermal springs (i.e. Puga and Chumathang in Indus valley-
production (71.3–142.8 mW/m2) in the EGMB is reported to be highest Ladakh, Panamik in Nubra valley- Ladakh, Parbati and Beas valley in
amongst the granulite belts of India (Kumar et al., 2007). Very high Himachal Pradesh, Tapoban and Badrinath in Uttarakhand) all other
levels of radioactive elements (U, 600 mg/kg; Th 1500 mg/kg and K, thermal springs fall in the medium to low enthalpy category (Singh
5%) are reported in metamorphic rocks (Baranwal et al., 2006). Another et al., 2016). Indian shield has an estimated potential of 10,000 MW
study conducted on gas emissions of these thermal springs indicated a (www.indiaenergyportal.org) the majority of which are low to moderate
considerable amount of helium gas (0.27 to 1.7% by volume) and the N2- enthalpy systems. Extensive investigations have been carried out with
Ar-He ternary diagram suggested a long residence time of the gas phase regard to geological, geochemical and geophysical aspects of
in the crust affected by U and Th decay (Mahala, 2019). From the cor­ geothermal systems but the drilling investigations have been very
relations among the temperature and water age, it is observed that limited. Though it was realized a long time ago that India has good
temperature of thermal waters increases with residence time of water. potential to harness geothermal energy (Krishnaswamy and Rav­
From this observation and tectonic and thermal information provided ishanker, 1982), the development of the geothermal resources has not
above, it can be concluded that the thermal waters in EGMB are pro­ progressed mainly due to inadequate understanding of the deeper
duced due to the long contact and deep circulation of the rainwater thermal regime, leading to uncertainty over the proposed reservoir
inside the Precambrian metamorphites (high heat producing) models as well as the sustainability of heat source.
comprising elevated levels of radioactive elements. There are only a couple of geothermal energy-based power plants
The water flows in geothermal systems are complex due to the constructed/planned at a few places, such as, a 25 MW geothermal
occurrence of complex subsurface stratigraphy and heterogeneity in power project in the West Coast hot spring belt (Sharma and Trikha
geology, temperature gradients and fractures or other conduits in the 2013), a 35 MWe geothermal power plant in Gujarat (Chandrasekharam
permeable zones. The E-W trending graben bounding faults provide and Chandrasekhar, 2015; Singh et al., 2016), a 5 KW pilot power plant
permeable pathways for a thermal water circulation system along the at Manikaran (Razdan et al., 2008), an approved 25 MWe geothermal
northern sector of EGMB. Based on the hydrochemical, isotopic and power project in Andhra Pradesh (Chandrasekharam and Chan­
residence time values as well as variation of temperatures in thermal drasekhar, 2015). But with regard to power production, the geothermal
waters, two modes of thermal flow paths can be suggested in EGMB. The resources are entirely undeveloped in the country at present.
first mode is N-S flow system that originates from Iron Ore Super Group Reservoir temperature, its depth and volume are the primary factors
(IOSG) plateau and proceed to graben floor through fractures and fis­ considered while planning geothermal energy exploitation. India fares
sures to ultimately issue on the surface along the contact zones of IOSG well with regard to the information on the geothermal reservoir tem­
plateau and northern edge of Mahanadi graben. The second mode is S-N perature and its depth, but the information on genesis, recharge history
directional flow system that originates from Eastern Ghats Super Group and movement of thermal fluids is very limited. There are reports on the
(EGSG) plateau and reach graben floor, which subsequently appear as reduction of thermal water discharges and drying up of reservoirs after
thermal sprouts along the contact zones of IOSG plateau and southern the commencement of thermal water exploitation (Chandrajith et al.,
edge of Mahanadi graben. 2013; Pingheng et al., 2017), which clearly demonstrate the importance
The rainwater which upon penetrating downwards along the faults of understanding the genesis and estimating the flow patterns of the
and fractures is heated by geothermal gradient and then rises to the thermal fluids in any geothermal system. Therefore, in order to expand
surface through the permeable zones. The non-thermal water also cir­ the footprint of geothermal energy use in India, it is very important to
culates in the same way, but do not flow through the permeable fault determine the genesis, recharge history and thermal water circulations
zones and hence restricted to shallow depths. The circulation depth of of geothermal systems. In this regard, isotope techniques are very
thermal waters is estimated using the equation (7); valuable tools that can address uncertainty over the sustainability of
Tz − To geothermal reservoirs as demonstrated in this study.
Z= + Zo (7)
G
6. Conclusions
where Z is the circulation depth (m), To is the local annual average
temperature (26.6 ◦ C from https://en.climate-data.org); G is thermal The genesis and circulation of thermal and non-thermal waters in
gradient of the region (0.04 ± 0.01 ◦ C/m after Padhi and Pitale, 1995), shear zones of Mahanadi graben of Eastern Ghats Mobile Belt region are
and Zo is the thickness of the constant temperature zone (taken as 0 m). investigated using multi-tracers including chemical and isotope tracers.
Tz is the temperature (◦ C) of the geothermal reservoir calculated by a The thermal waters issue along the faults of Mahanadi graben and are
reasonable geothermometer (IMG in this case, Table 4). The estimated classified as low to medium enthalpy systems (55 – 59 ◦ C). Thermal
circulation depths for Atri, Tarabalo, Athmalik and Taptapani sites are waters close to graben are alkaline (pH = 8.75 – 9.13), less saline (EC
2.3 ± 0.6 km, 2.1 ± 0.5 km, 2.4 ± 0.6 km and 1.6 ± 0.4 km respectively <1500 µS/cm) and Na-Cl type while away from the graben thermal
(Table 4). Reservoir depths of similar order (1–3 km) have been reported water has low temperature (42 ◦ C), neutral pH and is Na-Ca-HCO3 type.
in several geothermal systems of India (Chatterjee et al., 2016, 2017; The characteristic ionic ratios and major ion trends indicate that bulk of

16
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

Fig. 12. Conceptual map depicting the genesis and circulation patterns of thermal waters in shear zones of Mahanadi graben, EGMB. Transect a-a’ (shown in Fig. 1c)
runs parallel to the coast from Singhbhum craton to EGB province.

the chemical ions is derived from the silicate weathering in both the Data availability
thermal and non-thermal waters.
Isotope characteristics of the thermal waters indicate meteoric The data is provided in the supplimenatary file available online
source and presence of three different thermal reservoirs. Stable isotope
and environmental tritium data indicate that the non-thermal waters are Acknowledgements
recharged by modern precipitation. Influence of evaporative enrichment
in non-thermal waters is observed during pre-monsoon period. The Authors thank Dr. H.J. Pant, Head, Isotope and Radiation Applica­
thermal circulations seem to originate from northern side of the graben tion Division (BARC) for the constant encouragement and support dur­
in the case of Athmalik site and southern side in the case of Atri/Tar­ ing the course of this work. Mr. Ajay Jaryal and Mr. S.N. Kamble (Isotope
abalo sites. The estimated reservoir temperatures range between 92 and Hydrology Section, IRAD, BARC) and personnel from GSI are duly
124 ◦ C and the circulation depths are found to be 1.6 – 2.4 km. The acknowledged for their help during field sampling. Authors would like
radiocarbon age of 13.3 to 20.7 Ka for thermal waters suggests dominant to thank Dr. Diana Sharma (NPDF, Isotope Hydrology Section, BARC) for
recharge during LGM, which is characterized by arid climate and low critically editing the manuscript and helping in map preparation.
rainfall. The stable isotope and hydrochemical inferences clearly
establish the improbability of marine sources while the age and tem­
Appendix A. Supplementary data
perature correlations indicate thermal gradient as the source of heat.
The emergence of thermal waters is controlled through fracture systems,
Supplementary data to this article can be found online at https://doi.
and the hydrogeochemical and isotopic compositions reflect minimum
org/10.1016/j.jhydrol.2022.128172.
mixing with non-thermal waters during their rise to surface. However,
thermal water located away from graben indicates significant mixing
with non-thermal water and hence lower outlet temperature and References
depleted isotopic content.
Aggarwal, J.K., Palmer, M.R., Bullen, T.D., Ragnarsdottir, K.V., Arnorsson, S., 2000. The
boron isotope systematics of Icelandic geothermal waters: 1. meteoric water charged
CRediT authorship contribution statement systems. Geochim. Cosmochim. Acta. 64, 580–585.
Aravena, R., Wassenaar, L.I., Plummer, L.N., 1995. Estimating 14C Groundwater Ages in
a Methanogenic Aquifer. Water Resour. Res. 31 (9), 2307–2317. https://doi.org/
Tirumalesh Keesari: Conceptualization, Methodology, Investiga­ 10.1029/95WR01271.
tion, Data curation, Writing – original draft. Sitangshu Chatterjee: Arnorsson, S., 1983. Chemical equilibria in Icelandic geothermal systems-implications
Conceptualization, Software, Visualization, Investigation, Validation. for chemical geothermometry investigations. Geothermics 12, 119–128.
Arnorsson, S., D́ Amore, F., Gerardo, J., 2000. Isotopic and chemical techniques in
Mukund Kumar: Methodology, Investigation. Hemant Mohokar: . geothermal exploration (ed. S. Arnórsson). Vienna, International Atomic Energy
Uday Kumar Sinha: Supervision, Resources. Annadasankar Roy: Agency, 351p.
Formal analysis, Investigation. Diksha Pant: Formal analysis, Investi­ Baranwal, V.C., Sharma, S.P., Sengupta, D., Sandilya, M.K., Bhaumik, B.K., Guinc, R.,
Saha, S.K., 2006. A new high background radiation area in the Geothermal region of
gation. Suraj D. Patbhaje: Methodology, Investigation. Eastern Ghats Mobile Belt (EGMB) of Orissa. India. Rad. Measure. 41, 602–610.
Benedetti, M.F., van Riemsdijk, W.H., Koopal, L.K., 1996. Humic substances considered
as a heterogeneous Donnan gel phase. Environ Sci Technol 30 (6), 1805–1813.
Declaration of Competing Interest Bhowmik, S.K., Dasgupta, S., Hoernes, S., Bhattacharya, P.K., 1995. Extremely high
temperature calcareous granulites from the Eastern Ghats, India: evidence for
The authors declare that they have no known competing financial isobaric cooling, fluid buffering and terminal channelized fluid flow. Europ. J. Min.
7, 689–703.
interests or personal relationships that could have appeared to influence Blaser, P.C., Coetsiers, M., Aeschbach-Hertig, W., Kipfer, R., Van Camp, M., Loosli, H.H.,
the work reported in this paper. Walraevens, K., 2010. A new groundwater radiocarbon correction approach

17
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

accounting for palaeoclimate conditions during recharge and hydrochemical Keesari, T., Sinha, U.K., Deodhar, A., Krishna, S.H., Ansari, A., Mohokar, H.V., Dash, A.,
evolution: The Ledo-Paniselian Aquifer. Belgium. Appl. Geochem. 25, 437–455. 2016. Occurrence of fluoride in groundwater from an industrial area in Odisha State,
Cerling, T.E., Solomon, D.K., Quade, J., Bowman, J.R., 1991. On the isotopic composition India – A geochemical perspective. Environ Earth Sci. 75, 1090. https://doi.org/
of carbon in soil carbon dioxide. Geochem. Cosmochim. Acta. 55, 3403–3405. 10.1007/s12665-016-5874-0.
Chandrajith, R., Barth, J.A.C., Subasinghe, N.D., Merten, D., Dissanayake, C.B., 2013. Keesari, T., Chatterjee, S., Pant, D., Singh, M., Sakhare, V., Sinha, U.K., Mohokar, H.,
Geochemical and isotope characterization of geothermal spring waters in Sri Lanka: Jaryal, A., Roy, A., Maitra, A., 2019. Dating of hot springs at Atri, Tarabalo and
Evidence for steeper than expected geothermal gradients. J. Hydrol. 476, 360–369. Athmalik sites in Odisha, India using radiocarbon technique. J. Radioanal. Nucl.
Chandrasekhar, T., Minissale, A., Vaselli, O., Chandrasekharam, D., D., Singh, H.K.,, Chem. https://doi.org/10.1007/s10967-019-06867-1.
2018. Understanding the evolution of thermal fluids along the western continental Krishnaswamy, V.S., Ravishanker,, 1982. Scope of development, exploration and
margin of India using geochemical and boron isotope signatures. Geothermics. preliminary assessment of geothermal resources potential of India. Records GSI, VIII
https://doi.org/10.1016/j.geothermics.2018.03.007. 2, 17–40.
Chandrasekharam, D., Chandrasekhar, V. 2010. Energy independence through CDM Kulkarni, K. M., Navada, S. V., Nair, A.R., Rao, S. M., Shivanna, K., Sinha, U. K., Sharma,
using geothermal resources: Indian scenario. Proceedings of the World Geothermal S. 1998. Drinking water salinity problem in coastal Orissa, India: Identification of
Congress 2010, Bali, Indonesia, 5 pp. past transgressions of sea water by isotope investigation. IAEA: N. 1998. Web. 293-
Chandrasekharam, D., Chandrasekhar, V. 2015. Geothermal energy resources, India: 306.
country update. World Geothermal Congress. 2015. Kumar, B., Rai, S.P., Kumar, U.S., Verma, S.K., Pankaj, G., Vijaya Kumar, S.V., Rahul
Charlet, L., Wersin, P., Srumm, W., 1990. Surface charge of MnCO3 and FeCO3. Geochim Jaiswal, B.K., Purendra, S.R., Kumar, N.G., 2010. Isotopic characteristics of Indian
Cosmochim Acta 54 (8), 2329–2336. precipitation. Water Resour. Res. 46, W12548. https://doi.org/10.1029/
Chatterjee, S., Sinha, U.K., Biswal, B.P., Jaryal, A., Jain, P.K., Patbhaje, S., Dash, A., 2009WR008532.
2019. An Integrated Isotope-Geochemical Approach to Characterize a Medium Kumar, P.S., Menon, R., Reddy, G., 2007. The role of radiogenic heat production in the
Enthalpy Geothermal System in India. Aqua. Geochem. 25, 63–89. https://doi.org/ thermal evolution of a Proterozoic granulite-facies orogenic belt: Eastern Ghats.
10.1007/s10498-019-09352-z. Indian Shield. Earth Planet. Science. Lett. 254 (1), 39–54.
Chatterjee, S., Ansari, M.A., Deodhar, A.S., Sinha, U.K., Dash, A., 2017. A multi-isotope Kundu, N., Panigrahi, M., Sharma, S., Tripathy, S., 2002. Delineation of fluoride
approach (O, H, C, S, B and Sr) to understand the source of water and solutes in some contaminated groundwater around a hot spring in Nayagarh, Orissa, India using
the thermal springs from West Coast geothermal area. India. Arab. J. Geosci. 10, 242. geochemical and resistivity studies. Environ. Geol. 43, 228–235.
https://doi.org/10.1007/s12517-017-3022-0. Lambeck, K., Rouby, H., Purcell, A., Sun, Y., Sambridge, M. 2014. Sea level and global ice
Chatterjee, S., Sharma, S., Ansari, M.A., Deodhar, A.S., Low, U., Sinha, U.K., Dash, A., volumes from the Last Glacial Maximum to the Holocene. PNAS October. 28. 2014.
2016. Characterization of subsurface processes and estimation of reservoir 111 (43). 15296-15303.
temperature in Tural and Rajwadi geothermal fields, Maharashtra. India. Geotherm. Lapworth, D.J., Zahid, A., Taylor, R.G., Burgess, W.G., Shamsudduha, M., Ahmed, K.M.,
59, 77–89. https://doi.org/10.1016/j.geothermics.2015.10.011. Mukherjee, A., Gooddy, D.C., Chatterjee, D., MacDonald, A.M., 2018. Security of
Chatterjee, S., Sinha, U.K., Ansari, M.A., Mohokar, H.V., Dash, A., 2018. Application of Deep Groundwater in the Coastal Bengal Basin Revealed by Tracers. Geophysical
lumped parameter model to estimate mean transit time (MTT) of the thermal water Research Letters 45 (16), 8241–8252. https://doi.org/10.1029/2018GL078640.
using environmental tracer (3H): Insight from Uttarakhand geothermal area (India). Limberger, J., Boxem, T., Pluymaekers, M., Bruhn, D., Manzella, A., Calcagno, P.,
Appl. Geochem. 94, 1–10. https://doi.org/10.1016/j.apgeochem.2018.04.013. Beekman, F., Cloetingh, S., van Wees, J.-D., 2018. Geothermal energy in deep
Chetty, T.R.K., 2001. Eastern Ghats Mobile Belt, India: a collage of juxtaposed terranes aquifers: A global assessment of the resource base for direct heat utilization.
(?). Gond. Res. 4, 319–328. Renewable and Sustainable Energy Reviews 82 (1), 961–975.
Clark, I.D., Fritz, P., 1997. Environmental Isotopes in Hydrogeology. Lewis Publishers, Mahala, S.C. 2019. Geology, Chemistry and Genesis of Thermal Springs of Odisha, India,
New York. Springer Briefs in Earth Sciences, https://doi.org/10.1007/978-3-319-90002-5.
Coplen, T.B., 1996. New guidelines for reporting stable hydrogen, carbon and oxygen Maitra, A., Chatterjee, A., Keesari, T., Gupta, S., 2020. Forming topography in granulite
isotope-ratio data. Geochim Cosmochim. Acta 60, 3359–3360. terrains: the (un) importance of chemical weathering. J. Earth Syst. Sci. 129, 17.
Craig, J., Absar, A., Bhat, G., Cadel, G., Hafiz, M., Hakhoo, N., Kashkari, R., Moore, J., https://doi.org/10.1007/s12040-019-1293-4.
Ricchiuto, T.E., Thurow, J., Thussu, B., 2013. Hot springs and the geothermal energy Marty, B., Jambon, A., 1987. C/He in volatile fluxes from the solid Earth: Implications for
potential of Jammu & Kashmir State, N.W. Himalaya. India. Earth-Science Rev. 126, carbon geodynamics. Earth Planet. Sci. Lett. 83, 16–26.
156–177. Minissale, A., Vaselli, O., Chandrasekharam, D., Magro, G., Tassi, F., Casiglia, A., 2000.
Dansgaard, W., 1964. Stable isotopes in precipitation. Tellus 16, 436–468. Origin and evolution of ‘intra cratonic’ thermal fluids from central-western
Das, A., Krishnaswami, S., Bhattacharya, S., 2005. Carbon isotope ratio of dissolved peninsular India. Earth Planet. Sci. Lett. 181 (3), 377–394.
inorganic carbon (DIC) in rivers draining the Deccan Traps, India: sources of DIC and Morgenstern, U., Taylor, C.B., 2009. Ultra low-level tritium measurement using
their magnitudes. Earth Plan. Sci. Lett. 236, 419–429. https://doi.org/10.1016/j. electrolytic enrichment and LSC. Isotopes Environ Health Stud. 45 (2), 96–117.
epsl.2005.05.009. https://doi.org/10.1080/10256010902931194.
Dasgupta, S., 1993. Contrasting mineral parageneses in high temperature calc-silicate Nair, A.R., Joseph, T.B., Rao, S.M., 1990. Measurement of low-level tritium in water by
granulites: examples from the Eastern Ghats. India. J. Metamor. Geo. 11, 193–202. electrolytic enrichment and liquid scintillation counting. Bulletin of Radiation.
Dasgupta, S., Bose, S., Das, K., 2013. Tectonic evolution of the Eastern Ghats Belt. Navada, S.V., Rao, S.M., 1991. Isotope studies of some geothermal waters in India.
Precam. Res. 227, 247–258. Isotopenpraxis. 27, 153–163.
Evans, M.J., Derry, L.A., France-Lanord, C., 2008. Degassing of metamorphic carbon Navada, S.V., Nair, A.R., Sharma, S., Kulkarni, U.P. 1995. Geochemical and isotope
dioxide from the Nepal Himalaya. Geochem. Geophys. Geosyst. 9 (4) https://doi. studies of the geothermal areas of Central and Northern India. In: IAEA (Ed.), Isotope
org/10.1029/2007GC001796. and Geochemical Techniques Applied to Geothermal Investigation. IAEA, Vienna.
Exley, R.A., Mattey, D.P., Clague, D.A., Pillinger, C.T., 1986. Carbon isotope systematics 63–82.
of a mantle “hotspot”: a comparison of Loihi Seamount and MORB glasses. Earth Nicholson, K., 1993. Geothermal Fluids Chemistry and Exploration Techniques. Springer-
Planet. Sci. Lett. 78, 189–199. Verlag 1–215.
Fournier, R.O., 1977. Chemical geothermometers and mixing models for geothermal Padhi, R.N., Pitale, U.L., 1995. Potential geothermal fields in the context of national
systems. Geothermics 5, 40–41. scenario on non-conventional energy resources development programme in India.
Fournier, R.O., 1979. A revised equation for Na/K geothermometer. Geotherm. Proc. WGC 1995, 525–530.
Resources Council Trans. 3, 221–224. Pang, Z.H., Reed, M.H., 1998. Theoretical chemical thermometry on geothermal waters:
Fournier, R.O., Truesdell, A.H., 1970. Chemical indicators of sub-surface temperatures problems and methods. Geochim Cosmochim Acta 62, 1083–1091.
applied to hot spring waters of Yellowstone national park, USA. Proc. UN Symp. On Pingheng, Y., Qun, C., Shiyou, X., Jianli, W., Longran, C., Qin, Y., Zhaojun, Z., Feng, C.,
development and utilization of geothermal resources. Pissa, 2 (1). Geothermics Spl 2017. Hydrogeochemistry and geothermometry of deep thermal water in the
Issue. 2, 529–535. carbonate formation in the main urban area of Chongqing. China. J. Hydrol. 549,
Fournier, R.O., Truesdell, A.H., 1973. An empirical Na-K-Ca geothermometer from 50–61.
natural waters. Geochim. Cosmochim. Acta 37, 1255–1275. Ramakrishnan, M., 1987. Stratigraphy, sedimentary environment and evolution of the
Giggenbach, W.F., 1988. Geothermal solute equillibria. Derivation of Na-K-Mg-K late Proterozoic Indravati basin, Central India. Mem. Geol. Soc. India. 6, 139–160.
geoindicators. Geochim. Cosmochim. Acta 52, 2749–2765. Ravenscroft, P., McArthur, J.M., Rahman, M.S., 2018. Identifying multiple deep aquifers
Giggenbach, W.F., 1992. Isotope shift in waters from geothermal and volcanic systems in the Bengal Basin: Implications for resource management. Hydrol. Proc. 32,
along convergent plate boundaries and their origin. Earth and Planetary Sci. Lett. 3615–3632. https://doi.org/10.1002/hyp.13267.
113, 495–510. Razdan, P.N., Agarwal, R.K., Singh, R., 2008. Geothermal Energy Resources and its
Gsi.,, 1987. Records of the geological survey of India. Geological Survey of India, Potential in India. e-J. Earth Scien. Ind I I, 30–42. http://www.earthscienceindia.in
Calcutta. fo/.
Gsi.,, 1991. Geothermal atlas of India. Geological Survey of India, Calcutta. Roy, S., Mareschal, J.C., 2011. Constraints on the deep thermal structure of the Dharwar
Hoefs, J., 1997. Stable Isotope Geochemistry, Fourth Edition. New York, Springer-Verlag. craton, India, from heat flow, shear wave velocities, and mantle xenoliths.
Hoque, M.A., Burgess, W.G., 2012. 14C dating of deep groundwater in the Bengal Aquifer J. Geophys. Res. 116 (B02409) https://doi.org/10.1029/2010JB007796.
System, Bangladesh: Implications for aquifer anisotropy, recharge sources and Sammel, E.A., Craig, R.W., 1981. The geothermal hydrology of Warner valley, Oregon: a
sustainability. Journal of Hydrology 444–445 (2012), 209–220. reconnaissance study. US Geological Survey Professional Paper. 1044 (1), 47.
Hounslow, A.W., 1995. Water quality data- Analysis and Interpretation. Lewis Santhosh, M., 2012. Geological Society. London, Special Publications. 365, 263–288.
Publishers, CRC Press, Florida, USA. https://doi.org/10.1144/SP365.14.
Keenan, J.H., Keyes, F.G., Hill, P.G., Moore, J.G., 1969. Steam Tables – Thermodynamic Sarolkar P.B. 1996. Characteristics of Thermal water from shallow reservoir at Tattapani
Properties of Water Including Vapour, Liquid and Solid Phases (international edition geothermal field, Dist. Surguja, Madhya Pradesh, India, Geothermal Energy in India,
– metric units). Wiley, New York, p. 162. Eds. U.L. Pitale and R.N. Padhi, GSI. Spl. Pub. No. 45. 1996. 75-82.

18
T. Keesari et al. Journal of Hydrology 612 (2022) 128172

Sharma, O.P., Trikha, P., 2013. Geothermal energy and its potential in India. Renewable Ladakh and Himachal (India): Evidence for CO2discharge in northwest Himalaya.
Energy. AkshayaUrja. 7 (1), 14–18. Geothermics. 64, 314–330.
Singh, H.K., Chandrasekharam, D., Trupti, G., Mohite, P., Singh, B., Varun, C., Sinha, S. Valdiya, K.S., 2016. The making of India. Geodynamic evolution, (2nd ed.),. Springer,
K., 2016. Potential Geothermal Energy Resources of India: A Review. Curr. p. 923 pp..
Sustainable Renewable Energy Rep. https://doi.org/10.1007/s40518-016-0054-0. Wang, H., Mao, X.M., Wang, T., Feng, L., Liang, L., Zhu, D., Yang, K., 2019.
Singh, H.K., Satish, K.S., Mohammad, A.A., Dornadula, C., 2020. Tracing the evolution of Hydrogeochemical characteristics of hot springs exposed from fault zones in western
thermal springs in the Hazaribagh area of Eastern Peninsular India through Guangdong and their 14C age correction. J. of Groundwater Sci. and Eng. 7 (1),
hydrogeochemical and isotopic analyses. Geotherm. 85, 101817 https://doi.org/ 1–14. https://doi.org/10.19637/j.cnki.2305-7068.2019.01.001).
10.1016/j.geothermics.2020.101817. Wright, J.D., Miller, K.G., Fairbanks, R.G., 1991. Evolution of modern deepwater
Spycher, N., Peiffer, L., Sonnenthal, E.L., Saldi, G., Reed, M.H., Kennedy, B.M., 2014. circulation: evidence from the late Miocene Southern Ocean. Paleoceanography. 6,
Integrated multicomponent solute geothermometry. Geothermics. 51, 113–123. 275–290. https://doi.org/10.1029/90PA02498.
Thussu, J.L., 2002. Geothermal energy resources of India. Geol. Surv. India. Spec. Publ. Zimik, H.V., Farooq, S.H., Prustry, P., 2017. Geochemical evaluation of thermal springs
69, 210. in Odisha. India. Environ. Earth Sci. 76, 593. https://doi.org/10.1007/s12665-017-
Tiwari, S.K., Santosh, K.R., Sukesh, K.B., Anil, K.G., Manju, N., 2016. Stable isotopes 6925-x.
(δ13C-DIC, δD, δ18O) and geochemical characteristics of geothermal springs of

19

You might also like