Next Article in Journal
Insights into Ferric Leaching of Low Grade Metal Sulfide-Containing ores in an Unsaturated Ore Bed Using X-ray Computed Tomography
Previous Article in Journal
Kinetic Study of the Leaching of Low-Grade Manganese Ores by Using Pretreated Sawdust as Reductant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Geochemistry and Stable Isotopes of Travertine from Jordan Valley and Dead Sea Areas

by
Khalil M. Ibrahim
1,*,
Issa M. Makhlouf
1,
Ali R. El Naqah
2 and
Sana’ M. Al-Thawabteh
1
1
Department of Earth and Environmental Sciences, Hashemite University, P.O. Box 150459, Zarqa 13115, Jordan
2
Faculty of Natural Resources and Environment, Hashemite University, P.O. Box 150459, Zarqa 13115, Jordan
*
Author to whom correspondence should be addressed.
Minerals 2017, 7(5), 82; https://doi.org/10.3390/min7050082
Submission received: 15 February 2017 / Revised: 12 May 2017 / Accepted: 13 May 2017 / Published: 22 May 2017

Abstract

:
Travertine deposits in Deir Alla, Suwayma, and Az Zara areas were investigated. Mineralogy, geochemistry, stable isotopes and age dating indicate the presence of low-Mg calcite, with minor quartz components. The variable isotope (δ13C and δ18O) signatures indicate dependence on water temperature and water/rock isotopic exchange. In contrast, the high δ13C values in some travertine samples reflect 12CO2 degassing processes, increased input of 13C-enriched groundwater, and the presence of surface and groundwater hydrological systems. The high δ18O values may be attributed to evaporation effects and low water temperature during the formation of localized travertine. The age of travertine is the Late Pleistocene.

1. Introduction

Travertine refers to all non-marine carbonate precipitates formed at or near terrestrial springs, rivers, lakes, and caves [1], when CO2-rich, Ca-bearing fluids are exposed to low pressure conditions at surface conditions [2]. Travertine formation is controversial since it has been attributed to both biotic and abiotic processes [3,4]. The terms travertine and tufa are sometimes used synonymously [5]. Calcite precipitation occurs due to CO2-loss, and associated pH-reduction during fluid equilibration with the atmosphere [6,7]. Travertine precipitation seems to be controlled by a combination of several factors, such as: (i) a fast decrease of the hydrostatic pressure and the pCO2, resulting in a rapid degassing process; and (ii) microbes and algae activity [6]. The term travertine must be retained for continental carbonates mainly composed of calcium carbonate deposits produced from non-marine, supersaturated calcium bicarbonate-rich waters, typically hydrothermal in origin [5]. Based on the CO2 interaction with the groundwater, Pentecost [8] subdivided travertine into meteogene and thermogene deposits. CO2 in the meteogene deposits is developed at a shallower origin, as a result of organic activity in the soil [9]. Thermogene deposits of CO2 have a deeper origin either from magmatic degassing or from decarbonation processes [10]. Such deposits are typical of tectonically active areas where geothermal heat flux (endogenic or volcanic) is high [11]. The meteogene travertine is usually referred to as “tufa”, especially that which contains remains of micro and macrophytes, invertebrates and bacteria [11,12]. It is usually derived from lower water temperature, and lower contents of dissolved inorganic carbon (Dissolved inorganic carbon (DIC) < 10 mmol·L−1), CO2 partial pressure (pCO2 < 0.1 atm) and higher pH values (7–8) [10]. Meteogene usually displays low deposition rates (<10 mmol·cm−2 year−1) and shows a negative carbon isotopic composition (δ 13C vs. PDB between −12 and 0‰). However, thermogene travertine displays higher deposition rates and higher (δ 13C vs. PDB) values [5,10]. They are characterized chiefly by high depositional rates, regular bedding and fine lamination, low porosity, low permeability and an inorganic crystalline fabric [5]. The main distinctive characteristics of travertine and tufa are detailed in Capezzuoli et al. [5].
Travertine and tufa rocks are exposed in the eastern margins of the Dead Sea and Jordan Valley. Some investigations focused on their geotechnical properties as a decoration stone [13]. However, geochemical and mineralogical studies of such rocks are limited [14,15]. The present study aims to elucidate the mineralogy, geochemistry, stable isotope implications and age determination of the travertine and tufa formation along the eastern margins of the Dead Sea and Jordan Valley.

1.1. Location

The study area is located along the margins of the Dead Sea (Figure 1). Most of the travertine outcrops are located in the eastern Jordan Valley and the Dead Sea Transform Fault system. Three main travertine outcrops were studied including: Deir Alla, Suwayma and Az Zara (Figure 1). Representative travertine, tufa and water samples were collected from these areas.

1.2. Geological Setting

The travertine is exposed in a highly tectonized area associated with the main Dead Sea Transform Fault (Figure 2). It is an 1100 km long sinistral fault system that connects the Gulf of Aqaba-Red Sea spreading system to the convergence zone of the Taurus–Zagrous Mountains [16]. This fault is a left lateral transform plate boundary, separating the Arabian plate in the east and the Palestine-Sinai sub plate (part of the African plate) in the west [17]. It has been active since ca. 13–18 Ma ago with movement continuing today [17,18]. Most of the faults in the study area are related to the stress fields associated with the transform boundary.
The stratigraphic sequence in the studied areas, as illustrated in the simplified geological maps in Figure 2, comprises a wide range of rock units including Cambrian, Permian, Triassic, Jurassic and Cretaceous sandstones, shale and carbonates (Figure 2). Quaternary deposits occur widely in the form of marls, siltstone, basalts, travertine, fluvial and alluvial sediments.
The studied travertine facies are composed of crystalline crust, shrub, paper-thin raft, coated gas bubble, reed, lithoclasts, pebbly travertine, and palaeosols with a tufa carbonate facies of macrophyte encrustation deposits, bryophyte build-ups, biomicrites and green and grey marls (Figure 3). The Deir Alla travertine is formed in a terraced and smooth slope depositional setting, where it has been subjected to meteoric diagenesis.
The inactive thermogene Suwayma travertine is deposited in fissure-ridge and depression depositional setting, with meteogene and burial diagenetic processes [19]. Both localities are characterized by hard crystalline travertine with low organic content. Az Zara travertines were deposited under self-built channels’ depositional settings. Az Zara active tufa was formed in a predominantly fluvio-palaustrine environment, which is friable and rich in organics and clays.

1.3. Hydrology and Hydrogeology

The physiographic situation of the Dead Sea and Jordan Valley displays morphotectonic depressions (Figure 1). Therefore, most of the rainfall-infiltrated water emerges once again in the form of springs, and the general flow direction is due west, towards the Dead Sea and Jordan Valley. The water temperature of some springs is moderate (18–21 °C). The discharge amount is seasonal such that it increases during the winter and drops in the summer [20].
The groundwater around the Dead Sea is found in two aquifer complexes; the first is the upper limestone aquifer complex (Upper Cretaceous), and the second is the sandstone aquifer complex (Lower Cretaceous). The total amount of groundwater associated with the upper aquifer is around 87 million m3/year [20], and half of it discharges on the surface through springs. The total discharge of the lower aquifer is around 90 million m3/year, and the water is mostly thermal and mineralized [21]. The Jordan Valley floor consists of recent alluvial fan sediments inter-tonguing with Pleistocene salty, clayey sediments of the Lisan Formation. The available groundwater in this area ranges from 18–20 million m3/year [22].
Springs issued east of Deir Alla are alkaline with prevailing bicarbonate in summer, and alkaline with prevailing chlorides in winter. Fe oxides precipitate as soon as the water mixes with the Zarqa River water [23]. The number of thermal springs in Ma’in area is approximately 109, with temperature exceeding 34 °C. All such thermal springs are classified as hyperthermal [24]. Sixty-four hyperthermal springs are located in the northern part of Wadi Zarqa Ma’in forming the “Hammamt Ma’in”, with 19 million m3/year total discharge into the Dead Sea. The rest of the hyperthermal springs are aligned in a north–south trend at the Az Zara area, on the eastern shore of the Dead Sea. The total discharge into the Dead Sea from the Az Zara area is 17 million m3/year [23].

2. Materials and Methods

More than 80 travertine samples were collected from three investigated areas representing both vertical and horizontal sections. The mineralogy of 18 samples was determined by X-ray diffraction (XRD), using a Phillips X’ Pert Pro PW 3040/80 diffractometer (Lelweg 1, 7602, EA Almelo, The Netherlands). Twenty-two samples were selected for major and trace elements determination by X-ray fluorescence (XRF)—using a Phillips MagiX PW 2424 diffractometer (Lelweg 1, 7602, EA Almelo, The Netherlands). Radiogenic and stable isotope measurements were also carried out. The method used for δ18O measurements is CO2 equilibrium/Delta Plus XP Isotope Ratio Mass Spectrometer HDO (ThermoFisher Scientific, Walton, MA, USA) using a platinum catalyst modified from (International Atomic Energy Agency). Measurements of 14C and δ13C are done by a benzene synthesis line and Liquid Scintillation Counter. The method is modified by (IAEA) technical procedure #25. The measurements of the (13C/12C) isotopes are made relative to the Vienna Pee Dee Belemnite standard (VPDB), and the water oxygen is based on a Vienna Standard Mean Ocean Water (VSMOW).
The ratios of (13C/12C) and (18O/18O) are expressed in the δ notation. The values of δ18O are interchangeable according to such equations [8]: δ18O (VSMOW) = [1.03088 δ18O (VPDB) + 30.88] and δ18O (VPDB) = [0.97008 δ18O (VSMOW) − 29.94].
The calculation of travertine age was performed by the equation set by Vogel [25] and used by Thorpe et al. [26]. The decay rate constant (λ) is assumed to equal 1.209 × 10−4 year−1. The value dilution factor is 0.85 ± 0.05 [8,26].

3. Results and Discussion

Petrographic investigation reported the presence of eight travertine and four tufa lithofacies in the studied sites. The description of these lithofacies is summarized in Table 1.

3.1. Mineralogy

Semiquantitative determination of mineral phases was reported by the X-ray diffraction. In the present work, components other than calcium carbonate may be defined as accessories. X-ray diffraction of travertine from Deir Alla, Suwayma, and Az Zara indicates close similarity. They are predominantly composed of calcite (Figure 4). Deir Alla and Suwayma samples occasionally contain a very low percent of quartz and Fe-oxides. Sample (D26a) shows a frequent amount of aragonite (Figure 4). Traces of gypsum, dolomite and halite may also occur.
Petrographic study by Al-Thawabteh [19] indicated that most of the Fe present in travertine and tufa is autochthonous, i.e., formed or grown in situ and has not been transported. Precipitation of hydrated Fe-oxides occurs when groundwater containing ferrous ions encounters the atmosphere. In contrast, quartz enters the travertine realm as a detrital mineral, derived from the soil and bedrocks, and are therefore considered as allochthonous minerals.
Calculations of the theoretical mineralogical composition including MgCO3 percent are presented in Table 2. In Deir Alla, the detrital quartz grains are rare compared to Suwayma and Az Zara (Table 2). It is believed that they are derived from sandstone when hot groundwater emerges onto the surface. In addition, Deir Alla travertine includes different types of clay minerals that vary from 0.19–1.47 wt % (Table 2). In thermogene travertine of Suwayma and Az Zara, the presence of clay (2.82 wt % and 0.85 wt %, respectively) may be due to the hydrothermal alteration of the host rock.

3.1.1. MgCO3

The calculated MgCO3 represents all of Mg in the studied travertine. The Mg is usually incorporated in the calcite, aragonite, and dolomite crystals, or it may form magnesite. In Deir Alla travertine, the calculated MgCO3 content ranges from 0.3 wt % to 0.8 wt %, with a mean value of 0.4 wt %, whereas in Suwayma travertine, the range is from 0.9 wt % to 2.2 wt %, with an average value of 1.7%; meanwhile, the calculated MgCO3 content of Az Zara travertine and tufa ranges from 1.0% to 1.8%, with a mean value of 1.3% (Table 2). This indicates that the calcite in such travertines and tufas is invariably of low Mg calcite [27] (Table 2). Calcite that contains less than 8% of MgCO3 has been defined as low-Mg calcite, whereas calcite, with more than 8–11% of MgCO3, has been defined as high-Mg calcite [27].

3.1.2. Calcium Carbonate (CaCO3)

Deir Alla travertine consists mainly of calcium carbonate with an arithmetic average value of 95.5% (Table 2, shown as cc). According to Pentecost [8], meteogene deposits have a high average of CaCO3 content with a median of about 92.9%, and an upper range of 99.3%. In addition, Pentecost [8] indicated that active meteogene deposits have a lower CaCO3 content (mean 91.2%, No. of samples = 28) than their inactive meteogene counterparts (mean 94.8%, No. of samples = 22). This might be due to the loss of organic matter in the older deposits. Consequently, the Deir Alla travertine can be classified as inactive super-ambient meteogene deposits based on the CaCO3 content [8].
The mean calcium carbonate abundance in Suwayma travertine is about 89.18% (Table 2). The CaCO3 content of thermogene travertine is broadly similar to that of meteogene [8]. Therefore, the Suwayma travertine can be classified as thermogene deposits based on its CaCO3 content. Az Zara travertine mainly consists of calcium carbonate with a mean of 92% (Table 2). Az Zara travertines and tufa can be classified as inactive and active thermogenic deposits, respectively. Az Zara deposits are localized and are commonly associated with regions of Quaternary basaltic outcrops and recent tectonic activities related to the Dead Sea Transform Fault System.
In general, calcium carbonates in travertine are formed of various combinations of aragonite and calcite polymorphs [28]. No attempts were carried out in this study to investigate the controls of calcium carbonate polymorph precipitation. The factors that control calcite and aragonite precipitation in travertine are still open to debate [28,29,30,31]. The precipitation has been variously attributed to water composition, water temperature, growth inhibitors (e.g., Mg/Ca ratio), and/or CO2 degassing and saturation levels [29]. Aragonite may co-precipitate with calcite at a high CO2 content and rapid CO2 degassing, irrespective of the Mg/Ca ratio [30], or aragonite may precipitate from low levels of CO2 degassing, provided the Mg/Ca ratio is high enough to inhibit calcite precipitation [30].

3.2. Geochemistry

3.2.1. CaO and MgO

The XRF results indicate that CaO wt % in the travertine ranges from 48.38% to 55.8% (Table 3). The relationship between CaO% and SiO2% is presented in Figure 5a supported by the negative correlation coefficient of −0.82 in the correlation coefficient matrix (Figure 6). The inverse relationship between these two oxides (Figure 5a) indicates that they are involved in two different phases. A similar conclusion can be noted in Figure 5b, which exhibits the relationship between CaO and Al2O3.
It is evident that the studied travertine samples have low Mg content (Table 3), which is less than 1.0 wt % MgO with a mean value of 0.19 wt %. Consequently, calcites of such travertine are invariably of low Mg. The high value of Mg in some Suwayma travertine samples (Table 3) may have resulted from the input of Mg-enriched solutions that could lead to slight dolometization; i.e., formation of diagenetic dolomite. Mg forms a solid solution series with calcite and a wide range of compositions are possible. Figure 6 shows the negative correlation between CaO and MgO, which indicates that Mg substitutes Ca in calcite crystals. Previous studies by Pentecost [8] show that a few meteogene travertines are composed of calcite, which exceed 1% of Mg by weight, and yield a mean value of 0.29 wt %. Thermogene travertine consisting of calcite has similar Mg concentration similar meteogene travertine [8].

3.2.2. Fe2O3 and MnO

The Fe2O3 content of Deir Alla samples ranges from 0.24% to 3.18% (Table 3). In Suwayma travertine, Fe2O3 wt % ranges from 0.47% to 4.72%, and Az Zara travertine and tufa is between 0.18% and 0.88%. Such percents are for total iron, which include the Fe precipitated with calcite and that associated with detrital minerals. According to Pentecost [8], the mean Fe wt % for the meteogene and thermogene travertines is 0.18% to 0.28%, respectively, which is far less than the mean value of the studied travertines (0.83%). This is most probably related to the presence of allochthonous Fe-minerals within the travertine.
The mean value of MnO content in Deir Alla and Suwayma travertines is about 0.14% and 0.03%, respectively, while the mean value for meteogene and thermogene travertines is 0.0188% and 0.091%, respectively [8]. MnO in Deir Alla and Suwayma travertines is most probably incorporated into calcium carbonate. The high percent of Mn in D26a sample is zoned in aragonite (Table 3). The low percent of MnO content reflects a high rate of travertine deposition in such areas [29]. Consequently, the precipitation rate in Suwayma was higher than that in Deir Alla. High percent of MnO was also recorded in Z1* tufa sample (Table 3), due to the presence of plants, especially mosses, which contain manganese in their internal structure, similar to those described by Obeidat [15]. Such MnO content indicates a slow rate of deposition of tufa [32]. Caboi et al. [33] found that Mn is mainly associated with calcite, while Fe is associated with the detrital minerals.
Pentecost [8] observed that when Fe levels are high, Mn tends to precipitate with it, but when Fe is low, Mn is incorporated into calcium carbonate. The selected travertine samples define a negative correlation between CaO and MnO (Figure 6). This might indicate that Mn is incorporated into calcium carbonate and replaces Ca.

3.2.3. Alkali Metals

The alkali metals are believed to occupy interstitial positions in the calcite lattice with a frequency of occurrence: Li > Na > K > Rb [34]. However, in aragonite, they substitute for Ca2+ in the crystal structure. Therefore, they are more easily co-precipitated with aragonite than with calcite [34]. It is noted that some Na concentrations in the samples are associated with Cl to form halite. In Deir Alla travertine, Na and K contents average 0.085% and 0.022%, respectively, and, similar to those found in the super-ambient travertines of Matlock Bath, UK [35]. Therefore, Deir Alla travertine is considered as super-ambient meteogene travertine, especially when the water temperature rises above ambient issuing hot springs (>37 °C) [8]. Figure 5c shows that K is correlated with Al, with a positive 0.98 correlation coefficient (Figure 6). This may indicate that they were found during the same mineral phases, such as clay minerals (potassium aluminum silicates). This is supported by the positive 0.98 correlation coefficient with SiO2 (Figure 6).

3.2.4. Sr

The Sr content of Deir Alla samples ranges from 423 to 4189 mg/kg. In Suwayma travertine, Sr ranges from 511 to 7987 mg/kg, whereas Az Zara travertine and tufa Sr ranges from 585 to 2213 mg/kg (Table 3). The range of Sr concentration in the meteogene travertine is from 9 to >2930 mg/kg, and in the thermogene travertine is 20–14,000 mg/kg. In general, the Sr proportion in calcite is less than that in aragonite [8]. The transformation of aragonite to calcite during meteoric water circulation in depth leads to precipitation of calcite with less Sr content [8].
There is probably no single universal factor that controls calcite and aragonite precipitation in spring systems. For example, calcite was precipitated directly from waters with temperatures of >90 °C in Kenya [29] and New Zealand [36]. According to Fouke et al. [1], only aragonite precipitates at water temperatures >44 °C, whereas calcite and aragonite precipitate at 30–43 °C, while calcite alone forms when temperature falls below 30 °C. The Sr proportion increases in high-temperature travertine facies [1]. Therefore, the low temperature Deir Alla travertines have low Sr concentrations and are commonly composed of pure calcite. Some samples that contain a high percent of Sr may reflect their high aragonite content, especially those in D26a samples that showed aragonite recrystallization under the microscope [19]. High Sr content in Suwayma travertine reflects deposition under high water temperature like aragonite, before its complete transformation into calcite. The association of aragonite and calcite in Az Zara samples indicates their formation under high water temperatures (25–45 °C), before the complete transformation of aragonite into calcite.
Kinsman [37], Veizer and Demovic [38] considered that the diagenetic solution is one of the main factors that control the distribution of Sr in the carbonate rocks. In addition, Cipriani et al. [39,40] found an appropriate correlation between Sr content and porosity, which means that calcite must contain less Sr as porosity decreases. Deir Alla travertine was affected by diagenetic processes, such as their tight lithification, high compaction, low pore spaces, and good cementation. In contrast, Suwayma travertine was affected much less by diagenetic processes, as indicated by their lose lithification, low compaction, high pore spaces, and weak cementation. A high Sr value was recorded in the Z1 sample of the spongy tufa that was slightly affected by diagenetic processes, as compared to hard travertine samples Z3 and Z4.
The negative correlation between Ca and Sr abundances in travertine samples indicates that some Ca was replaced by Sr in the calcite lattice. Ichikuni [41] observed that the partition coefficients of Sr are higher in both aragonite and calcite, if a small amount of Mn has substituted for Ca in the lattice.

3.3. Stable Isotopes

3.3.1. Background Information

The (13C/12C) and (18O/18O) ratios undergo discrete change during biogeochemical cycling, allowing inferences about their origin and past history. Such ratios provide an expanded scale for the comparative small differences observed [42]. For travertine, fractionation of carbon and oxygen isotopes is of particular interest. They provide information about the carbon dioxide source, the physio-chemical conditions of the precipitation event (rate and temperature), and the influence of metabolic processes. In favorable circumstances, they provide information about the temperature at which deposition occurred, and assist in radiometric dating [42].
Stable oxygen isotope ratios are of great interest to geochemists and carbonate sedimentologists. As the lithosphere water exchanges oxygen atoms with the dissolved carbon dioxide, and where equilibrium is established, the difference between the ratio of oxygen isotopes in carbonates and depositing water can be used to estimate past water temperatures, assuming that the fluid source has not changed through time. Information can also be obtained on rates of water evaporation with time [8,42].
The carbon stable isotopes also undergo fractionation during chemical reactions. The fractionations of δ13C occur between gas phase CO2 and the dissolved carbonate species. The CO2 gas is slightly depleted in 13C, whilst bicarbonates and carbonates are enriched by about 9% at 20 °C [8].
Carbonates have δ18O (VSMOW) values ranging from about 0 to +35%. Their interpretation is generally less straightforward than carbon isotope values because carbonate oxygen readily exchanges with oxygen in water molecules [8]. Biotical and geochemical processes such as photosynthesis and temperature decrease during flow process, evaporation, mixing from soil profile or surface waters will exert an influence on the final oxygen isotope value for any carbonate sample [43,44].

3.3.2. Deir Alla Travertine

The results indicate that the δ13C isotope values in the eight samples from Deir Alla are between −1.45 and 2.99 (‰VPDB) (Table 4), while the δ18O values vary from −8.13 to 8.17 (‰VPDB). Deir Alla samples are described as super-ambient meteogene travertines [8]. The range of stable carbon isotope value in such a travertine type is from −12 to +2 (‰VPDB), and the median of stable oxygen isotope is −7.92 (‰VPDB) [8]. The bivariate distribution of δ18O and δ13C (‰VPDB) values in Deir Alla samples (Figure 7) shows four different travertine lithofacies that are: shrub, coated bubble, paper-raft + reed, and crystalline crust lithofacies. The stable isotope values are variable in the different sub-environments, due to possible variations in conditions (i.e., water temperature, flow velocity and distance from spring, surface area and biological effects). Figure 7 indicates that Deir Alla travertine is enriched in 13C. The high δ13C values in some travertine facies (as shown in D26b and D29 samples) can be explained by a combination of several processes: (i) increased CO2 degassing processes related to travertine formation [45]; and (ii) increased input from heavier thermal-spring water 13C-enriched groundwater. The cooling process of spring water during the flow tract is responsible for the loss of δ18O from the water system by evaporation; as a result, the δ18O value will increase. This case was observed only in the D3 sample.

3.3.3. Suwayma Travertine

The δ13C isotope values in the six samples of Suwayma travertine vary between −0.18 and 1.28 (‰VPDB), while δ18O values vary from −5.38 to 8.12 (‰VPDB) as shown in Table 4. Suwayma travertine is considered as thermogene travertine. The stable carbon isotope in such travertine type ranges from −1 to +10 (‰VPDB), and the median of stable oxygen isotope is −0.95 (‰VPDB) [8].
The δ18O and δ13C (%VPDB) bivariate distribution of Suwayma travertines shows three different lithofacies: lithoclast, reed, and coated bubble (Figure 7). Such results explain the enrichment of δ13C, which depends on the depositional environment fluctuations. The δ13C is likely enriched due to the increase of degassing processes that are associated with the minor microbiologic and photosynthetic activities. The occurrence of 13C-enriched carbonate in the Suwayma slope indicates the presence of 13C-enriched surface water and groundwater that might be washed-out during hydrological open systems. It is quite evident that the low δ18O values are related to the poor evaporation effects and high water temperature during travertine formation. Additionally, the δ18O depletion in the thermogene water was displayed by the isotopic exchange with the host rock [46].

3.3.4. Az Zara Tufa

The δ18O and δ13C values in the Az Zara tufa deposits are distinctively low compared with the other areas (Figure 7). This may indicate different isotopic signatures. The average values of carbon and oxygen stable isotopes are −2.3‰ ± 0.2‰ VPDB and −8.58‰ ± 0.7‰ VPDB, respectively.
One per mill (1%) drop in the carbonate δ18O value is equivalent to about a 4.5 °C temperature rise under equilibrium conditions [8]. The δ18O compositions reflect the short stagnancy time of water. The lack of evaporation or continuous recharge could be the reason in such a fluvio-palustrine depositional environment. The low δ18O values may also have resulted from isotopic exchange between the circulating thermal waters and the country rock and/or water adsorbed in the clay minerals.
The occurrence of 13C-poor carbonate in the Az Zara area indicates the presence of 13C-poor surface and groundwater that might be washed-out when the hydrological systems are closed. The variations in δ13C are minimal (−3.52‰ to −1.34‰), reflecting similar carbon sources for all facies. The carbon dioxide is believed to originate from Upper Cretaceous limestone decarbonation reaction as a result of the Karst process; however, a magmatic component may also be significant and originates from the Pleistocene basaltic activity, which is dominant within two of the studied sites (Figure 2), and probably records the influence of meteoric and soil-derived CO2.

3.4. Travertine Dating

3.4.1. Background Information

According to Srdoč et al. [47,48], if travertine is to be dated using 14C, then the sources of carbon must be known. The 14C dating method is not reliable for thermogene travertine because their carbon content is fossil (dead carbon/inorganic carbon). More reliable dates may be obtained by using a U-Series date [49,50] or by using the Electron Spin Resonance method [50]. For instance, where cosmogenic carbon is the only source, then dating is straightforward and the level of 14C in the sample gives a direct measure of age [8]. The decay of isotopes 14C has been used most extensively, and can date deposits up to 30 Ka. The isotope 14C has a half-life of 5730 years and forms continuously in the atmosphere by the interaction of cosmic rays with molecules of nitrogen.

3.4.2. Age of Travertine

The previous studies did not provide any absolute ages of travertine in Jordan. The present work applied the radiocarbon method for dating Deir Alla and Suwayma travertines. Two rock samples were selected for radiocarbon dating. The travertine age calculations unfortunately yielded unreliable ages, and these indicate an age of 32.84 ± 0.005 Ka (Late Pleistocene).

4. Conclusions

The mineralogical compositions of the total travertine samples consist mainly of calcite with varying amounts of quartz (i.e., detrital quartz). Moreover, some aragonites are found in Deir Alla and Suwayma travertines. Based on geochemical studies, low Mg calcite is the main travertine component in all locations examined here. Accordingly, the Deir Alla travertines were classified as inactive super-ambient meteogene deposits. The Suwayma and Az Zara travertines were classified as inactive thermogene. Moreover, Az Zara tufa was classified as an active thermogene deposits.
The Sr abundances indicate pure calcite in the Deir Alla travertines and the diagenetic effect to be more than that recorded in Suwayma and Az Zara travertines. The Suwayma travertine was deposited as aragonite that completely transformed into calcite. Moreover, Az Zara travertine and tufa were deposited under high temperature water as the component of aragonite and calcite. Such aragonites are also completely transformed into calcite. The high δ13C values in the travertine facies are related to evaporation effects, CO2 degassing processes, increased input of 13C-enriched groundwater, and the presence of surface and groundwater transportation medium in hydrological open systems. The high δ18O values may be attributed to evaporation effects and low water temperature that prevailed during travertine formation. The travertine formed in the Late Pleistocene.

Acknowledgments

The authors thank the Deanship of Research and Graduate Studies of Hashemite University very much for funding this work through the project entitled “Genesis and palaeo-environmental significance of hot spring travertines in Jordan”. A huge thanks is due to Engneer Qais Al Qaisi, the Director of Travco Company (Amman, Jordan), who granted fieldwork access to study the travertine at the company site at Deir Alla. In addition, earnest thanks are due to Al al-Bayt University for the XRF analysis and the Water Authority for the radiogenic and stable isotope measurements. A special thanks is also extended to the anonymous reviewers of this article for their critical comments and suggestions.

Author Contributions

Khalil Ibrahim, Sana’ M. Al-Thawabteh, Issa M. Makhlouf and Ali R. El Naqah conducted the field work; Sana’ M. Al-Thawabteh performed the mineral and chemical analysis; Khalil Ibrahim and Sana’ M. Al-Thawabteh analyzed the data; and Khalil Ibrahim wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fouke, B.W.; Farmer, J.D.; Des Marais, D.J.; Pratt, L.; Sturchio, N.C.; Burns, P.C.; Discipulo, M.K. Depositional facies and aqueous-solid geochemistry of travertine-depositing hot springs (Angel Terrace, Mammoth Hot Springs, Yellowstone National Park, USA). J. Sediment. Res. 2000, 70, 565–585. [Google Scholar] [CrossRef]
  2. Jamtveit, B.; Hammer, Ø.; Andersson, C.; Dysthe, D.K.; Heldmann, J.; Vogel, M.L. Travertines from the Troll thermal springs, Svalbard. Norwegian. J. Geol. 2006, 86, 387–395. [Google Scholar]
  3. Fouke, B.W. Hot-spring systems geobiology: Abiotic and biotic influences on travertine formation at Mammoth hot springs, Yellowstone National Park, USA. Sedimentology 2011, 58, 170–219. [Google Scholar] [CrossRef]
  4. Zhu, T.; Dittrich, M. Carbonate precipitation through microbial activities in natural environment, and their potential in biotechnology: A review. Front. Bioeng. Biotechnol. 2016, 4, 4. [Google Scholar] [CrossRef] [PubMed]
  5. Capezzuoli, E.; Candin, A.; Pedley, M. Decoding tufa and travertine (fresh water carbonates) in the sedimentary record: The state of the art. Sedimentology 2014, 61, 1–21. [Google Scholar] [CrossRef]
  6. Rodrigo-Naharro, J.; Delgado, A.; Herrero, M.J.; Granados, A.; Pérez del Villar, L. Current travertines precipitation from CO2-rich groundwaters as an alert of CO2 leakages from a natural CO2 storage at Gañuelas-Mazarrón Tertiary Basin (Murcia, Spain). Inf. Técnicos Ciemat 2013, 1279, 1–53. [Google Scholar]
  7. Dreybrodt, W.; Buhmann, D.; Michaelis, J.; Usdowski, E. Geochemically controlled calcite precipitation by CO2 outgassing: Field measurements of precipitation rates in comparison to theoretical predictions. Chem. Geol. 1992, 97, 285–294. [Google Scholar] [CrossRef]
  8. Pentecost, A. Travertine, 1st ed.; Springer: London, UK, 2005; ISBN 978-1-4020-3606-4. [Google Scholar]
  9. Pentecost, A.; Vile, H.A. A review and reassessment of travertine classification. Géogr. Phys. Quat. 1994, 48, 305–314. [Google Scholar] [CrossRef]
  10. Alessandro, D.W.; Giammanco, S.; Bellomo, S.; Parello, F. Geochemistry and mineralogy of travertine deposits of the SW flank of Mt. Etna (Italy): Relationships with past volcanic and degassing activity. J. Volcanol. Geotherm. Res. 2007, 165, 64–70. [Google Scholar] [CrossRef]
  11. Ford, T.D.; Pedley, H.M. A review of tufa and travertine deposits of the world. Earth Sci. Rev. 1996, 41, 117–175. [Google Scholar] [CrossRef]
  12. Arenas, C.; Gutiérrez, F.; Osàcar, C.; Sancho, C. Sedimentology and geochemistry of fluvio-lacustrine tufa deposits controlled by evaporate solution subsidence in the central Ebro Depression, NE Spain. Sedimentology 2000, 47, 883–909. [Google Scholar] [CrossRef]
  13. Jasser, D. Investigation of Deir Alla Travertine; Natural Resources Authority: Amman, Jordan, 1984.
  14. Khoury, H.; Salameh, E.; Udluft, P. On the Zerqa Ma’in (thermal kallirrhoes) travertine/Dead Sea (hydrochemistry, geochemistry and isotopic composition). Neues Jahrb. Geol. Palaontol. Mh 1984, H8, 472–484. [Google Scholar]
  15. Obeidat, O. Geochemistry, Mineralogy, and Petrography of Travertine of Deir Alla and Zerqa Ma’in Hot Springs. Master’s Thesis, Yarmouk University, Irbid, Jordan, 1992. [Google Scholar]
  16. Ibrahim, K.M.; Moh’d, B.K.; Masri, A.I.; Al-Taj, M.M.; Musleh, S.M.; Alzughoul, K.A. Volcanotectonic evolution of central Jordan: Evidence from the Shihan volcano. J. Afr. Earth Sci. 2014, 100, 541–553. [Google Scholar] [CrossRef]
  17. Smit, J.; Brun, J.P.; Cloetingh, S.; Ben-Avraham, Z. The rift-like structure and asymmetry of the Dead Sea Fault. Earth Planet. Sci. Lett. 2010, 290, 74–82. [Google Scholar] [CrossRef]
  18. Ferry, M.; Meghraoui, M.; Abou Karaki, N.; Al-Taj, M.; Khalil, L. Episodic behavior of the Jordan Valley section of the Dead Sea fault inferred from a 14-kyr long integrated catalogue of large earthquakes. Bull. Seism. Soc. Am. 2011, 101, 39–87. [Google Scholar] [CrossRef]
  19. Al-Thawabteh, S.M. Sedimentology, Geochemistry, and Petrographic Study of Travertine Deposits along the Eastern Side of the Jordan Valley and Dead Sea Areas. Master’s Thesis, Hashemite University, Zarqa, Jordan, 2008. [Google Scholar]
  20. Salameh, E.; Bannayan, H. Water Resources of Jordan: Present Status and Future Potentials; Friedrich Ebert Shiftung: Amman, Jordan, 1993. [Google Scholar]
  21. Salameh, E.; Udluft, P. The Hydrodynamic Pattern of Central Jordan; Geol Jb, C38: Hannover, Germany, 1985. [Google Scholar]
  22. Water Authority of Jordan. Open Files of the Water Authority of Jordan; Water Authority of Jordan: Amman, Jordan, 1991.
  23. Salameh, E. Curative Water in Jordan; WRSC Bull, 7th Issue; University of Jordan: Amman, Jordan, 1988. [Google Scholar]
  24. Abu Ajamieh, M. The Geothermal Resources of Zarqa Ma’in and Zara; Natural Recourses Authority: Amman, Jordan, 1980.
  25. Vogel, J.C. Carbon-14 dating of groundwater. In Isotopes Hydrology; IAEA: Vienna, Austria, 1970. [Google Scholar]
  26. Thorpe, P.M.; Holyoak, D.T.; Preece, R.C.; Willing, M.J. Validity of corrected 14C dates from calcareous tufa. In Actes Colloq. I’AGF Form. Carbonatées Externs Travertines Paris; Springer: London, UK, 1981; pp. 151–156. [Google Scholar]
  27. Greensmith, J.T. Petrology of the Sedimentary Rocks, 6th ed.; George Allen and UNWIN: London, UK, 1978; ISBN 0 04 552011 9. [Google Scholar]
  28. Kawano, J.; Shimobayashi, N.; Miyake, A.; Kitamura, M. Precipitation diagram of calcium carbonate polymorphs: Its construction and significance. J. Phys. Condens. Matter 2009, 21, 425102–425107. [Google Scholar] [CrossRef] [PubMed]
  29. Jones, B.; Renaut, R.W. Calcareous spring deposits in continental settings. In Carbonates in Continental Settings: Facies, Environments, and Processes; Alonso Zarza, A.M., Taner, L.H., Eds.; Elsevier Science: Amsterdam, The Netherlands, 2010; Volume 61, pp. 177–224. [Google Scholar]
  30. Brian, J. Review of Calcium Carbonate Polymorph Precipitation in Spring Systems. Sediment. Geol. 2017, 353, 64–75. [Google Scholar] [CrossRef]
  31. Özkul, M.; Kele, S.; Gökgöz, A.; Shen, C.C.; Jones, B.; Baykara, M.O.; Fόrizs, I.; Nemeth, T.; Chang, Y.W.; Alcicek, M.C. Comparison of the Quaternary travertine sites in the Denizli Extensional Basin based on their depositional and geochemical data. Sediment. Geol. 2013, 294, 179–204. [Google Scholar] [CrossRef]
  32. Rao, C.P. Petrography, trace elements and oxygen and carbon isotopes of Grodon group carbonates (Ordovician), Florentine Valley, Tasmania, Australia. J. Sediment. Geol. 1990, 80, 221–231. [Google Scholar] [CrossRef]
  33. Caboi, R.; Cidu, R.; Fanfani, L.; Zuddas, P.; Zuddas, P.P. Geochemistry of Funtana Maore travertines (Central Sardinia, Italy). Miner. Petrogr. Acta 1991, 34, 77–93. [Google Scholar]
  34. Okumura, M.; Kitano, Y. Coprecipitation of alkali metal ions with calcium carbonate. Geochim. Cosmochim. Acta 1986, 50, 49–58. [Google Scholar] [CrossRef]
  35. Pentecost, A. The origin and development of the travertines and associated thermal waters at Matlock Bath, Derbyshire. Proc. Geol. Assoc. 1999, 110, 217–232. [Google Scholar] [CrossRef]
  36. Jones, B.; Renaut, R.W.; Rosen, M.R. High-temperature (>90 °C) calcite precipitation at Waikite Hot Springs, North Island, New Zealand. J. Geol. Soc. 1998, 153, 481–498. [Google Scholar] [CrossRef]
  37. Kinsman, D.J. Interpretation of Sr+2 concentration in carbonate minerals and rocks. J. Sediment. Pterol. 1989, 39, 488–508. [Google Scholar] [CrossRef]
  38. Veizer, J.; Demovic, R. Strontium as a tool in facies analysis. J. Sediment. Pterol. 1974, 44, 93–115. [Google Scholar] [CrossRef]
  39. Cipriani, N.; Ercoli, A.; Malesani, P.; Vannucci, S. I travertini di Rapolano Term (Siena). Mem. Soc. Geol. Ital. 1972, 11, 31–48. [Google Scholar]
  40. Cipriani, N.; Malesani, P.; Vannucci, S. I travertine dell’Italia centrale. Boll. Serv. Geol. Ital. 1977, 98, 85–115. [Google Scholar]
  41. Ichikuni, M. Partition of strontium between calcite and solution: Effect of substitution by manganese. Chem. Geol. 1973, 11, 315–319. [Google Scholar] [CrossRef]
  42. Sun, H.; Liu, Z. Wet-dry seasonal and spatial variations in the δ13C and δ18O values of the modern endogenic travertine at Baishuitai, Yunnan, SW China and their paleoclimatic and paleoenvironmental implications. Geochim. Cosmochim. Acta 2010, 74, 1018–1029. [Google Scholar] [CrossRef]
  43. Chafetz, H.S.; Guidry, S.A. Bacterial shrubs, crystal shrubs, and ray-crystal crusts: Bacterially induced vs. a biotic mineral precipitation. Sediment. Geol. 1999, 128, 57–74. [Google Scholar] [CrossRef]
  44. Guo, L.; Andrews, J.; Riding, R.; Dennis, P.; Dresser, Q. Possible microbial effects on stable carbon isotopes in hot-spring travertines. J. Sediment. Res. 1996, 66, 468–473. [Google Scholar] [CrossRef]
  45. Turi, B. Stable isotope geochemistry of travertines. In Handbook of Environmental Isotope Geochemistry; Fritz, B.P., Fontes, L.J.C., Eds.; Elsevier: Amsterdam, The Netherlands, 1986; ISBN 9781483289830. [Google Scholar]
  46. Fritz, P. Der Isotopengehalt der Mineralwasser von Stuttgart und Umgebung und ihrer ittel pleistozaenen Travertin-Ablagerungen. Jber. Mitt. Oberrhein. Geol. Ver. 1968, 50, 53–58. [Google Scholar]
  47. Srdoč, D.; Chafetz, H.; Utech, N. Radiocarbon dating of travertine deposits, Arbuckle Mountains, Oklahoma. Radiocarbon 1989, 31, 819–828. [Google Scholar] [CrossRef]
  48. Nishikawaa, O.; Furuhashia, K.; Masuyamaa, M.; Ogatab, T.; Shiraishia, T.; Shenc, C. Radiocarbon dating of residual organic matter in travertine formed along the Yumoto Fault in Oga Peninsula, northeast Japan: Implications for long-term hot spring activity under the influence of earthquakes. Sediment. Geol. 2012, 243, 181–190. [Google Scholar] [CrossRef]
  49. Lovea, A.J.; Shanda, P.; Karlstromd, K.; Crosseyd, L.; Rousseau-Gueutina, P.; Priestleya, S.; Wholinge, D.; Fultonf, S.; Keppela, M. Geochemistry and travertine dating provide new insights into the hydrogeology of the Great Artesian Basin, South Australia. Procedia Earth Planet. Sci. 2013, 7, 521–524. [Google Scholar] [CrossRef]
  50. Lebatard, A.; Alcicek, M.C.; Rochette, P.; Khatib, S.; Vialet, A.; Boulbes, N.; Bourles, D.L.; Demory, F.; Guipert, G.; Mayda, S.; et al. Dating the Homo erectus bearing travertine from Kocabas, (Denizli, Turkey) at least 1.1 Ma. Earth Planet. Sci. Lett. 2014, 390, 8–18. [Google Scholar] [CrossRef]
Figure 1. Location map to show the studied travertine outcrops.
Figure 1. Location map to show the studied travertine outcrops.
Minerals 07 00082 g001
Figure 2. Geological maps of the studied travertine outcrops.
Figure 2. Geological maps of the studied travertine outcrops.
Minerals 07 00082 g002
Figure 3. Selected field photos of the studied travertine: (A) root casts, downward tapering, rare lateral roots, smooth tube-like, ranging from 0.5 to 7.0 cm in length and from 0.1 to 3.5 cm in diameter preserved in porous travertine from Deir Alla; (B) tiny travertine from Deir Alla with irregular elliptical cavities are partly to completely infilled by secondary materials; (C) interconnected root traces on the upper bedding plain indicating the plants types of Az Zara area; (D) stalagmites and stalactites features (red circle), which are formed near Az Zara hot spring; (E) wavy laminae resembles stromatolite in Suwayma travertine, bedding plains indicate three stages of travertine precipitation; (F) the mound of travertine standing in an erosional relief in Suwayma. The hammer is 28 cm long.
Figure 3. Selected field photos of the studied travertine: (A) root casts, downward tapering, rare lateral roots, smooth tube-like, ranging from 0.5 to 7.0 cm in length and from 0.1 to 3.5 cm in diameter preserved in porous travertine from Deir Alla; (B) tiny travertine from Deir Alla with irregular elliptical cavities are partly to completely infilled by secondary materials; (C) interconnected root traces on the upper bedding plain indicating the plants types of Az Zara area; (D) stalagmites and stalactites features (red circle), which are formed near Az Zara hot spring; (E) wavy laminae resembles stromatolite in Suwayma travertine, bedding plains indicate three stages of travertine precipitation; (F) the mound of travertine standing in an erosional relief in Suwayma. The hammer is 28 cm long.
Minerals 07 00082 g003
Figure 4. Selected X-ray diffractograms of the studied travertines (a) D8, (b) D26a.
Figure 4. Selected X-ray diffractograms of the studied travertines (a) D8, (b) D26a.
Minerals 07 00082 g004aMinerals 07 00082 g004b
Figure 5. Variation diagrams of: (a) CaO versus SiO2; (b) CaO versus Al2O3; and (c) Al2O3 versus K2O.
Figure 5. Variation diagrams of: (a) CaO versus SiO2; (b) CaO versus Al2O3; and (c) Al2O3 versus K2O.
Minerals 07 00082 g005
Figure 6. Correlation coefficient matrix for the chemical components in the studied travertine.
Figure 6. Correlation coefficient matrix for the chemical components in the studied travertine.
Minerals 07 00082 g006
Figure 7. Bivariate plot of δ18O and δ13C (‰VPDB) values of the studied travertine samples.
Figure 7. Bivariate plot of δ18O and δ13C (‰VPDB) values of the studied travertine samples.
Minerals 07 00082 g007
Table 1. Summary of travertine and tufa lithofacies in the studied travertine outcrops.
Table 1. Summary of travertine and tufa lithofacies in the studied travertine outcrops.
LocalityLithofaciesDescription
Deir Alla & Suwayma TravertineCrystalline crustDense, crudely fibrous, elongated calcite crystals (ray crystals)
ShrubSmall bush-like growths, on horizontal-semi-horizontal surfaces
RaftsThin, delicate and brittle crystalline layers
LithoclastPenecontemporaneous, angular to subangular travertine fragments in different size, formed by erosion of the upper slope and collapse of waterfalls and terrace cliffs
Coated gas bubbleGas bubbles are coated by rapid precipitation of calcium carbonate
ReedRich in molds of reed and coarse grass
PebblyRounded pebbles of limestone, chert and basalt
PalaeosolExposure of travertine surface to rainwater, subaerial desiccation, and biological activities associated with soil formation
Az Zara TufaMacrophyte encrustationCalcified stems, twigs and leaves that represent rushes, reeds and unidentifiable bushes.
Bryophyte build-upsLaminated deposits that commonly form asymmetrical mounds or small domes, up to 1.75 m thick or stacked sequences up to 2 m high.
BiomicritesTabular layers, tens of centimeters thick, mostly at the top of the tufa succession. They consist of massive biomicrites or, in some places, biosparites.
MarlsMassive lenticular beds that are associated with macrophyte palisades. They contain plant debris.
Table 2. Calculated theoretical mineralogical composition of travertine and tufa samples (wt %).
Table 2. Calculated theoretical mineralogical composition of travertine and tufa samples (wt %).
No.D3D8D11D18D23bD26aD26bD29
Clay0.80.70.41.50.40.71.30.2
Dol0.50.70.50.70.40.90.50.4
Cc98.494.397.293.094.393.798.598.8
Qzndndnd0.10.10.1ndnd
Mgs0.40.40.30.50.30.80.40.3
Others2.44.31.84.74.84.41.80.7
No.S1dS10S19S27S32S34S38S39S45S47b
Clay9.22.24.51.41.21.81.81.32.22.8
Dol2.12.32.71.33.03.42.02.22.52.9
Cc72.090.788.994.492.390.793.893.190.787.8
Qz3.70.51.50.20.41.10.40.30.91.2
Mgs1.91.591.990.91.92.21.31.41.71.9
Others12.94.04.42.733.02.32.73.55.3
No.Z1Z1*Z3Z4
Clay0.80.51.01.1
Dol2.51.81.52.2
Cc91.888.793.594.2
Qz0.81.80.10.2
Mgs1.81.11.01.4
Others3.97.13.72.1
Z1 and Z1* are tufa samples. (nd: is not detected; D: Deir Alla; S: Suwayma; Z: Az Zara; Dol: dolomite; Cc: calcite and/or aragonite; Qz: quartz; Mgs: magnesite).
Table 3. X-ray Fluorescence results of major and minor elements of travertine and tufa samples.
Table 3. X-ray Fluorescence results of major and minor elements of travertine and tufa samples.
Chemical ConstituentsD3D8D11D18D23bD26aD26bD 29S1dS10S19
wt %
CaO55.3855.3855.8152.4254.152.9255.3955.4548.3854.3152.22
SiO20.22nd0.220.850.240.440.820.098.391.853.77
Al2O30.120.130.080.290.070.150.280.041.820.440.89
TiO2ndndnd0.05nd0.05ndnd0.28nd0.13
Fe2O30.281.30.82.133.810.240.40.554.720.471.83
MgO0.170.210.180.280.140.270.190.120.90.740.91
MnOndndnd0.030.031.93nd0.010.080.080.08
Na2Ondndnd0.040.040.13nd0.040.130.020.03
K2O0.01ndndndnd0.030.04nd0.850.080.23
P2O5dl0.01dl0.020.010.010.01dl0.010.190.08
SO30.020.020.02nd0.010.070.020.010.030.110.08
mg/kg
Sr8424235394189833391912007885111051898
Zr31ndndndndndndndndndnd
Mo307357349377285287378379389350321
Rbndndndndndndndnd87ndnd
Band447nd378nd2825ndnd10358339758
Clndndndndnd88nd401883112
Brndndndndndndndndndndnd
Chemical ConstituentsS27S32S34S38S39S45S47bZ1Z1*Z3Z4
wt %
CaO55.2955.2853.5754.2353.8451.9351.4152.5452.4555.4155.03
SiO20.891.041.951.2112.012.812.031.240.810.82
Al2O30.280.230.350.320.280.440.550.10.150.20.22
TiO2ndnd0.080.050.090.080.11ndndndnd
Fe2O30.830.470.870.911.810.533.430.880.180.170.47
MgO0.420.91.030.820.890.790.920.540.740.470.85
MnO0.030.04ndnd0.040.120.152.10.03ndnd
Na2Ondnd0.020.020.050.280.041.170.73nd0.02
K2O0.050.050.080.050.040.10.180.10.110.040.05
P2O5dl0.010.030.020.020.020.030.010.010.010.01
SO30.020.040.090.050.120.140.140.070.180.070.1
mg/kg
Sr5885551008802980181179872213897585837
Zrndndndndndndnd33ndndnd
Mo343234385410328373228378358308331
Rbndndndndnd48ndndndndnd
Ba547535834110,958101911591812nd29558221305
Clndnd24nd3923297877154545338
Brndndndndndndnd54134ndnd
Z1 and Z1* are tufa samples. nd: is not detected; dl: detection limit; D: Deir Alla; S: Suwayma; Z: Az Zara.
Table 4. Values of δ18O and δ13C in the studied samples.
Table 4. Values of δ18O and δ13C in the studied samples.
SampleFaciesδ18O (‰VPDB)δ13C (‰VPDB)
D3coated bubble8.171.88
D8paper-raft + reed−3.320.89
D18Shrub−5.48−1.45
D23bpaper-raft + reed−5.220.05
D26bcrystalline crust + palaeosol−4.932.99
D29crystalline crust−8.132.83
S1dlithoclast + pebbly−5.10.89
S10Lithoclast−4.251.01
S27Lithoclast−5.380.85
S32Lithoclast−2.841.03
S38Reed−3.58−0.18
S45coated bubble8.121.28
Z1macrophyte tufa−7.34−2.45
Z1*macrophyte tufa−7.02−1.78
Z3coated bubble travertine−8.83−3.52
Z4lithoclast + paper-raft travertine−5.14−1.34
VPDB: Vienna Pee Dee belemnite, standard value for C isotopes.

Share and Cite

MDPI and ACS Style

Ibrahim, K.M.; Makhlouf, I.M.; El Naqah, A.R.; Al-Thawabteh, S.M. Geochemistry and Stable Isotopes of Travertine from Jordan Valley and Dead Sea Areas. Minerals 2017, 7, 82. https://doi.org/10.3390/min7050082

AMA Style

Ibrahim KM, Makhlouf IM, El Naqah AR, Al-Thawabteh SM. Geochemistry and Stable Isotopes of Travertine from Jordan Valley and Dead Sea Areas. Minerals. 2017; 7(5):82. https://doi.org/10.3390/min7050082

Chicago/Turabian Style

Ibrahim, Khalil M., Issa M. Makhlouf, Ali R. El Naqah, and Sana’ M. Al-Thawabteh. 2017. "Geochemistry and Stable Isotopes of Travertine from Jordan Valley and Dead Sea Areas" Minerals 7, no. 5: 82. https://doi.org/10.3390/min7050082

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop