Hostname: page-component-7bb8b95d7b-pwrkn Total loading time: 0 Render date: 2024-09-29T17:12:43.414Z Has data issue: false hasContentIssue false

Estimation of Groundwater Residence Time Using Radiocarbon and Stable Isotope Ratio in Dissolved Inorganic Carbon and Soil CO2

Published online by Cambridge University Press:  29 April 2024

Rahul Kumar Agrawal
Affiliation:
Geosciences Division, Physical Research Laboratory Ahmedabad 380009, Gujarat, India Indian Institute of Technology, Gandhinagar, India
Ranjan Kumar Mohanty
Affiliation:
Geosciences Division, Physical Research Laboratory Ahmedabad 380009, Gujarat, India
Ajayeta Rathi
Affiliation:
Geosciences Division, Physical Research Laboratory Ahmedabad 380009, Gujarat, India Indian Institute of Technology, Gandhinagar, India
Shreya Mehta
Affiliation:
Geosciences Division, Physical Research Laboratory Ahmedabad 380009, Gujarat, India Indian Institute of Technology, Gandhinagar, India
M G Yadava
Affiliation:
Geosciences Division, Physical Research Laboratory Ahmedabad 380009, Gujarat, India
Sanjeev Kumar
Affiliation:
Geosciences Division, Physical Research Laboratory Ahmedabad 380009, Gujarat, India
Amzad H Laskar*
Affiliation:
Geosciences Division, Physical Research Laboratory Ahmedabad 380009, Gujarat, India
*
Corresponding author: Amzad H Laskar; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Estimation of residence time of groundwater, particularly in regions with inadequate surface waters are very important for formulating sustainable groundwater management policies. We developed a technique for extracting dissolved inorganic carbon (DIC) quantitatively from water for measuring its 14C contents and presented the analytical details here. We also measured stable carbon isotope ratio (δ13C) in soil CO2 and groundwater DIC to correct the groundwater 14C ages. In addition, 14C in soil CO2 were measured for making necessary correction in the initial activity of the recharging water. The corrected 14C contents in the groundwater samples were used to estimate their residence times employing Lumped Parameter Models (LPM), a set of mathematical models to account for the processes that take place during transport from the recharge to the sampling spots. We present a case study focused on the calculation of radiocarbon ages and residence times for a groundwater sample collected from the campus of Physical Research Laboratory in Ahmedabad, Gujarat, India. The study also includes estimations of groundwater residence times using previously measured 14C ages of groundwater samples from Gujarat, India. Various factors controlling the groundwater ages in the LPM and their applicability are discussed.

Type
Research Article
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of University of Arizona

1. Introduction

Detailed knowledge of groundwater residence time is crucial for groundwater management, especially in regions with arid, semi-arid, and sub-humid climates where surface waters resources are inadequate to meet the day-to-day demand. In addition, contamination and risk assessment depend strongly on the residence time of groundwater. Groundwater residence time in an aquifer is the average time taken by a water molecule to travel from the recharge to the exit or monitoring site of the aquifer. Radiocarbon (14C) in dissolved inorganic carbon (DIC) is widely used to understand groundwater dynamics including estimation of its residence time, recharge and flow rates (Clark and Fritz Reference Clark and Fritz1997; Kalin Reference Kalin, Cook and Herczeg2000; Hamed et al. Reference Hamed, Dassi, Ahmadi and Dhia2008; Iverach et al. Reference Iverach, Cendón, Meredith, Wilcken, Hankin, Andersen and Kelly2017; Priestley et al. Reference Priestley, Wohling, Keppel, Post, Love, Shand, Tyroller and Kipfer2017; Herrera et al. Reference Herrera, Godfrey, Urrutia, Custodio, Jordan, Jodar and Delgado2021). With the development of accelerator mass spectrometry (AMS), precise determination of the ratios of 14C/12C and 14C/13C in sub milligram of carbon made it a popular technique for dating water samples. 14CO2 gets incorporated into groundwater through dissolution of CO2 from the atmosphere and soil during recharge. Soil CO2 mainly governs the DIC concentration of the recharging water while passing through the unsaturated zones as CO2 concentration is very high in this zone due to root respiration and microbial decomposition of organic matter. Once water percolates down to the unsaturated zone, DIC becomes isolated from the modern 14C and decays with time. Therefore, measurement of 14C in groundwater can give an estimate of the age of the water. 14C contents can be measured with a precision better than 1%.

Soil CO2 is mainly controlled by decomposition of recent organic matter and root respiration and hence mostly modern in 14C content although decomposition or oxidation of old organic matter may contribute significantly in some specific cases (Magnone et al. Reference Magnone, Richards, van Dongen, Bryant, Evans and Polya2019). If a significant fraction of old carbon contributes to the DIC in recharge zone, the age of the groundwater will be overestimated. This needs to be ascertained and corrected wherever required. Contrarily, addition of post 1950 CE bomb carbon, released to the atmosphere due to nuclear weapon testing can lead to underestimation of the groundwater ages. Recent addition of carbon in groundwater DIC can be assessed measuring tritium in the same groundwater (Telloli et al. Reference Telloli, Rizzo, Salvi, Pozzobon, Marrocchino and Vaccaro2022). Potential 14C contamination particularly, incorporation of atmospheric CO2 in the aquifer and poor constraints on the sources of DIC in groundwater often leads to large uncertainty in 14C content (Plummer and Glynn Reference Plummer and Glynn2013; Han and Plummer Reference Han and Plummer2016; Campeau et al. Reference Campeau, Wallin, Giesler, Löfgren, Mörth, Schiff, Venkiteswaran and Bishop2017) and lowering in the upper limit of the estimated radiocarbon ages of groundwater to less than 30 ka (Kalin Reference Kalin, Cook and Herczeg2000; Clark Reference Clark2015), which is still good enough for most of the regional aquifers. Additionally, when aquitards store significant amount of water with different water transport rates and hence 14C content, estimated groundwater ages may deviate significantly depending on the ratio of the fluid volume in aquitards to aquifers (Bethke and Johnson Reference Bethke and Johnson2002). Therefore, precautions need to be taken to account the exchange with the aquitards while estimating age of groundwater.

The objectives of the present study are to develop a simple method for extracting CO2 from groundwater DIC and measuring its 14C contents and estimation of its residence time using various correction schemes and models. The correction schemes include estimating the contribution of DIC from carbonate rocks using δ13C in DIC and initial 14C correction in the recharging water by measuring the 14C in soil CO2, the main component of the groundwater DIC. Finally, the corrected 14C ages of the DIC were used to estimate groundwater residence time using Lumped Parameter Models (LPM) that take care of the advection and dispersion during transit from the recharge to the monitoring site.

2. Materials and Methods

2.1. Groundwater and Soil CO2 Sampling

For demonstrating the performances of the developed techniques for 14C and δ13C analysis in DIC and soil CO2, we carried out sampling and analysis of groundwater and soil CO2 from the campus of Physical Research Laboratory (PRL) (Figure 1). Groundwater was collected from a borewell sucking water from a depth of ∼90 m from the surface. Water samples were collected in 60 mL polypropiline narrow mouth bottles (Tarson, Kolkata, India) directly from a pipe sucking the groundwater before falling on the reservoir from which water is distributed. For soil CO2 collection, a homemade customized vacuum system with pumping arrangement was utilized as shown in Figure 2. A stainless-steel pipe (∼2 cm diameter) was introduced into the soil to the depth from which the soil air was to be collected. The soil material inside the pipe was removed and a new empty pipe was introduced into the hole. The empty spaces on the sides of the pipe were packed with soil materials to avoid contact with atmospheric air. The glass flask along with the pipeline was evacuated using a rotary vacuum pump. The flask was then filled to 1 atmosphere with the soil air from the desired depth, the pressure inside the flask was monitored using a pressure gauge. More details about the soil CO2 extraction system are provided elsewhere (Mohanty et al. in preparation).

Figure 1 Location for the groundwater sampling site, the campus of the Physical Research Laboratory (PRL). Locations of the sampling sites whose groundwater samples were dated previously by Agarwal et al. (Reference Agarwal, Gupta, Deshpande and Yadava2006) are also shown (the site number were taken from their published article). Samples with radiocarbon ages neither modern nor beyond radiocarbon dating limit are listed here.

Figure 2 Soil air sampling setup. The soil air samples are collected from a specific soil depth introducing the pipe to that depth and filling the pre-evacuated high vacuum glass flask.

2.2. CO2 Extraction from DIC of Groundwater

A schematic diagram for the extraction of DIC from water is shown in Figure 3. The extraction system was modified from some existing techniques (e.g., Carmi et al. Reference Carmi, Kronfeld, Yechieli, Yakir, Stiller and Boaretto2007; Macchia et al. Reference Macchia, D’Elia, Quarta, Gaballo, Braione and Maruccio2013; Ge et al. Reference Ge, Wang, Zhang, Luo and Xue2016; Takahashi et al. Reference Takahashi, Handa and Minami2021). About 10 mL commercial orthophosphoric acid (85%) was taken in a flask of ∼1 L capacity and was evacuated to ∼1 × 10–2 mbar. Using a syringe, 20–200 mL water was introduced into the flask. The amount of water for reaction depends on the concentration of DIC. The flask was heated to ∼60ºC using a hot plate and a magnetic stirrer was used for fast reaction and release of CO2. The reaction was allowed to proceed for 15–20 minutes. The liberated CO2 was passed through a series of three U-tubes with first one immersed in alcohol-liquid nitrogen slush kept at –80ºC and the subsequent two in liquid nitrogen (LN2) at –196ºC. The U-tube immersed in the slush, trapped most of the moisture while CO2 passed through it and freezes in the two LN2 traps. One of the LN2 traps was a spiral tube to increase the CO2 freezing efficiency (Figure 3). The extracted CO2 was cleaned by freeze and thaw technique a few times while transferring from one trap to the other to remove remaining traces of water and other impurities. The amount of CO2 liberated was measured by expanding it in a calibrated volume connected to a pressure gauge. Finally, pure CO2 was collected in the pre-evacuated high vacuum glass ampule to load into the graphite preparation system. The complete extraction of CO2 from a water sample takes almost an hour. The vacuum system takes an additional half an hour for cleaning by pumping between two samples.

Figure 3 Glass vacuum line used for extraction of carbon dioxide from dissolved inorganic carbon and soil CO2. Water samples are injected into the sample flask, reacted with orthophosphoric acid under vacuum, passed through tubes immersed in slush (−80°C) and liquid nitrogen (−196°C) and finally collected in the sample bottle.

2.3. Estimation of CO2 Yield from DIC

The liberated CO2 from the water samples was allowed to expand in a pre-calibrated volume attached to a pressure gauge (Setra DATUM 2000TM) to measure the CO2 content and hence DIC concentration in the sample. The gauge and attached volume were calibrated using pure Na2CO3 powder. Precisely weighed Na2CO3 was reacted with orthophosphoric acid in the flask at 60ºC for half an hour and the liberated CO2 was extracted following the procedure discussed above. The amount of Na2CO3 reacted varied from 2 to 53 mg. The liberated CO2 was expanded in the volume attached to the gauge and a calibration curve was prepared by plotting the amount of CO2 in micromole against the gauge reading (Figure 4). The amount of liberated CO2 in micromole was calculated from the actual weight of the Na2CO3 powder used in the reaction. The amount of CO2 liberated from any sample with unknown concentration can be estimated directly from the gauge reading using the equation shown in Figure 4.

Figure 4 Calibration plot used for estimating the amount of CO2 liberated from any sample.

2.4. Extraction of CO2 from Soil Air

CO2 from soil air for δ13C and 14C measurements was extracted using the same glass vacuum line used for extracting DIC (Figure 3). Instead of using water flask, the glass flask containing soil air was connected and CO2 was quantitatively extracted from the soil air using cryogenic technique. From the total amount of CO2 and the volume of the flasks, the concentration of CO2 was measured with a precision of ∼1% (Mohanty et al. in preparation).

2.5. Preparation of Standards

In absence of water standards for radiocarbon dating, we used carbonate standards for estimating the accuracy of the radiocarbon measurements for the present technique. The International Atomic Energy Agency’s two carbonate standards viz. Carrara Marble (IAEA-C1) and Travertine (IAEA-C2) were used. The Carrara marble serves as background and Travertine has a radiocarbon content of 41.14 ± 0.03 pMC (Rozanski Reference Rozanski1991). The carbonate samples were finely powdered with agate-mortar and dried samples were loaded in the reaction flask followed by evacuation. About 20–30 mL of Milli Q water was introduced into the flask using a syringe through a rubber septum under vacuum condition. The carbonate powder was mixed with water to make the extraction procedure similar to that of water samples. About 5 mL orthophosphoric acid was then injected into the flask using a syringe. The CO2 evolved was extracted following the same procedure as discussed for the water samples. This gives the estimate of the background and accuracy of the procedure followed for radiocarbon measurements in DIC samples.

2.6. Graphite Preparation and 14C Dating Using AMS

The purified CO2 samples were converted into graphite using an in-house graphite preparation line (Figure 5). About 50 to 100 micromole CO2 was transferred to a vacuum system in which pre-conditioned iron (∼5 mg) and zinc (∼25 mg) powders were loaded in the two arms of an L-shaped interconnected vacuum sealed quartz tubes (Jull et al. Reference Jull, Donahue, Hatheway, Linick and Toolin1986). Iron and zinc powders were heated to appropriate temperatures for CO2 reduction to elemental carbon (graphite) which deposit on the surface of iron powder. Graphite precipitation using this technique takes 3 to 5 hours. The graphite-coated iron was then pressed into pellet and loaded into the source of the 1 MV accelerator mass spectrometer for 14C measurements at Physical Research Laboratory (PRL) Ahmedabad, India (Bhushan et al. Reference Bhushan, Yadava, Shah and Raj2019). Using 14C/12C or 14C/13C for samples, ages were estimated (Donahue et al., Reference Donahue, Linick and Jull1990). Along with the samples, measurements of international standards such as Oxalic Acid I, Oxalic Acid II and background materials (anthracite) were made to estimate the background subtracted activity of samples and measure their radiocarbon ages. The ages were corrected for the isotopic fractionation during sample preparation and within the AMS (Donahue et al. Reference Donahue, Linick and Jull1990).

Figure 5 Graphite preparation system. Fe: iron powder, Zn: zinc powder, PG: pressure gauge, VG: vacuum gauge, P: pump, SB: sample bottle to load CO2 into the system.

2.7. Radiocarbon Dating in Groundwater

Conventionally radiocarbon age (using Libby half-life of 5568 yr) of any sample is calculated using the basic decay equation given by

(1) $$t = - {{5568} \over {\ln (2)}}\ln \left( {{A \over {{A_0}}}} \right)$$

Where A is the activity measured in a sample and A 0 is the initial activity. These activities are corrected for isotopic fractionation also called normalized activities (Stuiver and Polach Reference Stuiver and Polach1977) “Closed system” is a basic requirement for radiocarbon age determination i.e., the system should remain closed to subsequent gains or losses of the carbon except through radioactive decay. However, most of the natural samples particularly subsurface water samples experience many physical and chemical processes and mixing with many different water channels during transport and within the aquifers altering the radiocarbon contents (Plummer and Glynn, Reference Plummer and Glynn2013 and references therein). Decomposition of old organic matter or dissolution of carbonates may cause incorporation of DIC into groundwater in addition soil CO2. These two components might have different 14C content compared to the original groundwater. In such cases actual 14C content present in a groundwater sample can be different from that attained just by decay of the original 14C content of the sample. Though, there is no standard way to correct for the incorporation of carbon from the surroundings, some schemes for radiocarbon age corrections in groundwater samples are discussed below.

If DIC in groundwater is solely derived from atmospheric CO2 and/or the modern soil CO2 and no mixing with other flow paths or incorporation of carbon from other sources take place, the determined age will be close to its exact value. It is to be noted that the determined age should be corrected for age of the soil CO2 particularly when the ages of CO2 in the unsaturated zone are significantly old or if a significant fraction of carbon is originated from the post 1950 CE atmosphere (bomb carbon). For example, 14C content of the soil CO2 at Ti Tree (Australia) at a depth of 10 m is ∼50 pMC though it is close to 100 pMC near the surface (Wood et al. Reference Wood, Cook and Harrington2015). Low 14C contents in soil CO2 in deeper soil layers were also observed at Saskatoon (Canada), North Dakota and Navada (USA) (Bacon and Keller Reference Bacon and Keller1998; Thorstenson et al. Reference Thorstenson, Weeks, Haas and Fisher2016; Haas et al. Reference Haas, Fisher, Thorstenson and Weeks2016). On the other hand, presence of bomb carbon in unsaturated zone CO2 or in organic matter (e.g., Blume et al. Reference Blume, Guitard, Crann, Orekhov, Amos, Clark, Lapen, Blowes, Ptacek, Craiovan and Sunohara2022; Laskar et al. Reference Laskar, Yadava and Ramesh2012) may lead to underestimation of groundwater ages if not corrected. In most of the studies related to the groundwater dating, 14C content in CO2 in the unsaturated zone is assumed to be modern which may lead to significant discrepancies between the actual and measured ages. The 14C contents are also expressed as percent modern carbon pMC = (A/A0)100% or in the form of delta notation such as D14C = (A/A0–1)1000‰ and or fraction modern F14C = A/A0. A sample with pMC greater than 100% or positive D14C or F14C greater than 1 indicate presence of bomb carbon.

Incorporation of carbon with different 14C signature in groundwater DIC pool in some cases can be corrected using stable carbon isotope ratio (δ13C) in DIC as follows. Average δ13C values of C3 and C4 vegetation are ∼ –27‰ and –13‰, respectively (Deines Reference Deines1980; Khon Reference Kohn2010). The organic carbon present in soil under a particular vegetation type has similar or slightly enriched δ13C values (Laskar et al. Reference Laskar, Sharma, Ramesh, Jani and Yadava2010, Reference Laskar, Yadava, Sharma and Ramesh2013, Reference Laskar, Yadava and Ramesh2016). The enrichment is due to incorporation of 13C depleted CO2 in the atmosphere during the industrial era (Suess Effect) (McCarrol and Loader Reference McCarroll and Loader2004; Paul et al. Reference Paul, Balesdent and Hatté2020) and preferential removal of lighter carbon during decomposition by microbes (Laskar et al. Reference Laskar, Yadava and Ramesh2016). Soil CO2 has a δ13C value controlled by the decomposition of organic matter, root respiration and diffusive fractionation. As soil CO2 mainly controls the DIC contents in the groundwater, its δ13C value can be used to estimate its fraction in DIC taking into account the appropriate fractionation factor (Szaran Reference Szaran1997). Therefore, by measuring δ13C values in soil CO2 in recharge region and in DIC, contribution from other sources (bedrock) can be estimated and corrected for the 14C activity in the sample. If 14C dilution is caused by dissolution of carbonate of marine origin, the correction can be relatively easily made using δ13C values as these carbonates are free from 14C and their δ13C values are 0 ± 2‰ (Ripperdan Reference Ripperdan2001) which is very different from soil CO2 (∼ –20‰). On the other hand, if dissolution takes place from soil carbonates formed by evaporation or lacustrine carbonates it may be difficult to estimate the contributions simply using δ13C of DIC due to weak constraints and possible overlapping in the range of their δ13C values with that dissolved from soil CO2. In addition, groundwater does not flow like the flow in a tube (piston flow), but encounter dispersion, advection and mixing making it a more complex system. Therefore, appropriate models are required to handle those mixing issues as discussed in Section 2.8.

A simple approach for correction of 14C age in groundwater due to contamination from carbonate dissolution and decomposition of old organic matter is as follows. Suppose a groundwater sample is diluted due to incorporation of carbon from bedrocks and q is the dilution factor (see Equation 3 below), then the 14C activity in the recharge area would be qAo, and the corresponding radiocarbon age of such a groundwater, as discussed elsewhere (e.g., Clark and Fritz Reference Clark and Fritz1997; Aggarwal et al. Reference Agarwal, Gupta, Deshpande and Yadava2006; Cartwright Reference Cartwright, Currell, Cendón and Meredith2020) would be

(2) $$t = - {{5568} \over {\ln (2)}}\ln \left( {{A \over {q{A_0}}}} \right)$$

Unfortunately, determination of the dilution factor q is challenging in most of the cases. In case of dilution by marine carbonate under closed condition, the dilution factor q can be calculated using the following two-component isotopic mass balance equation:

(3) $${\rm{{\delta ^{13}}{C_{DIC}} = {\rm{ }}{\delta ^{13}}{C_{carb}} \times q + {\delta ^{13}}{C_{recharge}} \times \left( {1-q} \right)}}$$

where δ13CDIC, δ13Crecharge and δ13Ccarb are the δ13C values observed in the DIC in the groundwater, DIC of recharging water and carbonate rocks in the aquifers, respectively. δ13Crecharge can be obtained from the δ13C value of soil CO2 as δ13C in bicarbonate (most of the DIC at neutral pH) is ∼7.4‰ higher than δ13C value of the soil CO2 at 30ºC (Vogel et al. Reference Vogel, Grootes and Mook1970; Mook et al. Reference Mook, Bommerson and Staverman1974; Deines et al. Reference Deines, Langmuir and Harmon1974; Szaran Reference Szaran1997). In the above isotopic mixing calculation (Equation 3), contribution from atmospheric CO2 is neglected because of the fact that most of the dissolution of CO2 in groundwater takes place in the unsaturated zone due to high concentration of CO2 in the latter. Soil CO2 from decomposition/oxidation of old organic matter if significant can be corrected using the 14C in soil CO2 in the recharge areas. The values of q can also be calculated from major ion geochemistry, graphical techniques and isotopic mass balance as discussed by many previous researchers (e.g., Vogel and Ehhalt Reference Vogel and Ehhalt1963; Vogel et al. Reference Vogel, Grootes and Mook1970; Pearson and Hanshaw Reference Pearson and Hanshaw1970; Mook Reference Mook1972; Tamers Reference Tamers1975; Fontes and Garnier Reference Fontes and Garnier1979; Clark and Fritz Reference Clark and Fritz1997; Geyh Reference Geyh2000; Agarwal et al. Reference Agarwal, Gupta, Deshpande and Yadava2006; Coetsiers and Walraevens Reference Coetsiers and Walraevens2009; Blaser et al. Reference Blaser, Coetsiers, Aeschbach-Hertig, Kipfer, Camp, Loosli and Walraevens2010; Clark Reference Clark2015; Han and Plummer Reference Han and Plummer2016; Cartwright et al. Reference Cartwright, Currell, Cendón and Meredith2020).

There can be more potential sources of DIC with different 14C content compared to the soil CO2 such as soil carbonate, carbonate nodules and addition of bomb carbon in groundwater which can alter the q value. Addition of such carbon in DIC would result in over or underestimation of groundwater residence time if not corrected. Estimation of the input of 14C-free DIC or any input into the DIC pool of an aquifer with different initial 14C content than the aquifer water is a major challenge in interpreting radiocarbon data in groundwater (Vogel and Ehhalt Reference Vogel and Ehhalt1963; Fontes Reference Fontes1992; Aravena et al. Reference Aravena, Wassenaar and Plummer1995; Kalin Reference Kalin, Cook and Herczeg2000; Coetsiers and Walraevens Reference Coetsiers and Walraevens2009; Han et al. Reference Han, Plummer and Aggarwal2012; Plummer and Glynn Reference Plummer and Glynn2013; Han and Plummer Reference Han and Plummer2016; Nydahl et al. Reference Nydahl, Wallin, Laudon and Weyhenmeyer2020; Cartwright et al. Reference Cartwright, Currell, Cendón and Meredith2020). As shown in Equation 3, for a two-component mixing, values of q may be calculated from the δ13C values of soil CO2, DIC in groundwater and carbonate rocks (Pearson and Hanshaw Reference Pearson and Hanshaw1970; Tamers Reference Tamers1975; Clark and Fritz Reference Clark and Fritz1997; Geyh Reference Geyh2000; Coetsiers and Walraevens Reference Coetsiers and Walraevens2009; Clark Reference Clark2015; Han and Plummer Reference Han and Plummer2016).

The ages estimated using decay equations (1 and 2) need to be converted into calendar ages using radiocarbon calibrations (e.g., Reimer et al. Reference Reimer, Austin, Bard, Bayliss, Blackwell and Bronk Ramsey2020; Hogg et al. Reference Hogg, Heaton, Hua, Palmer and Southon2020). Calibration is necessary to compare the obtained ages with other paleoclimate datasets. Other factors (e.g., hard-water effect) may cause erroneous estimation of radiocarbon ages and need to be carefully accounted before calibrations. The calibrated ages of groundwater can be considered as the residence times in some limited cases if all the water in an aquifer is recharged at the same time without much dispersion and advection during its transport. Hardly such simple systems exist as groundwater flows along paths of varying lengths and undergoes hydrodynamic dispersion and advection causing 14C content modification. In other words, a groundwater sample may contain many fractions of water from different origins and have a range of residence times rather than being of a single age (Maloszewski and Zuber Reference Maloszewski and Zuber1982; Maloszewski Reference Maloszewski2000; Cox Reference Cox2003; Suckow Reference Suckow2014).

2.8. Estimation of Residence Time of Groundwater Using Radiocarbon

There is no standard way to account all the complexities discussed above. Researchers address these complexities using different models (Seltzer et al. Reference Seltzer2021; Starn et al. Reference Starn, Kauffman, Carlson, Reddy and Fienen2021; Fareze et al. Reference Fareze, Rajaobelison, Ramaroson, Razafitsalama and Rakotomalala2022). One approach is to use Lumped Parameter Models (LPM) (Maloszewski and Zuber Reference Maloszewski and Zuber1982; Maloszewski and Zuber Reference Maloszewski and Zuber1992; Maloszewski Reference Maloszewski2000; McGuire and McDonnell Reference McGuire and McDonnell2006; Jurgens et al. Reference Jurgens, Böhlke and Eberts2012; Cartwright et al. Reference Cartwright, Currell, Cendón and Meredith2020) to estimate the groundwater residence times. LPM are mathematical models that deal with the transport of water from recharge zones to discharge regions based on aquifer geometry and flow configurations. The model’s approach involves convolution of tracer input and decay functions with a weight factor and the obtained output is matched to a known time series (Maloszewski and Zuber Reference Maloszewski and Zuber1982; Maloszewski et al. Reference Maloszewski, Stichler and Zuber2004). The models allow the variability of the 14C input, define a recharge zone, time taken by the water to penetrate the soil zone and the ratios of advection to dispersion. In other words, the model assumes more realistic and flexible flow path geometries, and the effects of dispersion may be taken into account to a large extent.

In the LPM, at steady state, the 14C content of groundwater 14C (t) at a time t in any outlet position in the aquifer can be calculated from the corrected 14C content of DIC in the recharging water (14Ci) using exit age distribution function and the radioactive decay via the following convolution integral (Maloszewski and Zuber, Reference Maloszewski and Zuber1982):

(4) $${}^{14}{C_t} = \int\limits_0^\infty {{C_i}(t - \tau )} g(\tau )\exp ( - \lambda \tau )d\tau $$

Where τ is the mean residence time, t- τ is the time when the water recharged, λ is the decay constant and g(τ) is the response function describing the distribution of flow paths and residence times in the aquifer. Depending on the aquifer geometries, different LPMs can be applied (e.g., Maloszewski and Zuber, Reference Maloszewski and Zuber1982; Maloszewski and Zuber Reference Maloszewski and Zuber1992; Amin and Campana, Reference Amin and Campana1996; Cook and Bohlke Reference Cook, Bohlke, Cook and Herczeg2000; Maloszewski Reference Maloszewski2000; Jurgens et al. Reference Jurgens, Böhlke and Eberts2012).

Here we used a tracer LPM program developed by Jurgens et al. (Reference Jurgens, Böhlke and Eberts2012) to estimate the residence times of groundwater extracted from the PRL Campus. We also applied the LPM program on previously measured 14C ages in groundwater from different parts of Gujarat, Western India. We applied three different LPMs viz. piston flow model (PFM), dispersion model (DM) and binary mixing model (BMM) to show the variations in residence times under different dispersion and mixing parameters. Schematics of the PFM and DM are shown in Figure 6.

Figure 6 Schematics of (A) piston flow model (after Bethke and Johnson, Reference Bethke and Johnson2008) and (B) dispersion model (after Maloszewski and Zuber Reference Maloszewski and Zuber1982).

The PFM assumes that a tracer (14C) travels from the recharge area to the outlet position (e.g., a well) without mixing or hydrodynamic dispersion. For PFM, radiocarbon output for a constant input (14Ci), the solution of Equation (4) can be expressed as (Zuber and Maloszewski Reference Maloszewski2000):

(5) $${C_{\left( t \right)}} = {C_i}\left( {t - \tau } \right){\rm{exp}}\left( { - {\rm{\lambda \tau }}} \right)$$

The PFM is applicable to hydrogeologic settings with low dispersion, high linear velocity or short flow path from recharge to discharge region. 14C tracer can be assumed to follow piston-flow behaviour in shallow, short screened monitoring wells in unconfined aquifers or in confined aquifers with a relatively small recharge area.

The DM is based on a solution of a one dimensional advection dispersion equation for a semi-infinite medium with an instantaneous injection and detection of the tracer in the fluid flux. A solution for the Equation (4) for a constant input (14Ci) with advection and dispersion can be expressed as (Zuber and Maloszewski Reference Maloszewski2000):

(6) $${\rm{^{14}{C_t}{ = ^{14}}{C_i}\ exp\left[ {{{\left( {2{P_D}} \right)}^{-1}} \times \{ 1 - {{\left( {1 + 4\lambda {P_D}\tau } \right)}^{1/2}}\} } \right]}}$$

Where PD is the dispersion parameter which is the reciprocal of Peclet number describing the relative importance of dispersion and advection. The Peclet number depends on the velocity of flow field and characteristics length of the system (Rapp Reference Rapp2017). The DM can give an approximate description of age distributions in samples from multiple aquifer configurations.

Binary mixing model (BMM) describes two-component mixtures in which each component can be obtained by a model. For example, a “BMM-PFM-DM” model describes a binary mixture in which one component of the mixture is modelled by using PFM (14C1) and the other component by DM (14C2) and the corresponding fractions are f1 and f2(=1-f1). The output in a BMM is given by:

(7) $${\rm{^{14}{C_{out}} = {f_1}^{14}{C_1} + {\left( {1 - {f_1}} \right)^{14}}{C_2}}}$$

Binary mixing models can be appropriate for wells screened across multiple aquifer units and aquifers with short-circuit pathways that result in age mixtures of significantly different mean ages.

3. Results and Discussion

3.1. Yield, Accuracy, and Reproducibility

The average yield of the DIC extraction of the present method is 98 ± 5% (Table 1). The accuracy of measurements is estimated with IAEA standards Carrara Marble and Travertine (Table 2). Carrara marble is free of 14C and acts as a background while Travertine has a consensus radiocarbon content of 41.14±0.03 pMC. These carbonate standards are solid samples, but our actual samples are liquid. To make the processing of these solid standards identical with liquid samples, these samples were crushed and mixed with Milli Q water and then reacted with orthophosphoric acid. The measured 14C contents of Travertine is similar to the consensus value within the analytical uncertainty but it is slightly higher for the Carrara marble (background). The higher 14C content for the later is probably due to incorporation of air CO2 into the MilliQ water, used with the carbonate powder for making the CO2 extraction identical with the water samples. The effect of incorporation of air CO2 significantly modified the 14C content of the background sample (Carrara marble) but insignificant for relatively younger samples. The actual groundwater samples are expected to have less air CO2 contamination compared to these carbonate standards as they were carefully handled without exposing to the air.

Table 1 Yield of DIC extraction, measured using the present technique. Water samples with different DIC concentration are prepared by dissolving known amount of sodium carbonate. The solutions are then reacted with orthophosphoric acid and the amount of liberated CO2 were measured. The measured amounts of CO2 for different samples are compared to that actually present and the percentage yields are calculated

Table 2 Radiocarbon dates of two international standards measured using the present technique (details of the standards are available in Rozanski et al. Reference Rozanski1991)

The reproducibility of 14C measurements for the present technique has been estimated using multiple measurements of the groundwater samples collected from the PRL Campus. Radiocarbon age of this groundwater are presented in Table 3. The reproducibility of the radiocarbon age (standard deviation) for five PRL groundwater samples is ∼160 yr (Table 3).

Table 3 Radiocarbon ages of PRL groundwater samples (with Libby half-life). A δ13C value of soil CO2 of –19.4‰ (average in the soil profile, see Table 4) is considered for the calculation. The median of the calibrated age is used for estimating mean residence time

* Value of (1–q) in Equation (3).

Correction is done using Equation (2).

3.2. Radiocarbon Age of PRL Groundwater

Groundwater samples collected from the PRL Campus were analyzed for δ13C and 14C and are presented in Table 3. To constrain the contribution of soil CO2 in DIC, we measured the δ13C values in soil CO2 up to a depth of 90 cm inside the PRL campus. δ13C values and 14C ages of the soil CO2 are presented in Figure 7 and Table 4, respectively. Within the soil profile, δ13C varies between –18.7 and –20.3‰ with an average of –19.4 ±0.4‰. Considering the fractionation factor of 7.4‰ at 30ºC (Szaran, Reference Szaran1997; Deines et al., Reference Deines, Langmuir and Harmon1974) between the CO2 and dissolved bicarbonate (dominant DIC species at a pH of ∼7), δ13C values in the latter is expected to be –12.1 ± 0.4‰. However, δ13C values of the present groundwater varies between –8.8 to –9.1‰ indicating that the groundwater samples are significantly contaminated by carbon from other sources. We assume that the δ13C value in soil CO2 measured inside the PRL campus is the same as that in the recharge region as we do not see much spatial variation in the δ13C values of soil CO2 across Gujarat (Mohanty et al. in preparation). Assuming bedrock carbonates as the other contributor to the DIC (hard-water effect) with a δ13C value of ∼0 ± 2‰ (Ripperdan Reference Ripperdan2001) and free from 14C, we correct its contribution to the radiocarbon ages calculating the factor q (Equation 3, Table 3). The factor q can also be affected by dissolution of soil carbonate from shallow zones with 14C content neither zero nor modern (e.g., Solder and Jurgens Reference Solder and Jurgens2020) though the deep soil carbonates are mostly free from 14C. The contribution from soil carbonates and carbonate nodules are neglected in the present case as the contribution from dissolution of such carbonates in alluvial settings with relatively fast recharge is expected to be low. However, this needs to be tested measuring 14C content in soil DIC and soil carbonate. To validate the other assumption that the soil CO2 in the recharge zones that dissolves to from DIC, is modern, we measured the 14C content in soil CO2 at different depths inside PRL campus and found F14C values in the range of 0.97 to 1 fraction modern or age from 0 to 210 BP with an average of 0.99 fraction modern or 90 BP in the top 90 cm of the soil profile (Table 4). Therefore, the 14C content of the soil CO2 in the recharge zone is mostly modern hence neglected in calculating the age of the groundwater.

Figure 7 Variation in δ13C in soil CO2 with depth in a soil profile inside the Campus of Physical Research Laboratory Ahmedabad.

Table 4 Stable carbon isotopic composition (δ13C) and radiocarbon content (F14C) in soil CO2 at different depths in a soil profile from the Campus of Physical Research Laboratory. The reproducibility in δ13C measurement is ∼0.1‰. 2 to 3 samples from the same depths (different samples) for some samples were collected, processed and analyzed for establishing the reproducibility of the method. Radiocarbon ages in BP are also presented

* Not measured.

The measured radiocarbon age of the PRL groundwater is 5490 ± 160 BP (Table 3). This age is corrected using the measured δ13C value of the DIC and calculating the factor q (Equation 3 and Table 3). The corrected age is 2980 ± 110 BP. This age is calibrated using OxCal version 4.4. The calibrated age of the PRL groundwater lies in the range of 2855 to 3442 cal BP with a median value of 3140 cal BP. The correction has been made using Equation (3). The mean residence times, calculated using the median of the calibrated age range are 3240 and 3360 yr, estimated by PFM and DM, respectively. For the DM, a dispersion parameter of 0.1 is assumed for the residence time calculation. This dispersion parameter corresponds to a Peclet number of ∼10 (see Section 3.3). A model estimated residence time is not a single value but a range with varying probability and the mean value is reported here.

3.3. Variation in Residence Times of Groundwater with Varying Radiocarbon Ages

Variation in residence times of groundwater with variation in radiocarbon ages for PFM, DM and BMM are shown in Figure 8. The actual radiocarbon ages and the residence times obtained from PFM are not very different as in this model it is assumed that a water parcel once detached from the soil zone directly travels to the target site without significant advection and/or dispersion. However, water before reaching the target site of an aquifer, may encounter dispersion and advection causing significant change in its radiocarbon content which need to be taken into account before estimating its residence time. In the DM, these processes along the transport path and aquifer are taken into account by the dispersion parameter which is given by the inverse of Peclet number (Zuber and Maloszewski Reference Maloszewski2000). Peclet number for recharge patterns, topography and aquifer set up for the present case would vary in the range of 10–20 as discussed in Wilske et al. (Reference Wilske, Suckow, Mallast, Meier, Merchel, Merkel, Pavetich, Rödiger, Rugel, Sachse, Weise and Siebert2020) giving a dispersion parameter in the range of 0.05–0.1. For this range of dispersion parameters, the variation in the residence times of groundwater with radiocarbon ages are shown in Figure 8.

Figure 8 Estimated residence times with change in radiocarbon ages using piston flow model (PFM), dispersion model (DM) with a dispersion parameter of 0.1 and binary mixing model (BMM). Residence times corresponding to dispersion parameters of 0.05 and 0.1 for DM are shown. In the BMM, calculation is made with 10% contribution from PFM and 90% from DM.

We also presented residence times estimated using a BMM with 10% contribution from PFM and 90% from DM with a dispersion parameter of 0.1. The BMM gives residence times intermediate of the PFM and DM (Figure 8). For radiocarbon ages of less than ∼5000 BP, the difference between the residences time estimated by various models is not significant and hence PFM or even the calibrated radiocarbon ages with proper corrections can represent the residence time. With increase in age, the deviation in the estimations of residence times increases compared to the radiocarbon ages or that estimated by PFM and hence it is better to use DM or other more appropriate models for older samples as recommended previously (e.g., Jurgens et al. Reference Jurgens, Böhlke and Eberts2012).

3.4. Residence Times of Groundwater with Available 14C Dates from Gujarat, Western India

We also estimated the residence times of some groundwater samples from Gujarat which were dated previously using 14C technique (Agarwal et al. Reference Agarwal, Gupta, Deshpande and Yadava2006). The locations of the sampling sites are given in Figure 1. The original radiocarbon ages and the residence times obtained using PFM and DM are presented in Figure 9. The PFM gives residence times close to that of the radiocarbon ages. However, the residence times deviates significantly with the DM especially for older samples. The residence times for DM is estimated using a dispersion parameter of 0.1. The groundwater sample from PRL campus whose residence times is estimated here was also dated by Agarwal et al. (Reference Agarwal, Gupta, Deshpande and Yadava2006) and estimated a radiocarbon age of 7580 BP which is significantly higher than that measured in this study (Table 3). The difference is probably due to sampling from different aquifers or increased addition of modern waters due excessive extraction of groundwater during the last two decades.

Figure 9 Residence times of the groundwater samples of Gujarat region using available radiocarbon dates from Agarwal et al (Reference Agarwal, Gupta, Deshpande and Yadava2006) (see Figure 1 for locations). Residence times estimated using piston flow model (PFM) and Dispersion Model (DM) with a dispersion parameter of 0.1 are shown along with the measured radiocarbon ages for different samples.

The advantage of the present technique of CO2 extraction from DIC is simple, quantitative and does not require career gas to extract the liberated CO2. The major limitations of estimating residence time using 14C in DIC is the poor constraints of the sources of DIC and exchange of waters between aquifers and aquitards. Further investigations with measurements of more parameters particularly to better constrain the carbonate dissolution would help to estimate the 14C contents of the groundwater DIC. Also, better mapping of the water contents in aquitards can better constrain the aquifer water age and hence groundwater residence time.

Conclusions

We presented a simple CO2 extraction system from DIC for radiocarbon dating of groundwater samples and estimated the 14C ages of groundwater from the campus of Physical Research Laboratory Ahmedabad. δ13C values and 14C ages of soil CO2 up to a depth of 90 cm were measured to estimate the contribution of soil CO2 in the DIC and age of the DIC in the recharge zone. δ13C values of the DIC were used to estimate the contribution of bedrock carbonates in groundwater and hence 14C age correction. Soil CO2 was found to be modern and hence the influence of age of it in the recharge zone in modulating the radiocarbon ages of groundwater was neglected. Residence times estimated using different LPM for a groundwater sample collected from the PRL Campus and some previously measured radiocarbon ages from Gujarat region were presented and discussed. Groundwater residence times can be estimated reasonably well using Piston Flow Model for young samples (<5000 yr) while for older samples, Dispersion Model or other more appropriate models should be used.

Acknowledgments

This work was funded by the Science and Engineering Research Board of Department of Science and Technology, India, grant number SRG/2021/001338 to AHL. AHL thanks Mr. A. Shivam for measuring 14C contents of the samples in AMS.

References

Agarwal, M, Gupta, SK, Deshpande, RD, Yadava, MG. 2006. Helium, radon and radiocarbon studies on a regional aquifer system of the North Gujarat–Cambay region, India. Chemical Geology 228(4):209232.CrossRefGoogle Scholar
Amin, IE, Campana, ME. 1996. A general lumped parameter model for the interpretation of tracer data and transit time calculation in hydrologic systems. J. Hydrol. 179:121. https://doi.org/10.1016/0022-1694(95)02880-3 CrossRefGoogle Scholar
Aravena, R, Wassenaar, LI, Plummer, LN. 1995. Estimating 14C groundwater ages in a methanogenic aquifer. Water Resour. Res. 31:23072317. https://doi.org/10.1029/95WR01271 CrossRefGoogle Scholar
Bacon, DH, Keller, CK. 1998. Carbon dioxide respiration in the deep vadose zone: implications for groundwater age dating. Water Resour. Res. 34:30693077. https://doi.org/10.1029/98WR02045 CrossRefGoogle Scholar
Bethke, CM, Johnson, TM. 2002. Paradox of groundwater age. Geology 30(2):107110.2.0.CO;2>CrossRefGoogle Scholar
Bethke, CM, Johnson, TM. 2008. Groundwater age and groundwater age dating. Annu. Rev. Earth Planet. Sci. 36: 121152. doi: 10.1146/annurev.earth.36.031207.124210.CrossRefGoogle Scholar
Bhushan, R, Yadava, MG, Shah, M, Raj, H. 2019. Performance of a new 1MV AMS facility (AURiS) at PRL, Ahmedabad, India. Nuclear Inst. and Methods in Physics Research B 439:7679.CrossRefGoogle Scholar
Blaser, PC, Coetsiers, M, Aeschbach-Hertig, W, Kipfer, R, Camp, MV, Loosli, HH, Walraevens, K. 2010. A new groundwater radiocarbon correction approach accounting for palaeoclimate conditions during recharge and hydrochemical evolution: the Ledo-Paniselian Aquifer, Belgium. Applied Geochemistry 25:437455.CrossRefGoogle Scholar
Blume, O, Guitard, E, Crann, C, Orekhov, M, Amos, R, Clark, I, Lapen, D, Blowes, D, Ptacek, C, Craiovan, E, Sunohara, M. 2022. Relationships between carbon age and CO2 efflux in agricultural and drainage ditch soils using the thermonuclear bomb pulse. Vados Zone J. 21:e20208.CrossRefGoogle Scholar
Campeau, A, Wallin, MB., Giesler, R, Löfgren, S, Mörth, CM, Schiff, S, Venkiteswaran, JJ, Bishop, K. 2017. Multiple sources and sinks of dissolved inorganic carbon across Swedish streams, refocusing the lens of stable C isotopes. Sci Rep 7:9158. https://doi.org/10.1038/s41598-017-09049-9 CrossRefGoogle ScholarPubMed
Carmi, I, Kronfeld, J, Yechieli, Y, Yakir, D, Stiller, M, Boaretto, E. 2007. quantitative extraction of dissolved inorganic carbon (as CO2) and water by vacuum distillation from sediments of the unsaturated zone for carbon isotope analysis (13C and 14C). Radiocarbon 49(1):83–94.CrossRefGoogle Scholar
Coetsiers, M, Walraevens, K. 2009. A new correction model for 14C ages in aquifers with complex geochemistry—application to the Neogene Aquifer, Belgium. Appl. Geochem. 24:768776.CrossRefGoogle Scholar
Cartwright, I, Currell, MJ, Cendón, DI, Meredith, KT. 2020. A review of the use of radiocarbon to estimate groundwater residence times in semi-arid and arid areas, Journal of Hydrology 580:124247. https://doi.org/10.1016/j.jhydrol.2019.124247 CrossRefGoogle Scholar
Clark, ID. 2015. Groundwater geochemistry and isotopes. Boca Raton, FL: CRC Press.CrossRefGoogle Scholar
Clark, ID, Fritz, P. 1997. Environmental isotopes in hydrogeology. Boca Raton, London, New York: CRC Press, Taylor & Francis Group.Google Scholar
Coetsiers, M, Walraevens, K. 2009. A new correction model for 14C ages in aquifers with complex geochemistry—application to the Neogene Aquifer, Belgium. Appl. Geochem. 24:768776. https://doi.org/10.1016/j.apgeochem.2009.01.003 CrossRefGoogle Scholar
Cook, PG, Bohlke, JK. 2000. Determining timescales for groundwater flow and solute transport. In: Cook, PG, Herczeg, AL, editors. Environmental tracers in subsurface hydrology. Kluwer, Boston. p. 130.CrossRefGoogle Scholar
Cox, SE. 2003. Estimates of residence time and related variations in quality of ground water beneath Submarine Base Bangor and Vicinity, Kitsap County, Washington. US Geological Survey Water Resources Investigation Report 03-4058.Google Scholar
Donahue, DJ, Linick, TW, Jull, AJT. 1990. Isotope-ratio and background corrections for accelerator mass spectrometry radiocarbon measurements. Radiocarbon 32(2):135142. https://doi.org/10.1017/S0033822200040121 CrossRefGoogle Scholar
Deines, P. 1980. The isotopic composition of reduced organic carbon. In: Fritz P, Fontes JC, editors. Handbook of environmental isotope geochemistry. Vol. 1. The terrestrial environment. 3rd edition. Amsterdam: Elsevier.CrossRefGoogle Scholar
Deines, P, Langmuir, D, Harmon, RS. 1974. Stable carbon isotope ratios and the existence of a gas phase in the evolution of a carbonate groundwaters. Geochim. Cosmochim. Acta 38:11471164.CrossRefGoogle Scholar
Fareze, LP, Rajaobelison, J, Ramaroson, V, Razafitsalama, FFA, Rakotomalala, CU. 2022. Groundwater Residence Time Modelling in the Aquifers of the Ampasindava Peninsula, Northwestern of Madagascar. Open Access Library Journal 9:118. doi: 10.4236/oalib.1109007 Google Scholar
Fontes, JC. 1992. Chemical and isotopic constraints on 14C dating of groundwater. In: Taylor RE, Long A, Kra RS, editors. Radiocarbon after four decades. Springer. p. 242–261.CrossRefGoogle Scholar
Fontes, JC, Garnier, JM. 1979. Determination of the initial 14C activity of the total disolved carbon: a review of the existing models and a new approach. Water Resour. Res. 15:399413. 10.1029/WR015i002p00399 CrossRefGoogle Scholar
Ge, T, Wang, X, Zhang, J, Luo, C, Xue, Y. 2016. Dissolved Inorganic Radiocarbon in the Northwest Pacific Continental Margin. Radiocarbon 58(3):517529. doi: 10.1017/RDC.2016.23 CrossRefGoogle Scholar
Geyh, MA. 2000. An overview of 14C analysis in the study of groundwater. Radiocarbon 42:99114.CrossRefGoogle Scholar
Haas, H, Fisher, DW, Thorstenson, DC, Weeks, EP. 2016. 13CO2 and 14CO2 Measurements on soil atmosphere sampled in the sub-surface unsaturated zone in the Western Great Plains of the US. Radiocarbon 25:301314.CrossRefGoogle Scholar
Hamed, Y, Dassi, L, Ahmadi, R, Dhia, HB. 2008. Geochemical and isotopic study of the multilayer aquifer system in the Moulares-Redayef basin, southern Tunisia. Hydrol. Sci. J. 53:12411252.CrossRefGoogle Scholar
Han, LF, Plummer, LN. 2016. A review of single-sample-based models and other approaches for radiocarbon dating of dissolved inorganic carbon in groundwater. Earth Sci. Rev. 152:119142. 10.1016/j.earscirev.2015.11.004 CrossRefGoogle Scholar
Han, LF, Plummer, LN. 2016. A review of single-sample-based models and other approaches for radiocarbon dating of dissolved inorganic carbon in groundwater. Earth Sci. Rev. 152:119142.CrossRefGoogle Scholar
Han, LF, Plummer, LN, Aggarwal, P. 2012. A graphical method to evaluate predominant geochemical processes occurring in groundwater systems for radiocarbon dating. Chem. Geol. 318:88112.CrossRefGoogle Scholar
Herrera, C, Godfrey, L, Urrutia, J, Custodio, E, Jordan, T, Jodar, J, Delgado, K, Barrenechea. 2021. Recharge and residence times of groundwater in hyper arid areas: the confined aquifer of Calama, Loa River Basin, Atachama Desert, Chile. Sci. Total Env. 752:141847.CrossRefGoogle ScholarPubMed
Hogg, A, Heaton, T, Hua, Q, Palmer, J, Southon, J, et al. 2020. SHCal20 Southern Hemisphere calibration, 0–55,000 years cal BP. Radiocarbon 62(4):759778.CrossRefGoogle Scholar
Iverach, CP, Cendón, DI, Meredith, KT, Wilcken, KM, Hankin, SI, Andersen, MS, Kelly, BF. 2017. A multi-tracer approach to constraining artesian groundwater discharge into an alluvial aquifer. Hydrol. Earth Syst. Sci. 21:59535969.CrossRefGoogle Scholar
Jull, AJT, Donahue, DJ, Hatheway, AL, Linick, TW, Toolin, LJ. 1986. Production of graphite targets by deposition from CO/H2 for precision accelerator 14C measurements. Radiocarbon 28:191197.CrossRefGoogle Scholar
Jurgens, BC, Böhlke, JK, Eberts, SM. 2012. Tracer LPM (Version 1): an Excel® workbook for interpreting groundwater age distributions from environmental tracer data: U.S. Geological Survey Techniques and Methods Report 4-F3:60.Google Scholar
Kalin, RM. 2000. Radiocarbon dating of groundwater systems. In: Cook, P, Herczeg, A, editors. Environmental tracers in subsurface hydrology. New York: Kluwer. p. 111144.CrossRefGoogle Scholar
Kohn, MJ. 2010. Carbon isotope compositions of terrestrial C3 plants as indicators of (paleo)ecology and (paleo) climate. Proc. Natl. Acad. Sci. USA 107:1969119695.CrossRefGoogle ScholarPubMed
Laskar, AH, Sharma, N, Ramesh, R, Jani, RA, Yadava, MG. 2010. Paleoclimate and paleovegetation of Lower Narmada Basin, Gujarat, Western India, inferred from stable carbon and oxygen isotopes. Quaternary Interanational 227:183189.Google Scholar
Laskar, A, Yadava, M, Sharma, N, Ramesh, R. 2013. Late-Holocene climate in the Lower Narmada valley, Gujarat, western India, inferred using sedimentary carbon and oxygen isotope ratios. The Holocene 23(8):11151122. https://doi.org/10.1177/0959683613483621 CrossRefGoogle Scholar
Laskar, AH, Yadava, MG, Ramesh, R. 2016. Stable and radioactive carbon in forest soil of Chhattisgarh, Central India: implications for tropical soil carbon dynamics and stable carbon isotope evolution. J. Asian Earth Sci. 123:4757. doi: 10.1016/j.jseaes.2016.03.021 CrossRefGoogle Scholar
Laskar, AH, Yadava, MG, Ramesh, R. 2012. Radiocarbon and stable carbon isotopes in two soil profiles from Northeast India. Radiocarbon 54(1):8189.CrossRefGoogle Scholar
Macchia, M, D’Elia, M, Quarta, G, Gaballo, V, Braione, E, Maruccio, L, et al. 2013. Extraction of Dissolved Inorganic Carbon (DIC) from Seawater Samples at CEDAD: results of an intercomparison exercise on samples from Adriatic Sea shallow water. Radiocarbon 55(2):579584.CrossRefGoogle Scholar
Magnone, D, Richards, LA, van Dongen, BE, Bryant, C, Evans, JA, Polya, DA. 2019. Calculating 14C mean residence times of inorganic carbon derived from oxidation of organic carbon in groundwater using the principles of 87Sr/86Sr and cation ratio mixing. Geochim Cosmochim Acta 267: 322340.CrossRefGoogle Scholar
Maloszewski, P. 2000. Lumped-parameter models as a tool for determining the Hydrological parameters of some groundwater systems based on isotope data. IAHSAISH Publication. p. 271276.Google Scholar
Maloszewski, P, Stichler, W, Zuber, A. 2004. Interpretation of environmental tracers in groundwater systems with stagnant water zones. Iso. Env. Health Stud. 40(1):2133.CrossRefGoogle ScholarPubMed
Maloszewski, P, Zuber, A. 1982. Determining the turnover time of groundwater systems with the aid of environmental tracers: 1. Models and their applicability. J. Hydrol. 57:207231. https://doi.org/10.1016/0022-1694(82)90147-0 CrossRefGoogle Scholar
Maloszewski, P, Zuber, A. 1992. On the calibration and validation of mathematical models for the interpretation of tracer experiments in groundwater. Adv. Water Resour. 15:4762. https://doi.org/10.1016/0309-1708(92)90031-V CrossRefGoogle Scholar
McCarroll, D, Loader, NJ. 2004. Stable isotopes in tree rings. Quart. Sci. Rev. 23:771801.CrossRefGoogle Scholar
McGuire, KJ, McDonnell, JJ. 2006. A review and evaluation of catchment transit time modeling. J. Hydrol. 330:543563. https://doi.org/10.1016/j.jhydrol.2006.04.020 CrossRefGoogle Scholar
Mook, WG. 1972. On the reconstruction of the initial 14C content of groundwater from the chemical and isotopic composition, Proceedings of Eighth International Conference on Radiocarbon Dating 1:342–352. Royal Society of New Zealand, Wellington.Google Scholar
Mook, WG, Bommerson, JC, Staverman, WH. 1974. Carbon isotope fractionation between dissolved bicarbonate and gaseous carbon dioxide. Earth Planet. Sci. Let. 22:169176.CrossRefGoogle Scholar
Nydahl, AC, Wallin, MB, Laudon, H, Weyhenmeyer, GA. 2020. Groundwater carbon within a boreal catchment: Spatiotemporal variability of a hidden aquatic carbon pool. Journal of Geophysical Research: Biogeosciences 125:e2019JG005244. https://doi.org/10.1029/2019JG005244 CrossRefGoogle Scholar
Paul, A, Balesdent, J, Hatté, C. 2020. 13C-14C relations reveal that soil 13C-depth gradient is linked to historical changes in vegetation 13C. Plant Soil 447:305317. https://doi.org/10.1007/s11104-019-04384-4 CrossRefGoogle Scholar
Pearson, FJ, Hanshaw, BB. 1970. Sources of Dissolved Carbonate Species in Groundwater and their Effects on Carbon-14 dating, Isotope Hydrology 1970. IAEA, Vienna. p. 271286.Google Scholar
Plummer, LN, Glynn, PD. 2013. Radiocarbon dating in groundwater systems. Vienna: IAEA. p. 3389.Google Scholar
Priestley, SC, Wohling, DL, Keppel, MN, Post, VEA, Love, AJ, Shand, P, Tyroller, L, Kipfer, R. 2017. Detecting inter-aquifer leakage in areas with limited data using hydraulics and multiple environmental tracers, including 4He, 36Cl/Cl, 14C and 87Sr/86Sr. Hydrogeol. J. 25:20312047. https://doi.org/10.1007/s10040-017-1609-x CrossRefGoogle Scholar
Rapp, BE. 2017. Microfluidics, modelling, mechanics and mathematics. Oxford: Elsevier. p. 243263.CrossRefGoogle Scholar
Reimer, P, Austin, W, Bard, E, Bayliss, A, Blackwell, P, Bronk Ramsey, C. et al. 2020. The IntCal20 Northern Hemisphere radiocarbon age calibration curve (0–55 cal kBP). Radiocarbon 62(4):725757. doi: 10.1017/RDC.2020.41 CrossRefGoogle Scholar
Ripperdan, RL. 2001. Stratigraphic variation in marine carbonate carbon isotope ratios. Reviews in Mineralogy and Geochemistry 43(1):637662.CrossRefGoogle Scholar
Rozanski, K. 1991. Report on the Consultants’ group meeting on 14C reference materials for radiocarbon laboratories, February 18–20, 1991, Vienna, Austria.Google Scholar
Seltzer, A M, et al. 2021. Groundwater residence time estimates obscured by anthropogenic cabonate. Sci. Adv. 7:eabf3503(2021). doi: 10.1126/sciadv.abf3503 CrossRefGoogle Scholar
Solder, JE, Jurgens, BC. 2020. Evaluation of soil zone processes and a novel radiocarbon correction approach for groundwater with mixed sources. Journal of Hydrology 588:124766. https://doi.org/10.1016/j.jhydrol.2020.124766 CrossRefGoogle Scholar
Starn, JJ, Kauffman, LJ, Carlson, CS, Reddy, JE, Fienen, MN. 2021. Three-dimensional distribution of groundwater residence time metrics in the glaciated United States using metamodels trained on general numerical simulation models. Water Resources Research 57:e2020WR027335. https://doi.org/10.1029/2020WR027335 CrossRefGoogle Scholar
Stuiver, M, Polach, HA. 1977. Discussion: reporting of 14C data. Radiocarbon 19(3):355363.CrossRefGoogle Scholar
Suckow, A. 2014. The age of groundwater—definitions, models and why we do not need this term. Appl. Geochem. 50: 222230. https://doi.org/10.1016/j.apgeochem.2014.04.016 CrossRefGoogle Scholar
Szaran, J. 1997. Achievement of carbon isotope equilibrium in the system HCO3¯ (solution)—CO2 (gas). Chemical Geology 142: 7986.CrossRefGoogle Scholar
Takahashi, HA, Handa, H, Minami, M. 2021. A simple CO2 extraction method for radiocarbon analyses of dissolved inorganic carbon in water samples without a carrier gas. Radiocarbon 63:13391353.CrossRefGoogle Scholar
Tamers, MA. 1975. The validity of radiocarbon dates on groundwater. Geophys. Survey 2:217239. https://doi.org/10.1007/BF01447909 CrossRefGoogle Scholar
Telloli, C, Rizzo, A, Salvi, S, Pozzobon, A, Marrocchino, E, Vaccaro, C. 2022. Characterization of groundwater recharge through tritium measurements. Adv. Geosci. 57:2136.CrossRefGoogle Scholar
Thorstenson, DC, Weeks, EP, Haas, H, Fisher, DW. 2016. Distribution of gaseous 12CO2, 13CO2, and 14CO2 in the sub-soil unsaturated zone of the Western US great plains. Radiocarbon 25:315346. https://doi.org/10.1017/S0033822200005622 CrossRefGoogle Scholar
Vogel, JC, Ehhalt, D. 1963. The use of carbon isotopes in groundwater studies radioisotopes in Hydrology. IAEA, Vienna. p. 383395.Google Scholar
Vogel, JC, Grootes, PM, Mook, WG. 1970. Isotopic fractionation between gaseous and dissolved carbon dioxide. Z. Physics 230:255258.Google Scholar
Wilske, C, Suckow, A, Mallast, U, Meier, C, Merchel, S, Merkel, B, Pavetich, S, Rödiger, T, Rugel, G, Sachse, A, Weise, SM, Siebert, C. 2020. A multi-environmental tracer study to determine groundwater residence times and recharge in a structurally complex multi-aquifer system, Hydrol. Earth Syst. Sci. 24:249267.CrossRefGoogle Scholar
Wood, C, Cook, PG, Harrington, GA. 2015. Vertical carbon-14 profiles for resolving spatial variability in recharge in arid environments. J. Hydrol. 520:134142. https://doi.org/10.1016/j.jhydrol.2014.11.044 CrossRefGoogle Scholar
Zuber A, Maloszewski. 2000. Lumped parameter models. In: Yurtsever Y, editor. Environemntal Isotopes in the Hydrological Cycle 6. p. 497–512.Google Scholar
Figure 0

Figure 1 Location for the groundwater sampling site, the campus of the Physical Research Laboratory (PRL). Locations of the sampling sites whose groundwater samples were dated previously by Agarwal et al. (2006) are also shown (the site number were taken from their published article). Samples with radiocarbon ages neither modern nor beyond radiocarbon dating limit are listed here.

Figure 1

Figure 2 Soil air sampling setup. The soil air samples are collected from a specific soil depth introducing the pipe to that depth and filling the pre-evacuated high vacuum glass flask.

Figure 2

Figure 3 Glass vacuum line used for extraction of carbon dioxide from dissolved inorganic carbon and soil CO2. Water samples are injected into the sample flask, reacted with orthophosphoric acid under vacuum, passed through tubes immersed in slush (−80°C) and liquid nitrogen (−196°C) and finally collected in the sample bottle.

Figure 3

Figure 4 Calibration plot used for estimating the amount of CO2 liberated from any sample.

Figure 4

Figure 5 Graphite preparation system. Fe: iron powder, Zn: zinc powder, PG: pressure gauge, VG: vacuum gauge, P: pump, SB: sample bottle to load CO2 into the system.

Figure 5

Figure 6 Schematics of (A) piston flow model (after Bethke and Johnson, 2008) and (B) dispersion model (after Maloszewski and Zuber 1982).

Figure 6

Table 1 Yield of DIC extraction, measured using the present technique. Water samples with different DIC concentration are prepared by dissolving known amount of sodium carbonate. The solutions are then reacted with orthophosphoric acid and the amount of liberated CO2 were measured. The measured amounts of CO2 for different samples are compared to that actually present and the percentage yields are calculated

Figure 7

Table 2 Radiocarbon dates of two international standards measured using the present technique (details of the standards are available in Rozanski et al. 1991)

Figure 8

Table 3 Radiocarbon ages of PRL groundwater samples (with Libby half-life). A δ13C value of soil CO2 of –19.4‰ (average in the soil profile, see Table 4) is considered for the calculation. The median of the calibrated age is used for estimating mean residence time

Figure 9

Figure 7 Variation in δ13C in soil CO2 with depth in a soil profile inside the Campus of Physical Research Laboratory Ahmedabad.

Figure 10

Table 4 Stable carbon isotopic composition (δ13C) and radiocarbon content (F14C) in soil CO2 at different depths in a soil profile from the Campus of Physical Research Laboratory. The reproducibility in δ13C measurement is ∼0.1‰. 2 to 3 samples from the same depths (different samples) for some samples were collected, processed and analyzed for establishing the reproducibility of the method. Radiocarbon ages in BP are also presented

Figure 11

Figure 8 Estimated residence times with change in radiocarbon ages using piston flow model (PFM), dispersion model (DM) with a dispersion parameter of 0.1 and binary mixing model (BMM). Residence times corresponding to dispersion parameters of 0.05 and 0.1 for DM are shown. In the BMM, calculation is made with 10% contribution from PFM and 90% from DM.

Figure 12

Figure 9 Residence times of the groundwater samples of Gujarat region using available radiocarbon dates from Agarwal et al (2006) (see Figure 1 for locations). Residence times estimated using piston flow model (PFM) and Dispersion Model (DM) with a dispersion parameter of 0.1 are shown along with the measured radiocarbon ages for different samples.