Very Important CI

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Applied Geochemistry 113 (2020) 104492

Contents lists available at ScienceDirect

Applied Geochemistry
journal homepage: http://www.elsevier.com/locate/apgeochem

Geothermometry and geochemistry of groundwater in the Continental


Intercalaire aquifer, southeastern Algeria: Insights from cations, silica and
SO4–H2O isotope geothermometers
Adel Abdelali a, *, Imed Eddine Nezli a, Rabah Kechiched b, Said Attalah c,
Slimane Abdeldjabbar Benhamida a, d, Zhonghe Pang e
a
Laboratoire de G�eologie du Sahara, Universit�e Kasdi Merbah Ouargla, 30000, Algeria
b
Laboratoire des R�eservoirs Souterrains: P�etroliers, Gaziers et Aquif�eres, Universit�e Kasdi Merbah Ouargla, 30000, Algeria
c
School of Plant Sciences, University of Arizona, Tucson, AZ, USA
d
Agence Nationale des Ressources Hydriques (ANRH), Direction Sud-Ouargla, 30000, Algeria
e
Institute of Geology & Geophysics, Chinese Academy of Sciences, No. 19 Beitucheng West Road, Chaoyang District, Beijing, 100029, PR China

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

Editorial handling by H. Armannsson Characterization of geochemistry and geothermometry of the groundwater from the Continental Intercalaire (CI)
aquifer, one of the largest aquifers in the world, stretching over one million km2 surface area shared between
Keywords: Algeria, Tunisia and Libya, was conducted in southeastern Algeria nearby the border with Tunisia. Thirty-two
Continental intercalaire water samples were collected from boreholes to analyze the physicochemical parameters and stable isotopes
Isotope geochemistry
(δ18O(H2O), δ18O(SO4), δ 34S(SO4)) and determine the origin of water mineralization of the CI aquifer. Water
Geothermometry
temperature was assessed using several geothermometers, such as cations, silica and SO4–H2O stable isotopes.
Geothermal systems
Isotope geothermometry The CI aquifer displayed a discharge temperature varying from 45 to 65.1 � C due to water circulation at deeper
Mineral equilibrium depths ranging from 1000 to 2200 m. The Results show that CI water contains high total dissolved solids (TDS)
and neutral to slightly alkaline pH. The most frequent water types were Na–Cl–SO4 and Ca–SO4 indicating the
geological formation nature of CI aquifer which is mainly composed of evaporites. The application of Na–K and
Na–K–Ca geothermometers yielded unreliable temperatures. However, Na–Li geothermometer resulted in rela­
tively plausible temperatures ranging from 74.69 to 147.83 � C. Quartz and CaSO4–H2O isotope geothermometer
are the most suitable methods for temperature estimation, due to their attainment of equilibrium resulting in
temperatures ranging from 62 to 93 � C, and validation by the multiple mineral equilibrium approach. The
present study demonstrated a successful application of CaSO4–H2O isotope geothermometer in a low enthalpy
(<100 � C) geothermal system. The results corroborate previous findings on the same CI aquifer at the Tunisian
side and on carbonate-evaporite geothermal reservoirs. Furthermore, geothermal potential of the CI aquifer has
been highlighted suggesting its use as a source of renewable energy which could be applied in heating green­
houses and/or generating electricity.

1. Introduction times geothermal energy has become a valuable source for industry and
agriculture. The application of geothermal energy differs depending on
Geothermal sources are considered as a renewable energy potential the reservoir temperature. For instance, a geothermal reservoir with a
which originated from heat inside the earth globe. This energy is usually temperature of 80 � C could generate electricity, while with a tempera­
linked to several geological formations that contain water resources ture of 30 � C, it would rather be useful for heating greenhouses,
either with their geothermal gradient at high depth or with their warming soil, or improving fish farming (Lindal, 1973).
occurrence nearby heat sources such as magma. In ancient times, the use Deep aquifers may also represent a considerable source of
of geothermal springs was mostly limited to bathing, whereas in modern geothermal energy such as many underground aquifers in Algeria.

* Corresponding author.
E-mail addresses: adelabdelali5@gmail.com (A. Abdelali), imedinezli@yahoo.fr (I.E. Nezli), rabeh21@yahoo.fr (R. Kechiched), sattalah@email.arizona.edu
(S. Attalah), slimbenha@gmail.com (S.A. Benhamida), z.pang@mail.iggcas.ac.cn (Z. Pang).

https://doi.org/10.1016/j.apgeochem.2019.104492
Received 5 April 2019; Received in revised form 7 December 2019; Accepted 9 December 2019
Available online 12 December 2019
0883-2927/© 2019 Elsevier Ltd. All rights reserved.
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

Occurrence of thermal water from aquifers in northeastern Sahara in aquifer system.


Algeria, particularly in Ouargla and Touggourt areas, was formerly re­ The CT aquifer is a confined reservoir composed of several layers
ported (Saibi, 2009; Chaib and Kherici, 2014). This thermal water is from the upper Senonian to Miocene, which can be captured at a depth
used for bathing, heating greenhouses and domestic usage purposes. ranging from 100 to 600 m from the soil surface (Bel and Cuche, 1969;
Recently, a fish farming project has been successfully implemented UNESCO, 1972). The structure of the CT aquifer is characterized by two
using thermal water from The Albian CI aquifer in Ouargla, Algeria geological units: 1) Senonian at the bottom with mainly limestone, and
(Saibi, 2009). Although hydrogeological and hydrochemical aspects of 2) Miopliocene at the top with sands, clays and sandstones. These two
the Albian CI aquifer, which is recognized as one of the largest aquifers geological units are separated by an impermeable layer of marl and clays
in the world, has been largely studied (Guendouz and Michelot, 2006; from the Eocene epoch. According to a previous piezometric map
Chkir et al., 2009; Moulla et al., 2012; Petersen et al., 2013, 2018; (UNESCO, 1972; OSS, 2003), the main flow lines in the CT aquifer begin
Petersen, 2014; Gonçalv�es et al., 2015; Slimani et al., 2017), more de­ at the recharge areas in the south of Tinrhert basin, southern Algeria,
tails on the geothermal characterization need to be investigated, espe­ and end at the north eastern discharge of Chott Merouane, Chott Melghir
cially at a regional scale including several locations in the northeastern and Chott Djerid in Tunisia.
Sahara of Algeria. A natural extension of the CI aquifer within The CI aquifer, which is the main focus of the present study, lies
Algeria-Tunisia southern border has been considered as a geothermal under the CT aquifer in a confined reservoir containing lower cretaceous
reservoir (Ben Dhia and Meddeb, 1990; Kamel, 2012; Makni et al., geological formations dating back to Neocomian, Barremian, Aptian and
2013). Albian epochs (Cornet, 1964; Busson, 1972; Bishop, 1975; Fabre, 1976).
The CI aquifer where the present study was conducted is enclosed in The CI aquifer extends from the Saharan Atlas in the north to the Tassili
a stable, non volcanic area with a typical geothermal gradient of 2–3 � C/ mountains in the south, and from the Guir-Saoura valley in the west to
100m (Takherist and Lesquer, 1989) and a depth ranging from 1000 to the Libyan Desert in the east (Cornet, 1964; Busson, 1972; Fabre, 1976;
2200 m (Edmunds et al., 2003), thus making it a focus of attention as a Edmunds et al., 2003; Ould Baba Sy, 2005; Guendouz, 1985). It has been
geothermal reservoir. suggested that the CI aquifer contains two distinct sub-basins separated
Assessment of the subsurface temperature is usually challenging by the M’zab dorsal: 1) the Grand Erg Oriental and 2) the Grand Erg
because of limitations in direct measurement methods. For instance, Occidental (Cornet, 1964; UNESCO, 1972; Bishop, 1975; Guendouz,
bottom hole surveys provide accurate results, however; this direct 1985). A homogeneous layer of Cenomanian clays and evaporites
method is highly costly due to the drilling requirements. Thus, alter­ overlies the oriental basin of the CI aquifer along the area where the
native indirect methods such as geochemistry are used. These economic present study was conducted, displaying a large variability in terms of
alternatives use geothermometers that have been developed for the lithology, thickness, and depth due to its wide geographic extension. The
main purpose of estimating reservoir temperatures using water chemical CI aquifer is mainly composed of sediments from the Mesozoic era, with
or isotope composition (D’Amore and Arno �rsson, 2000). These tech­ a depth varying considerably from South to North exceeding 2000 m in
niques are commonly applied to volcanic geothermal systems under Chott Melghir and Chott Djerid, in Tunisia. The flow lines converge from
various conditions (Fournier, 1977), whereas with low temperature and the Saharan Atlas, the M’zab dorsal and the Dahar region in Tunisia
non-volcanic systems more caution is required (Kharaka and Mariner, towards the discharge areas in Chott Djerid, Chott Fedjedj and Gulf of
1989). Although geothermometry of carbonate-evaporite geothermal Gabes (UNESCO, 1972).
reservoirs has been extensively studied (Kharaka and Mariner, 1989; Apparently, there is a significant lack of tectonic deformations in the
Ben Dhia and Meddeb, 1990; Lo �pez-Chicano et al., 2001; Mohammadi studied area (Edmunds et al., 2003) except in the southern zones where
et al., 2010; Makni et al., 2013; Wang et al., 2015; Pasvanog �lu, 2015; an intense faulting such as Amguid and Elhamma faults, and tectonic
Belhai et al., 2016; Yang et al., 2017; Blasco et al., 2017, 2018), geo­ activity might occur. A previous study in the same area reported some
thermometers are neither designed for nor applied to such environment modifications in lithology, piezometry and water chemical composition
due to erroneous results they might provide. A previous study on the CI (Edmunds et al., 1997).
aquifer in Ouargla, southern Algeria, using some geothermometry A geological characterization has been conducted locally based on
compounds such as Na–K, Na–K–Ca and K–Mg (Abdelali et al., 2017) the available data from the studied wells. Lithological logs were oriented
showed promising results, especially with K–Mg geothermometer. SSW-NNE from Ouargla to Meghaier to track the evolution of lithologies
The goal of the present study was to highlight the importance of and the thickness of aquifers (Figs. 1 and 2). Logs show similar lithol­
geothermometry and geochemistry application to the CI aquifer using ogies over the entire studied area, whereas the thickness and depth vary
several geothermometers to cover a larger study area southern Algeria significantly. The thickness increases towards NNE direction in both CT
including Touggourt, Djama ^a, Meghaier and El Oued. Moreover, and CI aquifers. For instance, the Senonian-Eocene reservoir in the CT
recently drilled wells in CI aquifer that had not been studied previously aquifer displays a thickness of 215 m at HDB(26) in the south western
(Edmunds et al., 2003) have been herein included. The first objective part then increases to 770 m at Ain sebbala (22) in the north eastern part
was to conduct a geochemical study to identify the chemistry and the (Fig. 2). The wells drilled in CI aquifer reach different geological for­
origin of water mineralization, thus the type of water that CI aquifer mations depending on each drilling location. For instance, in Ouargla, CI
encloses. The second objective was to apply geothermometry to assess groundwater is withdrawn from sandstones that contain clayey material
the CI reservoir temperatures based on chemical and isotope geo­ dating back to Albian epoch, whereas in the northeastern area the li­
thermometers including δ18O (SO4–H2O) isotope, and mineral satura­ thology of the CI aquifer shows mainly a Barremian formation. More­
tion index. over, the depth of the wells varies from one location to another with the
deepest point of 1900 m reaching the top of CI aquifer at Ain Sebbala
2. Geological and hydrogeological context (22) (Fig. 2).
Illustrated correlation between the studied wells in the direction
The North Western Sahara Aquifer System (NWSAS) is one of the SSW-NNE (Fig. 2) shows a variation in lithology with respect to the
largest aquifer systems in the world with more than one million km2 increasing depth towards the northeast direction where detrital sedi­
surface area (UNESCO, 1972) stretching over large areas of three ments are separated by impermeable substratum of clays and anhydrite.
countries: Algeria, Tunisia and Libya. In addition to the quaternary The impermeable layer overlying CI bed contains Cenomanian com­
water table, which fluctuates approximately at 2 m depth from the soil posites including clays (in Ouragla area), limestone, dolomite and
surface, the NWSAS includes two overlaying large aquifers referred to as anhydrite (in Temacine area) with a significant thickness ranging from
1) the Complex Terminal (CT) aquifer; and 2) the Continental Interca­ 200 to 250 m which separates the CI aquifer from the CT aquifer (i.e.,
laire (CI) aquifer, respectively from the soil surface to the bottom of the Senonian-Eocene formation). Lithology of sandstones and sandy clays in

2
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

Fig. 1. Geological map of the studied area, modified after Busson (1972) including the locations of sampling wells denoted by red spots, numbers from 1 to 14; 16 to
30, and letters M3, G2 and Kh1. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

the CI aquifer showed composites from Albian, Aptian and Barremian HACH). Sodium (Naþ), potassium (Kþ) and lithium (Liþ) concentrations
formations which are located deeper than 1000 m in the lower beds. The were measured with a flame spectrophotometer (PFP 7, JENWAY) in the
Albian and Barremian, which are separated by a thin layer ranging from laboratory of Sahara Geology at the University of Ouargla, Algeria.
10 to 30 m of Aptian dolomite, contain mostly clastic formations of Sulfate (SO24 ), chlorides (Cl ), nitrates (NO3 ) and silicon (Si) were
sands, sandstones and clays with infrequent limestone and marl which measured using a spectrophotometer device (DR2000, HACH). Bi­
are rather abundant in the Varaconian formation in the extreme carbonates (HCO3 ) were determined by titration. Magnesium (Mg2þ)
northern side of the aquifer (Fabre, 1976). and calcium (Ca2þ) were determined by complexometric titration using
Ethylenediaminetetraacetic acid (EDTA) protocol (Mg2þ ¼ TH-Ca2þ) in
3. Materials and methods the laboratory of National Agency of Hydraulic Resources (ANRH) in
Adrar, Algeria. The percentage ion-balance of the studied samples was
3.1. Analytical techniques calculated by PHREEQC v.3 software (Parkhurst and Appelo, 2013)
following previously described standard methods (Clesceri et al., 1999).
From January 25th to February 14th, 2017, thirty-two water samples All samples displayed more or less than 5% charge imbalances indi­
were collected from wells drilled in CI aquifer at different locations cating their reliability, except Kh1 sample which showed more or less
denoted as follows: Ouargla (23, 24, 25, 26, Kh1, G2 and M3), Toug­ than 6%. Additionally, Sixteen water samples (2, 7, 9, 11, 12, 14, 16, 18,
gourt (1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11), Djam^
aa (12, 13, 14, 16, 17, 18), 19, 22, 23, 24, 26, G2, Kh1, M3) were collected from the study area on
Meghaier (19, 20, 21, 22) and El Oued (27, 28, 29, 30). Chemical February 2018 to conduct the analysis of lithium (Liþ), silicon (Si) and
characterization of the water samples was performed considering pre­ aluminum (Al3þ) using ICP-MS instrument (Agilent 7700, ICP-MS) in
viously established standards (Rodier, 2016). Some physico-chemical ALEC laboratory of the University of Arizona, USA, and the analysis of
parameters such as pH, Temperature and Electrical conductivity (EC) isotopic ratios of δ18O(H2O), δ18O(SO4), δ34S(SO4), in the environmental
were measured in the field using a Multi-parameter device (HQ40D, isotope laboratory at the University of Arizona, USA. The isotope

3
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

Fig. 2. SSW-NNE stratigraphic correlation between the studied locations based on drilling data. (1): sandy shale; (2): limestone, Dolomite, marl and anhydrite; (3):
limestone, dolomite and marl; (4): anhydrite, massive salt at the bottom; (5): marly limestone and dolomite; (6): dolomitic limestone, limestone, shale and anhydrite;
(7): calcareous marl and dolomite; (8): argillaceous sandstone; (9): dolomite and limestone; (10): argillaceous sandstone. Depth in meters (m).

δ18O(H2O) was measured with a gas-source isotope ratio mass spec­ (1) Water-rock equilibrium, a temperature dependent reaction which
trometer (Finnigan Delta S), then compared to V-SMOW (Vienna Stan­ happens deep in the aquifer.
dard Mean Ocean Water) standard (Hoefs, 2018). Samples were (2) Required minerals for the reaction to happen should be a part of
equilibrated with CO2 gas in an automated equilibration device coupled the aquifer.
to the mass spectrometer. The analysis of isotopes δ18O(SO4) and (3) Any chemical/isotopic composition change in the water during
δ34S(SO4) was performed by precipitation of dissolved sulfate as BaSO4 the ascension process indicates a shallower equilibrium or a
powder by adding a BaCl2 solution (Giggenbach and Goguel, 1989) then smaller water-rock interaction. Thus, geochemical processes such
converted to CO and SO2 gas, respectively, to conduct the mass spec­ as base exchanges, precipitation, dissolution, and mixing with
trometry measurements. The isotope δ18O(SO4) was measured on CO gas shallow aquifers are not allowed.
in a continuous-flow gas-ratio mass spectrometer (Thermo Electron
Delta V), coupled to a thermal combustion elemental analyzer (Ther­ Relevant references for calibration of the selected geothermometers
moQuest Finnigan). The isotope δ34S(SO4) was measured on SO2 gas in a in the present study are presented in Table 2S (Supplementary Material).
continuous-flow gas-ratio mass spectrometer (ThermoQuest Finnigan
Delta PlusXL), coupled to an elemental analyzer (Costech). Dissolved
sulfate isotope composition units were expressed in V-SMOW for 3.3. Multiple mineral equilibrium approach
δ18O(SO4) and in V-CDT (Vienna Canyon Diablo Troilite; Hoefs, 2018) for
δ34S(SO4). Dissolved sulfate isotopes (δ34S(SO4) and δ18O(SO4)) were The multiple mineral equilibrium helps to pinpoint the equilibrium
analyzed using previously established standards (Rees, 1984; Morrison temperature at the saturation index (SI ¼ 0) where most minerals
et al., 1996; Morrison, 1997), and δ18O(H2O) was analyzed as described intersect (Reed and Spycher, 1984). Modeling of saturation index at
previously (Begley and Scrimgeour, 1997; Wassenaar and Koehler, different temperatures was first suggested by Reed and Spycher (1984),
1999). The level of precision of each analysis was as follows: � 0.08‰ then adopted by several authors (Park et al., 2006; Belhai et al., 2016) to
for δ18O(H2O), �0.3‰ or better (1s) for δ18O(SO4) and �0.15‰ or better compare reservoir temperatures with discharge temperatures. Besides
(1s) for δ34S(SO4), based on repeated internal standards. modelling saturation index, various geothermometers were used to
compare between reservoir and discharge temperatures (Park et al.,
2006; Belhai et al., 2016).
3.2. Geothermometrical approach It is crucial to understand the variation level of SI and its effect on the
states of water when dealing with dissolution and precipitation pro­
The temperature of CI aquifer was calculated using geothermometers cesses. Water equilibrium (between dissolution and precipitation) is
with chemical or isotope composition (i.e., solute geothermometers; reached when SI ¼ 0. When SI < 0 water becomes undersaturated and
D’Amore and Arn� orsson, 2000) considering specific geothermometry causes dissolution, whereas when SI > 0 water is oversaturated and
conditions as described previously (Fournier, 1977): leads to mineral precipitation. To calculate SI at different temperatures,

4
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

PHREEQC version 3 software (Parkhurst and Appelo, 2013) including 4.2. Isotope composition
carbfix database (Voigt et al., 2018), was used in the present study.
The isotopic composition of 18O (H2O and SO4) and δ34S(SO4) of the
3.3.1. FixAl method sixteen additional water samples (n ¼ 16) is reported in Table 1S
The CO2 add-back and FixAl methods (Pang and Reed, 1998; (Supplementary Material). The ratios of 18O(H2O) varied from 8.8 in
Palandri and Reed, 2001) were adopted in the present study. Laboratory Ouargla area to 7.5‰ in Meghaier (average ¼ 8.22‰, σ ¼ 0.38‰)
pH was considered because HCO3 was analyzed in the laboratory showing a slight positive shift towards the SSW-NNE direction. The 18O
(Pang and Reed, 1998). pH was decreased by increasing both Hþ and (SO4) ratios were clustered within a range of 11.9–13.3‰ (average ¼
HCO3 until SI (quartz, chalcedony, calcite) reached equilibrium (SI ¼ 0) 12.64‰, σ ¼ 0.43‰) within the entire studied area. The ratios of
based on the following reaction: HCO3 þ Hþ ¼ CO2 (gas) þ H2O δ34
S(SO4) did not display significant spatial variability with values ranged
The decrease of pH and addition of HCO3 were performed by trial from 13.9 to 15.3‰ (average ¼ 14.65 ‰, σ ¼ 0.36 ‰).
and error until SI ¼ 0 for quartz, chalcedony, and calcite. Quartz and
chalcedony were considered because they are not affected by CO2 loss.
For a set of minerals that is close to equilibrium with water, 4.3. Geothermometry properties
aluminum concentration is supposed unknown. The FixAl approach
consists in forcing an aluminum silicate mineral (i.e., kaolinite or 4.3.1. Chemical geothermometers
muscovite in the present study) in equilibrium with other minerals from The cation and silica geothermometers which were used to assess the
the assemblage, providing an aluminum concentration for different temperature of the collected water samples resulted in substantial
temperatures. The fixed mineral has its SI ¼ 0, resulting in a different log temperature variations (Table 2S; Supplementary Material). A wide
(Q/K) graph (after FixAL and CO2 add-back process). range of unsteady temperatures was displayed using cation geo­
thermometers with different methods: 1) Na–K geothermometer
4. Results (Arno �rsson et al., 1983) provided temperatures varying from 165.64 to
295.46 � C (average ¼ 201.90 � C, σ ¼ 38.97 � C, n ¼ 32); 2) Na–K–Ca
4.1. Hydrochemistry properties geothermometer (Benjamin et al., 1983) resulted in temperatures
ranging from 48.10 to 60.34 � C (average ¼ 56.40 � C, σ ¼ 3.54 � C, n ¼
The physicochemical analysis of the water samples (n ¼ 32) is re­ 32); and 3) Na–Li geothermometer (Sanjuan et al., 2014) produced
ported in Table 1S (Supplementary Material). The physical properties temperatures extending from 74.69 to 147.83 � C (average ¼ 107.74 � C,
showed neutral to slightly alkaline pH ranging from 6.74 to 8.35 σ ¼ 18.38 � C, n ¼ 25). Silica geothermometers yielded different tem­
(average ¼ 7.21, σ ¼ 0.37), electrical conductivity (EC) varies from 1900 peratures depending on mineral phases as follows: 1) Quartz no steam
to 3830 μs/cm (average of 2806 μs/cm, σ ¼ 0.36 μs/cm), high discharge loss, Quartz max steam loss, and Chalcedony (Fournier, 1977) displayed
temperatures varying from 44.4 to 65.1 � C (average ¼ 56.59 � C, σ ¼ temperatures ranging from 62.45 to 91.00 � C (average ¼ 76.82 � C, σ ¼
4.38 � C) with an increasing trend towards the NNE direction reaching 7.57 � C, n ¼ 32), 68.07–93.18 � C (average ¼ 80.76 � C, σ ¼ 6.66 � C, n ¼
65.1 � C in the wells of Meghaier area, and high mineralization repre­ 32), and 30.33–60.33 � C (average ¼ 45.36 � C, σ ¼ 7.94 � C, n ¼ 32),
sented by total dissolved solids (TDS) ranging from 1170 to 2309 mg/l respectively; 2) Quartz formula (Verma, 2008) generated temperatures
(average ¼ 1544.25 mg/l, σ ¼ 204.56 mg/l). Chloride concentration varying from 53.51 to 85.18 � C (average ¼ 69.38 � C, σ ¼ 8.39 � C, n ¼
ranged from 9.01 to 18.31 meq/l (average ¼ 11.40 meq/l, σ ¼ 1.87 32); and 3) Amorphous Silica (Fournier, 1977) which did not provide
meq/l). Sulfates level ranged from 8.33 to 15.20 meq/l (average ¼ 12.07 any significant result.
meq/l, σ ¼ 1.79 meq/l) typifying the main anion composition, while
sodium and calcium dominated the cation composition with concen­ 4.3.2. Isotope geothermometers
trations ranging from 8.69 to 13.91 meq/l (average ¼ 10.92 meq/l, σ ¼ Calibration of SO4–H2O geothermometer was performed for isotope
1.25 meq/l) and 4.00–10.82 meq/l (average ¼ 6.22 meq/l, σ ¼ 2.13 geothermometry with the sixteen additional water samples (n ¼ 16).
meq/l, n ¼ 32), respectively. However, magnesium, bicarbonates, and The HSO4–H2O Geothermometer (Seal et al., 2000) provided tempera­
nitrates showed trivial concentrations ranging respectively as follows: tures ranging from 73.70 to 79.84 � C (average ¼ 76.73 � C, σ ¼ 1.56 � C),
0.82–12.82 meq/l (average ¼ 7.38 meq/l, σ ¼ 2.02 meq/l); 1.49–3.25 and CaSO4–H2O geothermometer (Boschetti, 2013) produced tempera­
meq/l (average ¼ 2.07 meq/l, σ ¼ 0.35 meq/l); and 0–0.74 meq/l tures varying from 83.80 to 90.40 � C (average ¼ 87.06 � C, σ ¼ 1.67 � C).
(average ¼ 0.14 meq/l, σ ¼ 0.19 meq/l). Potassium concentrations, The calibration of SO4–H2O as previously described (Zeebe, 2010) dis­
which ranged from 0.51 to 1.28 meq/l (average ¼ 0.69 meq/l, σ ¼ 0.21 played temperatures ranging from 32.23 to 37.21 � C (average ¼ 34.75
meq/l), were relatively high compared to the norms of natural water �
C, σ ¼ 1.3 � C; Table 2S; Supplementary Material).
(Edmunds et al., 2003). High Silica (SiO2) concentrations, varying from
19.47 to 39.38 mg/l (average ¼ 28.52 mg/l, σ ¼ 5.12 mg/l), were found
in all water samples. Silica showed an increasing tendency towards the
SSW- NNE direction with a minimum of 19.47 mg/l in Ouargla area and
a maximum of 39.38 mg/l in Meghaier. Thus, SiO2 may not be consid­
ered as trace element in the sampled wells. Several trace elements were
detected such as, Lithium with concentrations going from 0.07 to 0.23
mg/l (average ¼ 0.13 mg/l, σ ¼ 0.04 mg/l), and Aluminum with con­
centrations ranging from 0.7 to 3.20 μg/l (average ¼ 1.73 μg/l, σ ¼ 0.73
μg/l).
Overall, the results indicate that the chemical composition of the
sampled wells exceeds the WHO standards for drinking water (WHO,
2017). The quality of the sampled water in the present study falls into
two categories: Na–Cl–SO4 and Ca–SO4 types in accordance with pre­
vious findings (Piper diagram in Nezli, 2009, and Abdelali et al., 2017). Fig. 3. Langelier-Ludwig diagram showing the position of the collected water
Diagrams confirming types of water are reported further below samples, with (Na þ K)¼(Σr Na þ K/Σr cations) � 50 and (HCO3) ¼ rHCO3/Σr
(Figs. 3–5). anions) � 50. (Ca þ Mg) and (SO4þCl) are complementary values that are
not calculated.

5
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

5. Discussion

5.1. Hydrogeochemical characteristics

The collected water samples may be ranged within carbonate-


evaporite geothermal systems with high sulfate content. These
geothermal systems have been extensively studied (Park et al., 2006;
Goldscheider et al., 2010; Mohammadi et al., 2010; Boschetti, 2013;
Wang et al., 2015; Yang et al., 2017; Blasco et al., 2017, 2018). The high
mineralization of water indicates a long residence time accompanied by
high temperatures as suggested by Edmunds et al. (2003), Petersen
(2014) and Petersen et al. (2018). The results revealed that the water
mineralization was mainly caused by the leaching process of the hosting
geological formation. The water type of the collected samples is shown
in Langelier-Ludwig diagram with Na–Cl–SO4 composition (Fig. 3). In
accordance with previous findings (Nezli, 2009; Abdelali et al., 2017),
the data are scattered within Na-(Cl þ SO4) and Ca-(Cl þ SO4) sets.
Moreover, the diagram shows that 8 samples (14, kh1, G2, M3, 27, 28,
29, 30) are more enriched in Ca, thus their water type is CaSO4 with high
Cl and Na contents. The remaining samples display remarkable
competition between Cl and SO4. The position of these samples in
Langelier-Ludwig diagram indicates that Ca is not derived from HCO3
suggesting gypsum or anhydrite as their origin, since they are abundant
in the investigated wells.
Sodium (Na) is the main cation in the water samples with Na–SO4
composition (Fig. 4). SO4 concentration is twofold greater than Ca,
suggesting the occurrence of non-evaporite sulfate minerals (Edmunds
et al., 2003). Plotted data in Ternary diagrams (Spencer, 2000; and
Boschetti et al., 2013) show that 30–45% of sulfate in the collected
samples originated from MgSO4 minerals, such as bloedite, epsomite,
kainite, kieserite and hexahydrite, whereas 55–70% derived from
evaporite minerals, except for sample 13 of which sulfate emanated
from gypsum dissolution (Fig. 4a). Only few samples are close to the
bloedite mineral (Fig. 4b). Thenardite, mirabilite and hanksite are the
main sources of water mineralization with 50% of NaSO4 and 50% of
CaSO4. Note that both diagrams exhibit a non-significant contribution of
HCO3 compared to sulfate (Mg–SO4 and Na–SO4) origins. Furthermore,
it has been suggested that the oxidation of pyrite (FeS2) and hydrogen
Fig. 4. Ternary diagrams after Spencer (2000) and Boschetti et al. (2013) sulfide (H2S), in similar geothermal systems with sulfate predominance,
showing the origin of sulfate in CI waters. (a): Mg–SO4–HCO3 and (b): might be the sources of sulfate (Wang et al., 2015; Park et al., 2006;
Na–SO4–HCO3 (units are in meq/l). Crosses represent the chemical composition
Mohammadi et al., 2010). The water samples contained some H2S which
of minerals: bold lines are the divisions between brine types, and dashed lines
could be detected based on its distinctive smell.
represent mixing trends. Chemical composition of Sulfate minerals is inserted as
follows: Epsomite: MgSO4 7H2O; Kainite: KClMgSO4 3H2O; Kieserite: MgSO4
H2O; Hexahydrite: MgSO4 6H2O; Bloedite: Na2SO4MgSO4 4H2O; Glauberite: 5.2. Isotope geochemistry
Na2SO4CaSO4; Polyhalite: K2SO4MgSO4 2CaSO4 2H2O; Loweite: 6Na2SO4
7MgSO4 4H2O; Thenardite: Na2SO4; Mirabilite: Na2SO4 10H2O; Aphtitalite Isotopes are widely used in geochemical investigations such as the
(Glaserite): K2.25Na1.75(SO4)2; Eugsterite: Na4Ca(SO4)3 2H2O; Hanksite: identification of Sulfur origin. In the present study, isotopes δ34S(SO4)
9Na2SO4 2NaCO3 KCl; Burkeite: 2Na2SO4Na2CO3; Gypsum: CaSO4 2H2O; and and δ18O(SO4) of CI waters were compared to their respective typical
Anhydrite: CaSO4. ranges in terrestrial materials (Fig. 5a and b). This approach has been
largely adopted (Liu et al., 2017; Laouar et al., 2002, 2018). The
4.3.3. Saturation index comparative diagram (Fig. 5) shows δ34S(SO4) ratio ranging from 10 to
The saturation index (SI) at discharge temperatures showed calcite at 20‰ which indicates a sedimentary origin of Sulfur and removes any
oversaturated or near equilibrium state, whereas other minerals (i.e., other origin from igneous rocks. The ratio of δ18O(SO4) reflects enrich­
albite, k-feldspar, anhydrite, gypsum, chalcedony, amorphous silica and ment in heavy isotope while it indicates the origin of the sedimentary
CO2 pressure) were undersaturated with negative values, except for rocks (δ18O(SO4)> 8‰; Fig. 5b) as compared to previous data of sedi­
Quartz which was near equilibrium (SI ¼ 0; Table 3S; Supplementary mentary and metasedimentary rocks (Taylor and Sheppard, 1986). The
Material). To meet typical requirements of a water reservoir, SI was enrichment of δ18O(SO4) could be explained by the exchange between
calculated at different temperatures and graphically illustrated (Fig. 8). δ18O(SO4) and δ18O (H2O) at low temperature (Chiba et al., 1998).
Mineral equilibrium at typical reservoir conditions is useful to analyze The isotopic data from the collected water samples were also used to
some geothermometrical methods and determine the temperature at perform cross plot diagrams and investigate the origin and geochemical
equilibrium. evolution of sulfate (Fig. 6). The suggested diagrams are a) δ34S(SO4)
versus δ18O(SO4), and b) Sulfate isotope δ34S(SO4) versus total concentra­
tion of SO4. The isotopes δ34S(SO4) and δ18O(SO4) of seawater (δ34S(SO4) ¼
20‰ V-CDT and δ18O(SO4) ¼ 9.5‰ V-SMOW; data after Longinelli, 1989)
were included in the diagram of δ34S(SO4) versus δ18O (SO4) for compar­
ison purpose. These diagrams were used in literature to identify the

6
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

Fig. 5. (a) Sulfur isotopic composition of CI waters


compared to its typical range in natural materials
(Data after Thode and Monster, 1965; Coleman,
1977; 1979; Claypool et al., 1980; Ishihara and
Sasaki, 1979; 1989; Kyser, 1986; Longinelli, 1989;
Laouar et al., 1990; 2018; Awaleh et al., 2015a) (b)
Oxygen isotopic composition of CI waters compared
to its typical range in terrestrial materials (Data after
Craig, 1961; Panichi and Gonfiantini, 1977; Ohmoto,
1986; Taylor and Sheppard, 1986; Giggenbach,
1992).

Fig. 6. (a) Cross plot of δ34S(SO4) versus δ18O(SO4). SW ¼ seawater composition (data after Longinelli, 1989). (b) Cross plot of Sulfur isotope δ34S versus total con­
centration of SO4.

origin of seawater or sulfur which is produced by the oxidation of sulfide The results show depleted sulfate values in heavy isotope compared
minerals (Chiba et al., 1998; Dotsika, 2012, 2015; Awaleh et al., 2015a, to seawater. Based on the depletion of δ34S(SO4) and increase of sulfate in
b; Fu et al., 2018). Data from the collected samples were plotted within the sampled water, the origin of Sulfur might be related to the oxidation
the domain of isotopic composition of sedimentary sulfides (Coleman, of sulfide minerals, in agreement with a previous study (Dotsika, 2012).
1977) and sulfate minerals, such as anhydrite or gypsum (Thode and However, in the present study there is no correlation (R2 ¼ 0.09) be­
Monster, 1965; Awaleh et al., 2015a; Laouar et al., 2018). tween SO4 total concentration and δ34S (Fig. 6b), suggesting that the

7
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

previous findings (i.e., plotted data of CI waters in the KORJINSKY di­


agram from Nezli (2009) denote the Kaolinite proxy at 50 � C).
The application of Na–K geothermometer on similar waters, such as
carbonate-evaporite geothermal systems, provides overestimated tem­
peratures (Mohammadi et al., 2010; Wang et al., 2015). Besides high
temperatures (T > 150 � C) that Na–K geothermometer generated (see
Table 2S; Supplementary Material), the collected waters are considered
immature (Fig. 7) according to the diagram of Giggenbach (1988),
which is largely used for similar purpose (Pasvanog �lu, 2015; Alçiçek
et al., 2016; Chatterjee et al., 2016; Yang et al., 2017; Blasco et al., 2017,
2018). Considering these two criteria, Na–K geothermometer is not
herein recommended. Furthermore, the collected water samples do not
reach the equilibrium between Na, K, Mg cations, which is mandatory
for an appropriate application of this geothermometer. The concentra­
tion of Na, K, Mg cations in CI waters is controlled by alteration process
rather than equilibrium (Nezli, 2009).
SiO2 geothermometers (Fournier, 1977) generate a large range of
temperatures through several mineralogical phases of SiO2. In the pre­
sent study, Amorphous Silica geothermometer yielded negative and
unreliable results which might be due to low contents of SiO2 varying
from 19.5 to 39.4 mg/l in the water samples which did not reach the
equilibrium at low temperatures (Arno �rsson, 1975). Chalcedony geo­
Fig. 7. Na–K–Mg triangular Giggenbach diagram (Giggenbach, 1988) including
thermometer underestimates reservoir temperatures, probably due to
water samples from the present study.
lack of equilibrium in chalcedony phase, resulting sometimes in lower
values than discharge temperatures. Acceptable temperatures were ob­
oxidation hypothesis is not plausible.
tained with Quartz geothermometers, using previously developed for­
Furthermore, δ34S(SO4) ratios in the present study and previous ratios
mulas by Fournier (1977) and updated by Verma (2008). The results
of 87Sr/86Sr (Edmunds et al., 2003) suggest anhydrite and/or gypsum as
from the updated formula (Verma, 2008) were about 6–10 � C lower than
local origins. The mean value of δ34S(SO4) (14.65‰) and low ratios of
87 those obtained with the original formula (Fournier, 1977), and even
Sr/86Sr (average ¼ 0.70826424) are linked to Cretaceous origin of
lower than the discharge temperatures, in some cases. Quartz geo­
evaporite with approximately 120 Ma (Paytan and Gray, 2012). In
thermometers provided an average temperature of 76.6 � C with no
contrast, high values of non-isotopic strontium rather confirm the con­
steam loss, and an average of 80.6 � C with maximum steam loss, while
tinental contribution to evaporite occurrence (Edmunds et al., 2003). In
the application of the updated formula (Verma, 2008) resulted in an
summary, Sulfur in the collected water samples mainly originated from
average temperature of 69.38 � C. Quartz geothermometers appear to
leaching of gypsum and anhydrite.
provide a better estimation of temperature compared to Amorphous
Silica and Chalcedony geothermometers. These reliable estimations can
5.3. Geothermometry be used as a proxy for the reservoir temperature indicating quartz
equilibrium. It is worth noting that both SiO2 contents and estimated
Because of the wide range of temperatures geothermometry methods temperatures fall into the quartz field in the silica solubility diagram
provide, geothermometers should be applied with caution to avoid (Arno �rsson, 1975; Gunnarsson and Arno �rsson, 2000).
overestimation or underestimation of the reservoir temperatures. The The Na–Li geothermometer was used in the present study by
accuracy of geothermometers might be checked by mineral equilibrium adopting equation (4) from Sanjuan et al. (2014) based on the chlorine
approach. In the present study, a large range of reservoir temperatures concentration (Cl < 0.3M) which matches the Cl levels in the collected
was obtained using several geothermometers (see Table 2S; Supple­ water samples. The obtained temperatures using Na–Li were over­
mentary Material). estimated but more realistic than the ones obtained using Na–K,
The Na–K Geothermometer (Arno �rsson et al., 1983) overestimated Na–K–Ca, Chalcedony and amorphous silica geothermometers. Reliable
the temperatures, although it was calibrated for low temperature res­ results were also obtained in Tunisia using Na–Li geothermometer to
ervoirs (25–250 � C). This indicates non-equilibrium conditions between estimate reservoir temperatures of the CI aquifer (Kamel, 2012). High
Albite, K-Feldspar and water. It has been suggested that Na–K geo­ lithium concentrations appear to enhance the Na–Li geothermometer.
thermometer is not recommended in the case of waters with low tem­ According to Sanjuan et al. (2014), with Li > 1 ppm reliable tempera­
perature, high sodium and calcium contents (D’Amore and Arno �rsson, tures were obtained in separate geothermal fields, namely Djibouti and
2000), as in the case of CI aquifer. Iceland. Moreover, it has been reported that low enthalpy systems with
The Na–K–Ca geothermometer (Fournier and Truesdell, 1973; low salinity and low lithium concentrations (Li < 1 ppm) should not be
Benjamin et al., 1983) was proposed to avoid the effect of high calcium assessed with Na–Li geothermometer (D’Amore et al., 1987; and Güleç,
content. The formula of Fournier and Truesdell (1973) assumed equi­ 1994).
librium between K-feldspar and Albite at high temperature (100 < T <
350 � C). However, Benjamin et al. (1983) excluded any possible equi­ 5.4. Multiple mineral equilibrium approach
librium between Naþ, Kþ and Ca2þ originating from the alteration of
silicate material. According to D’Amore and Arno �rsson (2000), at low Most of the calculated mineral saturation index (SI) at different
temperature Naþ, Kþ and Ca2þ emanate from Biotite and K-feldspar temperatures did not converge at zero axis, except for quartz and chal­
during its mineralogical transformation to Kaolinite, which does not cedony (Fig. 8). Calcite and dolomite were supersaturated due to CO2
fulfill the geothermometry conditions (thermodynamic equilibrium). degassing, while Kaolinite and Muscovite tended to be undersaturated,
The application of Na–K–Ca geothermometer to the collected water which is common in carbonate-evaporite geothermal systems
samples resulted in underestimated temperatures with lower levels than (Lo
�pez-Chicano et al., 2001; Mohammadi et al., 2010; Kamel, 2012;
discharge temperatures. The results corroborate the assumptions of Wang et al., 2015). Among the set of minerals (Fig. 8) only quartz and
D’Amore and Arno �rsson (2000) as discussed above, and agree with chalcedony were in equilibrium, displaying the same range of estimated

8
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

Fig. 8. Original SI ¼ Log (Q/K) diagram versus temperature using laboratory pH (a, c, e, g; see Table 4S; Supplementary Material). SI ¼ Log (Q/K) diagram for the
same samples after CO2 add-back and FixAl application (b, d, f, h; Table 5S; Supplementary Material). SI was calculated using carbfix database of Voigt et al. (2018).
Water equilibrium with kaolinite and muscovite is shown for theses samples (fixed minerals are not shown because their SI ¼ Log (Q/K) ¼ 0; CO2 add-back means the
number of moles of Hþ (i.e., decreased pH) and HCO3 added to the solution to compensate for the degassed CO2).

temperatures (75–85 � C). This equilibrium condition might be due to applied to the collected water samples (Fig. 8) to verify the above as­
immature waters as shown in the Giggenbach diagram (Fig. 7), or sumptions. Degassed CO2 was simultaneously re-established in the
because water samples were probably affected by CO2 degassing as re­ water samples by adding HCO3 and Hþ to maintain pH at 6.5 average
ported previously (Reed and Spycher, 1984; Pang and Reed, 1998; level, then water was equilibrated with an aluminum silicate mineral.
L�opez-Chicano et al., 2001), but most likely not due to dilution since SI The recorded concentrations of HCO3 at pH equilibrium increased
curves did not shift towards undersaturation, knowing that equilibrium nearly to the same level in all the tested water samples as follows (see
is reached at SI ¼ 0 (Pang and Reed, 1998; Palandri and Reed, 2001). Table 5S; Supplementary Material): sample 2 (pH ¼ 6.5, HCO3 ¼ 3.383
FixAl method (Pang and Reed, 1998; Palandri and Reed, 2001) was mmol/kgw), sample 18 (pH ¼ 6.54, HCO3 ¼ 3.546 mmol/kgw), sample

9
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

24 (pH ¼ 6.6, HCO3 ¼ 3.200 mmol/kgw); and sample 22 (pH ¼ 6.53,


HCO3 ¼ 3.227 mmol/kgw). Therefore, Log (Q/K) diagram becomes
more accurate displaying convergent SI curves as in the case of quartz
and chalcedony (Pang and Reed, 1998; Palandri and Reed, 2001; Blasco
et al., 2017, 2018). With slightly acidic water (pH~6.5) kaolinite and
muscovite were fixed as suggested by Pang and Reed (1998). The results
were in accordance with previous findings (Nezli, 2009). Microcline and
Albite were not used due to their disequilibrium with water (under­
saturation) confirming Na–K geothermometry assumptions.
Based on the convergent SI of the minerals aggregation, the obtained
temperatures ranged from 75 � C to 85 � C depending on locations:
Ouargla (75–76 � C); Touggourt and Djamaa (78–80 � C); and Meghaier
(84–85 � C).

5.5. Sulfate-water isotope geothermometers

Equations of δ18O (SO4–H2O) isotope geothermometer were deter­


mined based on equilibrium between HSO4 and H2O (Seal et al., 2000),
equilibrium of SO4–H2O (Zeebe, 2010), and equilibrium of CaSO4–H2O
(Boschetti, 2013).
Normally, the isotope geothermometer HSO4–H2O was calibrated for
systems with temperatures higher than, or equal to 100 � C and neutral
pH (Lloyd, 1968; Mizutani and Rafter, 1969). However, it can also be
Fig. 9. Oxygen-Water isotope fractionation factors versus Temperature (in
applied to water systems with temperatures lower than 100 � C under a
Kelvin) with calibration curves CaSO4–H2O (Boschetti, 2013); HSO4–H2O (Seal
required 500 years minimum residence time to reach equilibrium
et al., 2000); SO4–H2O (continuous line; Zeebe, 2010); and SO4–H2O (dashed
(Lloyd, 1968). The residence time of the collected water samples was line; Halas and Pluta, 2000).
estimated at about 775000 years (Petersen, 2014; Petersen et al., 2018),
thus fulfilling HSO4–H2O equilibrium condition. With this residence
composition, which occurred due leaching of hosting geological for­
time condition and equilibrium state, HSO4–H2O Formula (Seal et al.,
mation. The water composition is substantially dominated by sulfate
2000) provided relatively reliable temperatures of 73.70–79.84 � C since
over the entire sampled area. Thus, Isotope ratios of δ18O(H2O), δ18O(SO4)
the water pH was slightly acidic. The results could be more accurate if
and δ34S(SO4) were used. Isotope geochemistry involving δ34S and δ18O
the formula was applied to water systems with HSO4 predominant
of dissolved sulfates were applied to identify the origin of the leaching
composition. In the present case, HSO4–H2O reached equilibrium at near
process in the CI aquifer. Besides, δ18O(H2O) and δ18O(SO4) enabled the
neutral pH, an appropriate condition for such application as previously
application of SO4–H2O geothermometer.
suggested (Chiba and Sakai, 1985). Furthermore, results from plotted
Selected geothermometers successfully assessed the temperatures in
Oxygen-Water isotope fractionation factors versus temperature showed
CI aquifer, with Quartz (Silica) and CaSO4–H2O isotope geo­
that the tested water samples were closer to HSO4–H2O than to SO4–H2O
thermometers suggested to be the most equilibrated minerals with
equilibrium (Fig. 9).
water. On the other hand, Na–K and Na–K–Ca showed their limits by
Updated SO4–H2O formula (Zeebe, 2010), which was originally
providing unreliable results, while Na–Li formula, although over­
calibrated for low temperature sedimentary reservoirs, was also
estimated the temperatures due to lack of equilibrium, was somehow
attempted in the present study but the results were unreliable (i.e.,
close to reality. The results also showed that CaSO4–H2O geo­
estimated temperatures lower than the ones obtained at discharge).
thermometer could be successfully used with reservoir systems at low-
Although, the studied water samples had predominant SO4 content, they
medium temperatures.
were not in equilibrium with H2O (Fig. 9). This lack of equilibrium
The recorded temperatures positively correlated with the aquifer
might be due to a slow exchange of 18O between SO4 and H2O at tem­
depths over the sampled locations, which could be validated by the
peratures lower than 200 � C and low pH (Lloyd, 1968; Chiba and Sakai,
multiple mineral equilibrium approach.
1985).
The present study highlighted the potential of a geothermal source in
On the other hand, higher temperatures (83.80–90.40 � C) were ob­
southeastern Algeria, considering reservoir temperatures ranging from
tained applying CaSO4–H2O formula (Boschetti, 2013). The functioning
~70 to ~93 � C and a flow rate greater than 100 l/s, especially in
of this formula depends on gypsum or anhydrite-water equilibrium,
Djama ^a, Meghaier and El Oued areas. This geothermal source could be
which is the case in the present study (Fig. 9). The estimated tempera­
used for electricity generation and other applications for a better sus­
ture validates the reliability of CaSO4–H2O isotope geothermometer
tainable development in the Sahara of Algeria.
even for reservoirs with temperatures lower than 100 � C. The results
obtained with CaSO4–H2O isotope geothermometer are similar to those
obtained with quartz geothermometer, and both corroborate previous Declaration of competing interest
findings (Blasco et al., 2017, 2018).
The authors declare that there is no conflict of interest.
6. Conclusion
Acknowledgment
The present research focused on geochemical and geothermo­
metrical characterization of groundwater from the Continental Inter­ We would like to thank Dr. David Dettman from the Environmental
calaire (CI) aquifer at a regional scale in the northeastern Sahara of Isotope Laboratory at the University of Arizona for his help with isotope
Algeria, using chemical and isotope data. The results showed that The CI analysis, and Amistadi Mary Kay for providing trace element analysis at
waters are neutral to slightly alkaline, with a high salinity. Water the ALEC Laboratory, University of Arizona. Many thanks to the staff
chemistry is governed by calcium, sodium and very high chloride and from the National Agency of Hydraulic Resources (ANRH) for the
sulfate concentration levels. Water types are of Na–Cl–SO4 and Ca–SO4 valuable help they provided during the field work. This article is a part

10
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

of a national research project (PRFU) under the number: Chkir, N., Guendouz, A., Zouari, K., Hadj Ammar, F., Moulla, A.S., 2009. Uranium
isotopes in groundwater from the continental intercalaire aquifer in Algerian
E04N01UN300120180002. Helpful comments by the editor and re­
Tunisian Sahara (Northern Africa). J. Environ. Radioact. 100, 649–656. https://doi.
viewers greatly improved the paper quality are gratefully appreciated. org/10.1016/j.jenvrad.2009.05.009.
Claypool, G.E., Holser, W.T., Kaplan, I.R., Sakai, H., Zak, I., 1980. The age curves of
sulfur and oxygen isotopes in marine sulfate and their mutual interpretation. Chem.
Appendix A. Supplementary data Geol. 28, 199–260. https://doi.org/10.1016/0009-2541(80)90047-9.
Clesceri, L.S., Greenberg, A.E., Eaton, A.D., 1999. Standard Methods for the Examination
Supplementary data to this article can be found online at https://doi. of Water and Wastewater, twentieth ed. American Public Health Association,
American Water Works Association, Water Environment Federation, Washington.
org/10.1016/j.apgeochem.2019.104492.
Coleman, M.L., 1977. Sulphur isotopes in petrology. J. Geol. Soc. 133, 593–608. https://
doi.org/10.1144/gsjgs.133.6.0593.
References Coleman, M.L., 1979. Isotopic analysis of trace sulphur from some S- and I-type granites:
heredity or environment. In: Atherton, M.P., Tarney, J. (Eds.), Origin of Granite
Batholiths: Geochemical Evidence Based on a Meeting of the Geochemistry Group of
Abdelali, A., Nezli, I.E., Benhamida, A.S., 2017. Geothermometry application to the
the Mineralogical Society. Birkh€ auser Boston, Boston, MA, pp. 129–133. https://doi.
Continental Intercalaire geothermal aquifer: a case study of Ouargla region. Energy
org/10.1007/978-1-4684-0570-5_11.
Procedia 119, 264–269. https://doi.org/10.1016/j.egypro.2017.07.079.
Cornet, A., 1964. Introduction a � l’hydrog� eologie saharienne. G�eog. Phys. et G�eol. Dyn.
Alçiçek, H., Bülbül, A., Alçiçek, M.C., 2016. Hydrogeochemistry of the thermal waters
VI, 5–72 fasc. 1.
from the yenice geothermal field (denizli basin, Southwestern Anatolia, Turkey).
Craig, H., 1961. Isotopic variations in meteoric waters. Science 133, 1702–1703. https://
J. Volcanol. Geotherm. Res. 309, 118–138. https://doi.org/10.1016/j.
doi.org/10.1126/science.133.3465.1702.
jvolgeores.2015.10.025.
D’Amore, F., Arn� orsson, S., 2000. Geothermometry. In: Arn� orsson, S. (Ed.), Isotopic and
Arn�orsson, S., 1975. Application of the silica geothermometer in low temperature
Chemical Techniques in Geothermal Exploration, Development and Use.
hydrothermal areas in Iceland. Am J Sci U. S. 275, 7. https://doi.org/10.2475/
International Atomic Agency, Vienna, pp. 152–199.
ajs.275.7.763.
Dotsika, E., 2012. Isotope and hydrochemical assessment of the Samothraki Island
Arn�orsson, S., Gunnlaugsson, E., Svavarsson, H., 1983. The chemistry of geothermal
geothermal area, Greece. J. Volcanol. Geotherm. Res. 233–234, 18–26. https://doi.
waters in Iceland. II. Mineral equilibria and independent variables controlling water
org/10.1016/j.jvolgeores.2012.04.017.
compositions. Geochem. Cosmochim. Acta 47, 547–566. https://doi.org/10.1016/
Dotsika, E., 2015. H–O–C–S isotope and geochemical assessment of the geothermal area
0016-7037(83)90277-6.
of Central Greece. J. Geochem. Explor. 150, 1–15. https://doi.org/10.1016/j.
Awaleh, M.O., Hoch, F.B., Boschetti, T., Soubaneh, Y.D., Egueh, N.M., Elmi, S.A.,
gexplo.2014.11.008.
Mohamed, J., Khaireh, M.A., 2015. The geothermal resources of the Republic of
D’Amore, F., Fancelli, R., Caboi, R., 1987. Observations on the application of chemical
Djibouti — II: geochemical study of the Lake Abhe geothermal field. J. Geochem.
geothermometers to some hydrothermal systems in Sardinia. Geothermics 16,
Explor. 159, 129–147. https://doi.org/10.1016/j.gexplo.2015.08.011.
271–282.
Awaleh, M.O., Hoch, F.B., Kadieh, I.H., Soubaneh, Y.D., Egueh, N.M., Jalludin, M.,
Edmunds, W.M., Guendouz, A.H., Mamou, A., Moulla, A., Shand, P., Zouari, K., 2003.
Boschetti, T., 2015. The geothermal resources of the Republic of Djibouti — I:
Groundwater evolution in the Continental Intercalaire aquifer of southern Algeria
hydrogeochemistry of the Obock coastal hot springs. J. Geochem. Explor. 152,
and Tunisia: trace element and isotopic indicators. Appl. Geochem. 18, 805–822.
54–66. https://doi.org/10.1016/j.gexplo.2015.02.001.
https://doi.org/10.1016/S0883-2927(02)00189-0.
Begley, I.S., Scrimgeour, C.M., 1997. High-precision δ2H and δ18O measurement for
Edmunds, W.M., Shand, P., Guendouz, A., Moulla, A.S., Mamou, A., Zouari, K., 1997.
water and volatile organic compounds by continuous-flow pyrolysis isotope ratio
Recharge characteristics and groundwater quality of the Grand erg orientale basin,
mass spectrometry. Anal. Chem. 69, 1530–1535. https://doi.org/10.1021/
Final report. EC (Avicenne) Contract CT93AVI0015, BGS Technical Report WD/97/
ac960935r.
46R, Hydrogeology series.
Bel, F., Cuche, D., 1969. Mise au point des connaissances sur la nappe du Complexe
Fabre, J.A., 1976. Introduction � a la g�eologie du Sahara alg� erien et des r�egions voisines.
Terminal. ERESS, Ouargla, Alg� erie, p. 20.
Soci�et�e Nationale d’Edition
� et de Diffusion, Alger, p. 421.
Belhai, M., Fujimitsu, Y., Bouchareb-Haouchine, F.Z., Haouchine, A., Nishijima, J., 2016.
Fournier, R.O., 1977. Chemical geothermometers and mixing models for geothermal
A hydrochemical study of the Hammam Righa geothermal waters in north-central
systems. Geothermics 5, 41–50. https://doi.org/10.1016/0375-6505(77)90007-4.
Algeria. Acta Geochim 35, 271–287. https://doi.org/10.1007/s11631-016-0092-8.
Fournier, R.O., Truesdell, A.H., 1973. An empirical Na-K-Ca geothermometer for natural
Ben Dhia, H., Meddeb, N., 1990. Application of chemical geothermometers to some
waters. Geochem. Cosmochim. Acta 37, 1255–1275. https://doi.org/10.1016/0016-
Tunisian hot springs. Geothermics 19, 87–104. https://doi.org/10.1016/0375-6505
7037(73)90060-4.
(90)90068-M.
Fu, C., Li, X., Ma, J., Liu, L., Gao, M., Bai, Z., 2018. A hydrochemistry and multi-isotopic
Benjamin, T., Charles, R., Vidale, R., 1983. Thermodynamic parameters and
study of groundwater origin and hydrochemical evolution in the middle reaches of
experimental data for the Na-K-Ca geothermometer. J. Volcanol. Geotherm. Res.,
the Kuye River basin. Appl. Geochem. 98, 82–93. https://doi.org/10.1016/j.
Geothermal Energy of Hot Dry Rock 15, 167–186. https://doi.org/10.1016/0377-
apgeochem.2018.08.030.
0273(83)90099-9.
Giggenbach, W.F., 1988. Geothermal solute equilibria. Derivation of Na-K-Mg-Ca
Bishop, W.F., 1975. Geology of Tunisia and adjacent parts of Algeria and Libya. AAPG
geoindicators. Geochem. Cosmochim. Acta 52, 2749–2765. https://doi.org/
Bull. 59, 413–450.
10.1016/0016-7037(88)90143-3.
Blasco, M., Auqu�e, L.F., Gimeno, M.J., Acero, P., Asta, M.P., 2017. Geochemistry,
Giggenbach, W.F., 1992. Isotopic shifts in waters from geothermal and volcanic systems
geothermometry and influence of the concentration of mobile elements in the
along convergent plate boundaries and their origin. Earth Planet. Sci. Lett. 113,
chemical characteristics of carbonate-evaporitic thermal systems. The case of the
495–510. https://doi.org/10.1016/0012-821X(92)90127-H.
Tiermas geothermal system (Spain). Chem. Geol. 466, 696–709. https://doi.org/
Giggenbach, W.F., Goguel, R.L., 1989. Collection and Analysis of Geothermal and
10.1016/j.chemgeo.2017.07.013.
Volcanic Water and Gas Discharges. Report No. CD 2401. Department of Scientific
Blasco, M., Gimeno, M.J., Auqu�e, L.F., 2018. Low temperature geothermal systems in
and Industrial Research, Chemistry Division, Petone, New Zealand.
carbonate-evaporitic rocks: mineral equilibria assumptions and geothermometrical
Goldscheider, N., M� adl-Sz}onyi, J., Er}
oss, A., Schill, E., 2010. Review: thermal water
calculations. Insights from the Arnedillo thermal waters (Spain). Sci. Total Environ.
resources in carbonate rock aquifers. Hydrogeol. J. 18, 1303–1318. https://doi.org/
615, 526–539. https://doi.org/10.1016/j.scitotenv.2017.09.269.
10.1007/s10040-010-0611-3.
Boschetti, T., 2013. Oxygen isotope equilibrium in sulfate–water systems: a revision of
Gonçalv�es, J., Vallet-Coulomb, C., Petersen, J., Hamelin, B., Deschamps, P., 2015.
geothermometric applications in low-enthalpy systems. J. Geochem. Explor. 124,
Declining water budget in a deep regional aquifer assessed by geostatistical
92–100. https://doi.org/10.1016/j.gexplo.2012.08.011.
simulations of stable isotopes: case study of the Saharan “Continental Intercalaire.
Boschetti, T., De Felice, V., Celico, F., 2013. The Pozzo del Sale Groundwaters (Irpinia,
J. Hydrol 531, 821–829. https://doi.org/10.1016/j.jhydrol.2015.10.044.
Southern Apennines, Italy): origin and Mechanisms of Salinization. Aquat. Geochem.
Guendouz, A., 1985. Contribution � a l’�
etude g�eochimique et isotopique des nappes
19, 303–322. https://doi.org/10.1007/s10498-013-9196-5.
profondes du Sahara Nord-Est septentrional, Alg� erie. Ph.D. Thesis. Paris 11, p. 243.
Busson, G., 1972. Principes, m� ethodes et r�esultats d’une �
etude stratigraphique du
Guendouz, A., Michelot, J.-L., 2006. Chlorine-36 dating of deep groundwater from
M�esozoïque saharien [Principles, methods and results of a stratigraphic study of the
northern Sahara. J. Hydrol 328, 572–580. https://doi.org/10.1016/j.
Mesozoic of the Sahara]. M�em. mus�eum Nat. Histoire Naturelle., Paris XXVI, p. 441.
jhydrol.2006.01.002. The ICWRER Symposium, Dresden, Germany.
Chaib, W., Kherici, N., 2014. Hydrochemistry and geothermometry of an albian aquifer
Güleç, N., 1994. Geochemistry of thermal waters and its relation to the volcanism in the
from oued righ region in northeastern Algerian Sahara. Geotherm. Energy 2. https://
Kizilcahamam (Ankara) area, Turkey. J. Volcanol. Geotherm. Res. 59, 295–312.
doi.org/10.1186/s40517-014-0003-3.
https://doi.org/10.1016/0377-0273(94)90084-1.
Chatterjee, S., Sharma, S., Ansari, M.A., Deodhar, A.S., Low, U., Sinha, U.K., Dash, A.,
Gunnarsson, I., Arn� orsson, S., 2000. Amorphous silica solubility and the thermodynamic
2016. Characterization of subsurface processes estimation of reservoir temperature
properties of H4SiO� 4 in the range of 0� to 350� C at Psat. Geochem. Cosmochim.
in Tural Rajwadi geothermal fields, Maharashtra, India. Geothermics 59, 77–89.
Acta 64, 2295–2307. https://doi.org/10.1016/S0016-7037(99)00426-3.
https://doi.org/10.1016/j.geothermics.2015.10.011.
Halas, S., Pluta, I., 2000. Empirical calibration of isotope thermometer of δ 18 O (SO 4 2-
Chiba, H., Sakai, H., 1985. Oxygen isotope exchange rate between dissolved sulfate and
)-δ 18 O (H 2 O) for low temperature brines. In: V Isotope Workshop, pp. 68–71.
water at hydrothermal temperatures. Geochem. Cosmochim. Acta 49, 993–1000.
Krak� ow, Poland.
https://doi.org/10.1016/0016-7037(85)90314-X.
Hoefs, J., 2018. Stable Isotope Geochemistry, 8th Ed, Springer Textbooks in Earth
Chiba, H., Uchiyama, N., Teagle, D.A.H., 1998. Stable isotope study of anhydrite and
Sciences, Geography and Environment. Springer International Publishing.
sulfide minerals at the tag hydrothermal mound, mid-Atlantic ridge, 26� N1. In:
Ishihara, S., Sasaki, A., 1989. Sulfur isotopic ratios of the magnetite-series and ilmenite-
Herzig, P.M., Humphris, S.E., Miller, D.J., Zierenberg, R.A. (Eds.), Proceedings of the
series granitoids of the Sierra Nevada batholith—a reconnaissance study. Geology
Ocean Drilling Program, vol.158. Scientific Results.

11
A. Abdelali et al. Applied Geochemistry 113 (2020) 104492

17, 788–791. https://doi.org/10.1130/0091-7613(1989)017<0788:SIROTM>2.3. Paytan, A., Gray, E.T., 2012. Sulfur isotope stratigraphy. A geological time scale. In:
CO;2. Gradstein, F., et al. (Eds.), The Geologic Time Scale, vol. 1. Elsevier, pp. 167–180.
Kamel, S., 2012. Application of selected geothermometers to Continental Intercalaire https://doi.org/10.1016/B978-0-444-59425-9.00009-3 (Chapter 9).
thermal water in southern Tunisia. Geothermics 41, 63–73. https://doi.org/ Petersen, J.O., 2014. Traçage isotopique (36CI, 4He, 234U) et mod�elisation
10.1016/j.geothermics.2011.10.003. hydrog�eologique du Syst� eme Aquif�ere du Sahara Septentrional. Application � a la
Kharaka, Y.K., Mariner, R.H., 1989. Chemical geothermometers and their application to recharge Quaternaire du Continental Intercalaire.. Ph.D. Thesis Aix-Marseille,
formation waters from sedimentary basins. In: Naeser, N.D., McCulloh, T.H. (Eds.), p. 325.
Thermal History of Sedimentary Basins. Springer New York, New York, NY, Petersen, J.O., Deschamps, P., Hamelin, B., Goncalves, J., Michelot, J.-L., Zouari, K.,
pp. 99–117. https://doi.org/10.1007/978-1-4612-3492-0_6. 2013. Water-rock interaction and residence time of groundwater inferred by 234U/
Kyser, T.K., 1986. Stable isotope variations in the mantle. Rev. Miner. 16, 141–164. 238U disequilibria in the Tunisian continental intercalaire aquifer system. procedia
Laouar, R., Boyce, A.J., Fallick, A.E., Leake, B.E., 1990. A sulphur isotope study on earth planet. In: Science Proceedings of the Fourteenth International Symposium on
selected Caledonian granites of Britain and Ireland. Geol. J. 25, 359–369. https:// Water-Rock Interaction, WRI, 14 7, pp. 685–688. https://doi.org/10.1016/j.
doi.org/10.1002/gj.3350250318. proeps.2013.03.206.
Laouar, R., Boyce, A.J., Ahmed-Said, Y., Ouabadi, A., Fallick, A.E., Toubal, A., 2002. Petersen, J.O., Deschamps, P., Hamelin, B., Fourr� e, E., Gonçalv�es, J., Zouari, K.,
Stable isotope study of the igneous, metamorphic and mineralized rocks of the Guendouz, A., Michelot, J.-L., Massault, M., Dapoigny, A., Team, A., 2018.
Edough complex, Annaba, Northeast Algeria. J. Afr. Earth Sci. 35, 271–283. https:// Groundwater flowpaths and residence times inferred by 14C, 36Cl and 4He isotopes
doi.org/10.1016/S0899-5362(02)00037-4. in the Continental Intercalaire aquifer (North-Western Africa). J. Hydrol 560, 11–23.
Laouar, R., Lekoui, A., Bouima, T., Salmi-Laouar, S., Bouhlel, S., Abdallah, N., Boyce, A. https://doi.org/10.1016/j.jhydrol.2018.03.003.
J., Fallick, A.E., 2018. Petrology, geochemistry and stable isotope studies of the Reed, M., Spycher, N., 1984. Calculation of pH and mineral equilibria in hydrothermal
Miocene igneous rocks and related sulphide mineralisation of Oued Amizour (NE waters with application to geothermometry and studies of boiling and dilution.
Algeria). Ore Geol. Rev. 101, 312–329. https://doi.org/10.1016/j. Geochem. Cosmochim. Acta 48, 1479–1492. https://doi.org/10.1016/0016-7037
oregeorev.2018.07.026. (84)90404-6.
Lindal, B., 1973. Industrial and other applications of geothermal energy. Geotherm. Rees, C.E., 1984. The Isotopic Analysis of Sulphur. McMaster University. Isotopic,
Energy 135–148. Nuclear and Geochemical Studies Group. Contribution No. 139.
Liu, M., Guo, Q., Zhang, C., Zhu, M., Li, J., 2017. Sulfur isotope geochemistry indicating Rodier, J., 2016. L’analyse de l’eau, 9e �edition. URL. https://www.chapitre.com/
the source of dissolved sulfate in gonghe geothermal waters, northwestern China. BOOK/rodier-jean-legube-bernard-merlet-nicole/l-analyse-de-l-eau-9e-edition,2297
Procedia Earth Planet. Sci. 17, 157–160. https://doi.org/10.1016/j. 4734.aspx. (accessed 2.16.19).
proeps.2016.12.039. Saibi, H., 2009. Geothermal resources in Algeria. Renew. Sustain. Energy Rev. 13,
Lloyd, R.M., 1968. Oxygen isotope behavior in the sulfate-water system. J. Geophys. Res. 2544–2552. https://doi.org/10.1016/j.rser.2009.06.019.
73, 6099–6110. https://doi.org/10.1029/JB073i018p06099. Sanjuan, B., Millot, R., Asmundsson,
� R., Brach, M., Giroud, N., 2014. Use of two new Na/
Longinelli, A., 1989. Oxygen-18 and sulphur-34 in dissolved oceanic sulphate and Li geothermometric relationships for geothermal fluids in volcanic environments.
phosphate. Mar. Environ. 219–255. Chem. Geol. 389, 60–81. https://doi.org/10.1016/j.chemgeo.2014.09.011.
L�
opez-Chicano, M., Cer� on, J.C., Vallejos, A., Pulido-Bosch, A., 2001. Geochemistry of Sasaki, A., Ishihara, S., 1979. Sulfur isotopic composition of the magnetite-series and
thermal springs, Alhama de Granada (southern Spain). Appl. Geochem. 16, ilmenite-series granitoids in Japan. Contrib. Mineral. Petrol. 68, 107–115. https://
1153–1163. https://doi.org/10.1016/S0883-2927(01)00020-8. doi.org/10.1007/BF00371893.
Makni, J., Bouri, S., Dhia, H.B., 2013. Hydrochemistry and geothermometry of thermal Seal, R.R., Alpers, C.N., Rye, R.O., 2000. Stable isotope systematics of sulfate minerals.
groundwater of southeastern Tunisia (Gabes region). Arab. J. Geosci. 6, 2673–2683. Rev. Mineral. Geochem. 40, 541–602. https://doi.org/10.2138/rmg.2000.40.12.
Mizutani, Y., Rafter, T.A., 1969. Oxygen isotopic composition of sulphates. 3. Oxygen Slimani, R., Guendouz, A., Trolard, F., Moulla, A.S., Hamdi-Aïssa, B., Bourri�e, G., 2017.
isotopic fractionation in the bisulfate ion-water system. N. Z. J. Sci. 12, 54–59. Identification of dominant hydrogeochemical processes for groundwaters in the
Mohammadi, Z., Bagheri, R., Jahanshahi, R., 2010. Hydrogeochemistry and Algerian Sahara supported by inverse modeling of chemical and isotopic data.
geothermometry of Changal thermal springs, Zagros region, Iran. Geothermics 39, Hydrol. Earth Syst. Sci. 21, 1669–1691. https://doi.org/10.5194/hess-21-1669-
242–249. https://doi.org/10.1016/j.geothermics.2010.06.007. 2017.
Morrison, J., 1997. Inorganic oxygen isotope analysis by EA-pyrolysis-IRMS. In: 4th Spencer, R.J., 2000. Sulfate minerals in evaporite deposits. In: Charles, N.A., Jambor, J.
Canadian Continuous Flow – IRMS Conference, Waterloo 24-27 August 1997. L., Nordstrom, D.K. (Eds.), Sulfate Minerals: Crystallography, Geochemistry, and
Morrison, J., Fallick, T., Donelly, T., Leossen, M., St Jean, G., Drimmie, R.J., 1996. δ34S Environmental Significance. Mineralogical Society of America, Geochemical Society,
Measurements of Standards from Several Laboratories by Continuous Flow Isotope pp. 173–192. https://doi.org/10.2138/rmg.2000.40.3.
Ratio Mass Spectrometry (CF-IRMS). Micromass UK Ltd. Technical Note TN 309, Takherist, D., Lesquer, A., 1989. Mise en �evidence d’importantes variations r�egionales du
April 1996. flux de chaleur en Alg� erie. Can. J. Earth Sci. 26, 615–626. https://doi.org/10.1139/
Moulla, A.S., Guendouz, A., Cherchali, M.E.-H., Chaid, Z., Ouarezki, S., 2012. Updated e89-053.
geochemical and isotopic data from the Continental Intercalaire aquifer in the Great Taylor, H.P.J., Sheppard, S.M.F., 1986. Igneous rocks : I. Processes of isotopic
Occidental Erg sub-basin (south-western Algeria). Quat. Int. 257, 64–73. https://doi. fractionation and isotope systematics. Rev. Mineral. 16, 227–271.
org/10.1016/j.quaint.2011.08.038. Palaeogroundwater dynamics and their Thode, H.G., Monster, J., 1965. Sulfur-isotope geochemistry of petroleum. Evaporites,
importance for past human settlements and today’s water management. and Ancient Seas 71, 367–377.
Nezli, I.E., 2009. Approche hydrog�eochimique � a l’�
etude des aquif�eres de la basse vall�
ee UNESCO, 1972. Projet Reg 100. Etude des ressources en eau du Sahara septentrional.
de l’Oued M’ya (Ouargla). Ph.D. Thesis. Universit�e Mohamed Kheider de Biskra, Rapport sur les r�esultats du projet. UNESCO, Paris.
Alg�erie, p. 146. Verma, M.P., 2008. Qrtzgeotherm: an ActiveX component for the quartz solubility
Ohmoto, H., 1986. Stable isotope geochemistry of ore deposit. In stable isotopes in high geothermometer. Comput. Geosci. 34, 1918–1925. https://doi.org/10.1016/j.
temperature geological processes. Rev. Mineral. 16, 491–559. cageo.2008.01.008.
OSS, 2003. Syst� eme Aquif�ere du Sahara Septentrional. Observatoire du Sahara et du Voigt, M., Marieni, C., Clark, D.E., Gíslason, S.R., Oelkers, E.H., 2018. Evaluation and
Sahel. Volume 2 : Hydrog�eologie. Projet SASS. Coupes. Planches. Annexes. Tunis, P. refinement of thermodynamic databases for mineral carbonation. Energy Procedia
275. Volume 4 : Mod�ele Math� ematique. Annexes. Tunis, P. 229. 146, 81–91. https://doi.org/10.1016/j.egypro.2018.07.012.
Ould Baba Sy, M., 2005. Recharge et pal�eorecharge du syst� eme aquif� ere du Sahara Wang, J., Jin, M., Jia, B., Kang, F., 2015. Hydrochemical characteristics and
septentrional. Ph.D. Thesis. Facult�e des Sciences de Tunis, Tunisie, p. 277. geothermometry applications of thermal groundwater in northern Jinan, Shandong,
Palandri, J.L., Reed, M.H., 2001. Reconstruction of in situ composition of sedimentary China. Geothermics 57, 185–195. https://doi.org/10.1016/j.
formation waters. Geochem. Cosmochim. Acta 65, 1741–1767. https://doi.org/ geothermics.2015.07.002.
10.1016/S0016-7037(01)00555-5. Wassenaar, L.I., Koehler, G., 1999. An on-line technique for the determination of the
Pang, Z.-H., Reed, M., 1998. Theoretical chemical thermometry on geothermal waters: δ(18)O and δ(17)O of gaseous and dissolved oxygen. Anal. Chem. 71, 4965–4968.
problems and methods. Geochem. Cosmochim. Acta 62, 1083–1091. https://doi. https://doi.org/10.1021/ac9903961.
org/10.1016/S0016-7037(98)00037-4. WHO, 2017. Guidelines for Drinking-Water Quality: Fourth Edition Incorporating the
Panichi, C., Gonfiantini, R., 1977. Environmental isotopes in geothermal studies. First Addendum. World Health Organization, Geneva. Licence: CC BY-NC-SA 3.0
Geothermics 6, 143–161. https://doi.org/10.1016/0375-6505(77)90024-4. IGO.
Park, S.-S., Yun, S.-T., Chae, G.-T., Hutcheon, I., Koh, Y.-K., So, C.-S., Choi, H.-S., 2006. Yang, P., Cheng, Q., Xie, S., Wang, J., Chang, L., Yu, Q., Zhan, Z., Chen, F., 2017.
Temperature evaluation of the Bugok geothermal system, South Korea. Geothermics Hydrogeochemistry and geothermometry of deep thermal water in the carbonate
35, 448–469. https://doi.org/10.1016/j.geothermics.2006.07.002. formation in the main urban area of Chongqing, China. J. Hydrol 549, 50–61.
Parkhurst, D.L., Appelo, C.A.J., 2013. Description of input and examples for PHREEQC https://doi.org/10.1016/j.jhydrol.2017.03.054.
version 3—a computer program for speciation, batch-reaction, one-dimensional Zeebe, R.E., 2010. A new value for the stable oxygen isotope fractionation between
transport, and inverse geochemical calculations. U.S. Geol. Surv. Tech. Methods dissolved sulfate ion and water. Geochem. Cosmochim. Acta 74, 818–828. https://
book 6, chap. A43, 497 pp. http://pubs.usgs.gov/tm/06/a43/. doi.org/10.1016/j.gca.2009.10.034.
Pasvano� glu, S., 2015. The Seben-Kesen€ ozü low-temperature geothermal prospect, NW
Turkey: study of an advective geothermal system. Environ. Earth Sci. 74, 363–376.
https://doi.org/10.1007/s12665-015-4041-3.

12

You might also like