Sunteți pe pagina 1din 13

Chemical Geology 413 (2015) 94–106

Contents lists available at ScienceDirect

Chemical Geology

journal homepage: www.elsevier.com/locate/chemgeo

3
H/3He, 14C and (U–Th)/He groundwater ages in the St. Lawrence
Lowlands, Quebec, Eastern Canada
Geneviève Vautour a, Daniele L. Pinti a,⁎, Pauline Méjean a, Marion Saby a, Guillaume Meyzonnat a,
Marie Larocque a, M. Clara Castro b, Chris M. Hall b, Christine Boucher a,1, Emilie Roulleau c,2, Florent Barbecot a,
Naoto Takahata c, Yuji Sano c
a
GEOTOP and Département des sciences de la Terre et de l'atmosphère, Université du Quebec à Montreal, CP8888 succ. Centre-Ville, Montreal, QC, Canada
b
Dept. of Earth and Environmental Sciences, University of Michigan, 1100 N. University, Ann Arbor, MI 48109-1005, USA
c
Atmosphere and Ocean Research Institute, The University of Tokyo, Kashiwa, Chiba 277-8564, Japan

a r t i c l e i n f o a b s t r a c t

Article history: This study attempts to place constraints on groundwater residence times using helium isotopes and 14C of a
Received 8 February 2015 regional groundwater flow system – the Becancour River watershed – located midway between Montreal and
Received in revised form 27 July 2015 Quebec City (Quebec, Canada). This densely populated region is one of the main targets for shale gas exploitation
Accepted 5 August 2015
in Eastern Canada. For this reason, this watershed has been the focus of a detailed aquifer study to gain a better
Available online 1 September 2015
understanding of groundwater resources, both in terms of availability and quality. In the current study, noble
Keywords:
gases were sampled and analyzed in twenty-eight wells from a Quaternary granular aquifer and a regional
Helium isotopes bedrock aquifer of Ordovician age. Tritium (3H) and radiocarbon (A14C) activities were measured on selected
3
H/3He ages wells. Helium isotopic ratios 3He/4He (R) normalized to that of the atmosphere (Ra = 1.386 × 10−6) range
U–Th/4He ages from R/Ra = 0.039 ± 0.003 to 3.109 ± 0.065. The helium isotopic signatures point to the presence of three
14
C ages water bodies: 1) modern infiltration water with nearly atmospheric helium isotopic ratio and little post-bomb
St. Lawrence Lowlands tritium recharging the shallower granular aquifer; 2) mid-50s tritium-rich water slightly mineralized; and 3)
Quebec an older water component rich in terrigenic helium flowing in from the bedrock fractured aquifer. Uncorrected
14
C ages range from 15 ka to modern. 14C is affected by several dead carbon reservoirs related to carbonate dis-
solution, cations exchange, oxidation of organic matter and methanogenesis. Adjusted 14C ages calculated with
NETPATH range from 7 ka to modern. Older 14C ages correspond to the end of the regional re-organization of
the hydrological system following deglaciation and isostatic rebound. Noble gas recharge paleotemperatures
are 4–9 °C colder than the present temperatures, although no clear relation with ages has been found. The
relationship between the helium isotopic ratios and 14C ages suggests that the regional bedrock aquifer is affected
by mixing between these three water sources. Calculated (U–Th)/4He ages can be partially explained by in situ
production of 4He in the aquifer rock and the addition of a radiogenic helium source external to the aquifer.
Calculated minimum helium fluxes of 1.0 × 10−8 to 1.8 × 10−7 cm3 STP cm−2 yr−1 are tens to hundreds of
times lower than the average continental crust flux of 3.3 × 10−6 cm3 STP cm−2 yr−1, suggesting local sources,
possibly from production of radiogenic helium in the shale gas formations underlying the studied aquifers. The
occurrence of old groundwater in this aquifer system clearly limits the renewable resource and increases the
risk of overexploitation in the case of increased use or in the case of pollution from different sources.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction including recharge conditions and geological characteristics of the


reservoirs (e.g., Carrera and Neuman, 1986; Harvey and Gorelick,
Groundwater flow rates in most aquifers are difficult to estimate be- 1995; Castro and Goblet, 2005). Cross-formational flow and water
cause they are dependent on multiple and highly variable parameters, mixing (e.g., Pinti and Marty, 1998; Ma et al., 2005, 2009) make the
estimation of groundwater flow rates even more complex. Using the
⁎ Corresponding author. distribution of natural tracer concentrations in an aquifer may contrib-
E-mail address: pinti.daniele@uqam.ca (D.L. Pinti). ute to quantify groundwater age. They can be further used to constrain
1
Now at: Centre de Recherches Pétrographiques et Géochimiques (CRPG), Nancy
Université, BP 20, 54501 Vandoeuvre-lès-Nancy Cedex, France.
groundwater flow, to identify recharge areas and estimate recharge
2
Now at: CEGA - Centro de Excelencia en Geotermia de Los Andes, Departamento de rates, as well as to understand groundwater susceptibility to contami-
Geologia, Plaza Ercilla 803, Santiago, Chile. nation (e.g., Solomon et al., 1993).

http://dx.doi.org/10.1016/j.chemgeo.2015.08.003
0009-2541/© 2015 Elsevier B.V. All rights reserved.
G. Vautour et al. / Chemical Geology 413 (2015) 94–106 95

Estimation of groundwater ages can be challenging (e.g., Goode, wide range of groundwater residence times, from modern water with
1996; Bethke and Johnson, 2008) and often calls upon the use of multi- the use of the 3H/3He method (e.g., Takaoka and Mizutani, 1987),
ple isotopic tracers such as stable isotopes, CFC, SF6 or noble gases to million years old water with the use of the U–Th/4He method
(e.g., Cook and Solomon, 1997). These tracers provide information on (e.g., Andrews, 1985; Torgersen and Clarke, 1985; Castro et al., 2000;
water mixing displaying different residence times and distinct flow Kulongoski et al., 2005; Aggarwal et al., 2015).
paths (e.g., Castro and Goblet, 2003). Because they are inert and stable, This study uses noble gases to constrain groundwater flow and to es-
noble gases are suited to place constraints on groundwater flow rates timate groundwater residence times in the regional groundwater flow
and flow direction (e.g., Phillips and Castro, 2003). In addition, a number system of the Becancour watershed (Fig. 1). The studied area is located
of noble gas isotopes are produced by radioactive decay of parent midway between Montreal and Quebec City, in the Province of Quebec,
elements with contrasting half-lives, facilitating the estimation of a eastern Canada. This region, corresponding geographically to the St.

Fig. 1. Simplified map of the Becancour watershed with potenziometric head isolines for the regional fractured bedrock aquifer, groundwater sampling locations and major tectonic
accidents (A). Profile along a NW–SE direction showing the architecture of the sampled aquifers (B).
96 G. Vautour et al. / Chemical Geology 413 (2015) 94–106

Table 1
He, Ne and Ar isotopic compositions of groundwater in the Becancour basin together with chemical and physical data.

Sample name X coord. Y coord. Altitude Water Depth Casing T TDS δ18O δD 4
Hemeas ±1σ 20
Ne ±1σ 36
Ar

UTM NAD 83 asl chemistry m length, m °C g L−1 cm3 STP g−1


H2O

x 10−8 x 10−7 x 10−6

Quaternary granular aquifer


BEC 105 (M) −72.0012 46.3643 91.7 Ca-HCO3 7.0 – 9.4 0.213 −11.71 −81.95 9.23 0.14 2.195 0.029 1.52
BEC 106 (M) −72.2271 46.3680 58.6 Ca-HCO3 20.0 – 8.3 0.062 −12.42 −87.59 7.12 0.11 2.130 0.030
BEC 113 (M) −71.5788 46.3699 138.6 Ca-SO4 9.1 – 7.8 0.270 −12.57 −88.81 11.08 0.17 3.502 0.046
BEC 117 (M) −72.1083 46.5447 40.2 Ca-HCO3 15.2 – 7.4 0.084 −12.12 −84.01 5.36 0.08 1.898 0.025
BEC 118 (P) −72.1274 46.5265 41.7 Ca-HCO3 6.1 – 9.9 0.298 −12.43 −86.06 6.04 0.09 2.275 0.030 1.49
BEC 124 (P) −71.6002 46.2505 317.3 Ca-Cl 2.2 – 10.2 0.053 −13.22 −93.75 4.81 0.07 1.546 0.020

Ordovician fractured aquifer


BEC 009 (P) −72.0076 46.2512 94.8 Ca-HCO3 50.3 18.6 10.0 0.197 −11.75 −83.55 8.02 0.12 2.209 0.029
BEC 010_1 (P) −71.9328 46.2696 105.2 Ca-HCO3 37.0 4.9 10.6 0.192 −10.88 −76.28 52.54 0.79 2.137 0.028
BEC 010_2 (P) 71.25 1.07 2.222 0.029
BEC 021 (P) −71.9511 46.3237 113.1 Ca-HCO3 25.9 12.2 10.6 0.294 −11.48 −79.48 5.98 0.09 1.907 0.025
BEC 034 (P) −71.8509 46.2415 130.1 Ca,Na-HCO3 24.4 6.0 9.3 0.192 −11.58 −80.73 5.68 0.09 1.443 0.019
BEC 101 (P) −72.1687 46.2760 90.6 Na-HCO3,Cl 47.2 12.6 12.0 0.780 −11.93 −83.60 1169.41 17.54 3.163 0.041 2.01
BEC 102 (M) −72.0291 46.4898 60.1 Ca,Na-HCO3 21.6 20.0 8.9 0.427 −11.29 −77.26 7.95 0.12 1.725 0.022
BEC 103 (M) −72.3201 46.1482 81.7 Ca,Na-HCO3 43.6 7.6 9.6 0.501 −12.57 −88.81 63.35 0.95 4.803 0.062 2.15
BEC 107 (P) −72.2478 46.2156 82.0 Ca-HCO3 36.6 17.5 13.3 0.188 −11.52 −80.31 39.65 0.59 2.260 0.029 1.48
BEC 110 (P) −72.5012 46.1853 49.7 Ca,Na-HCO3 37.8 26.1 8.9 0.346 −11.19 −77.80 60.65 0.91 2.217 0.029 1.48
BEC 112 (M) −71.8732 46.4124 115.0 Ca,Na-HCO3 38.1 4.7 8.6 0.286 −11.90 −84.70 6.70 0.10 2.191 0.028 1.58
BEC 119 (P) −72.0529 46.5095 54.8 Na-HCO3 45.7 34.1 12.0 0.471 −11.51 −81.49 62.92 0.94 2.360 0.031 1.58
BEC 126_1 (P) −71.5454 46.3040 210.2 Ca,Na-HCO3 49.1 3.8 8.8 0.227 −11.77 −82.03 4482.00 67.23 2.819 0.037 2.15
BEC 126_2 (P) 3.8 2662.41 39.94 2.508 0.033 2.08
BEC 137 (P) −71.6311 46.3518 141.6 Na-HCO3 23.7 10.0 8.6 0.350 −11.68 −82.13 11.32 0.17 2.251 0.029 1.65
BEC 138 (P) −71.7041 46.3403 130.6 Ca-HCO3 32.0 3.9 8.9 0.216 −11.85 −82.25 7.72 0.12 2.612 0.034 1.72
BEC 147 (P) −71.8162 46.3475 122.9 Ca-SO4 32.0 6.1 9.6 0.157 −11.87 −83.56 7.51 0.11 2.382 0.031 1.63
BEC 148 (P) −71.7726 46.3266 120.2 Na-SO4 64.6 13.7 8.7 0.109 −11.61 −80.37 12.38 0.19 2.579 0.034 1.63
BEC 149 (P) −71.8032 46.2562 128.1 Ca-HCO3 54.9 8.3 9.5 0.286 −11.87 −81.59 97.94 1.47 2.223 0.029 1.59
F2 PACES −72.2125 46.2252 130.0 Na-SO4 42.0 – 7.2 0.131 −11.53 −79.92 21.22 0.32 2.149 0.028 1.72
F3 PACES −72.2022 46.3511 84.9 Na-Cl 78.3 47.9 7.2 2.175 −14.15 −99.26 45.14 0.68 1.722 0.022 0.52
F4 PACES −71.8462 46.3822 128.5 Na-HCO3 36.6 9 8.5 0.339 −11.67 −79.46 62.30 0.93 1.901 0.025 4.59
F5 PACES −71.9141 46.4507 103.0 Na-HCO3 47.2 17.44 8.0 0.529 −12.25 −83.74 12.77 0.19 1.937 0.025 1.35
F9 PACES −72.4686 46.3260 10.0 Na-HCO3,Cl 35.7 5.1 8.5 0.222 −10.92 −74.44 272.50 4.09 0.996 0.013 1.15
Air/ASW 4.72 1.860 1.31

F = fractured; G = granular; P = private well; M = municipal well; PACES = instrumented wells, drilled in the framework of the PACES project (see text for details).
Air Saturated Water (ASW) values are indicative and calculated for a freshwater with TDS = 0 mg L−1 and T = 10 °C using Smith and Kennedy (1983) equations.
Casing is the length of the casing in the unconsolidated granular aquifer section of wells tapping water in the fractured bedrock.
Granular aquifer's wells have casing all along to the bottomhole which are equipped with a short screen (1 m).
When casing length is not reported for wells in the fractured aquifer it means that the fractured bedrock arrives at the surface with little or no cover (ex. BEC126).

Lawrence Lowlands, is the third most densely populated region in by an irregular topography with maximum elevations of 500 masl
Canada (158 inhabitants per km2), with agriculture and industry (Fig. 1B). The two regions correspond to the St. Lawrence Platform
being the province's main economic drivers. The St. Lawrence Lowlands and the Appalachian Mountains geological provinces, respectively. The
is also one of the main regions targeted for shale gas exploitation in St. Lawrence Platform is a Cambrian–Lower Ordovician siliciclastic and
Canada (Lavoie et al., 2013; Pinti et al., 2013; Moritz et al., 2015). For carbonate platform having a thickness of ca. 1200 m, overlain by ca.
these reasons, this watershed has been the focus of a detailed aquifer 1800 m of Middle–Late Ordovician foreland carbonate-clastic deposits
characterization project funded by the Quebec Ministry of Environment (Lavoie et al., 2013). The Ordovician geological units of the St. Lawrence
and by local municipalities (Larocque et al., 2013). Results from this pro- Platform outcropping in the study area are the Upper Ordovician
ject include analyses for solute ion contents, trace elements and stable turbiditic Nicolet Fm. (Fig. 1) of the Lorraine Group and the Saint-
isotopes of water (Larocque et al., 2013; Meyzonnat et al., in press). Sabine les Fonds Fm. of the Sainte-Rosalie Group. These two formations
The focus of the present study is to determine groundwater residence are dominated by thick successions of mudstones with subordinate
times using noble gases, tritium (3H) and radiocarbon (14C). The overall alternating sandstone and siltstone. Another unit outcropping in the
goal was to gain a better understanding of groundwater resources both St. Lawrence Platform is the Becancour Fm. of the Queenston Group
in terms of availability and quality, with an eye to future natural consisting of post-orogenic red shale, sandstone and conglomerate
resource exploitation. (Globensky, 1987).
The terrains outcropping in the Appalachian Mountains correspond
2. Geological and hydrogeological background to imbricated thrust sheets. In the Becancour watershed area we can
distinguish Cambrian green and red shales (Sillery Group); Ordovician
2.1. Geology bedded black and yellowish-weathered shaly matrix containing chaotic
blocks of shales, cherts, sandstone forming the “wildflysch” of the
The studied area is located in southwestern Quebec, Canada Etchemin River (Fig. 1); middle Ordovician dolomitic or calcitic schists
(Fig. 1A). It covers 2859 km2 in the southern portion of the Becancour of the Sweetsburg and West Sutton Fm. of the Oak Hill Group (Fig. 1)
River watershed and includes eight sub-watersheds that drain into the (Globensky, 1993).
St. Lawrence River (Fig. 1A). The northwestern part corresponds to the Unconsolidated Quaternary sediments derived from multiple
St. Lawrence Lowlands, a flat area with elevations less than 150 masl, glaciation–deglaciation cycles cover unconformably the Cambrian–
while the southeastern part is located in the Appalachians, marked Ordovician sequence of the St. Lawrence Platform (Lamothe, 1989).
G. Vautour et al. / Chemical Geology 413 (2015) 94–106 97

H
84 132 20
±1σ Kr ±1σ Xe ±1σ R/Ra ±1σ Ne/22Ne ±1σ 21
Ne/22Ne ±1σ 38
Ar/36Ar ±1σ 40
Ar/36Ar ±1σ

x 10−8 x 10−9

0.02 6.02 0.09 4.13 0.09 0.788 0.021 9.81 0.03 0.0282 0.0002 0.1875 0.0003 297.4 0.7
1.115 0.007
1.500 0.008
1.604 0.010
0.02 5.55 0.08 3.69 0.08 1.077 0.037 9.83 0.03 0.0283 0.0001 0.1873 0.0005 296.5 0.8
1.090 0.025

3.109 0.065
0.414 0.013
0.308 0.005
2.299 0.049
1.912 0.042
0.03 6.95 0.10 3.76 0.08 0.074 0.003 9.87 0.05 0.0288 0.0002 0.1907 0.0016 296.5 3.4
2.316 0.048
0.03 7.51 0.11 4.94 0.11 0.381 0.011 9.81 0.02 0.0285 0.0002 0.1875 0.0006 296.7 0.8
0.02 5.57 0.08 3.87 0.09 0.412 0.009 9.76 0.05 0.0288 0.0001 0.1879 0.0007 295.0 1.1
0.02 5.66 0.08 3.78 0.08 0.172 0.008 9.75 0.05 0.0287 0.0001 0.1876 0.0005 295.3 1.0
0.02 6.15 0.09 4.98 0.11 1.238 0.041 9.85 0.04 0.0282 0.0001 0.1873 0.0005 295.1 0.8
0.02 5.94 0.09 4.07 0.09 0.283 0.014 9.75 0.03 0.0291 0.0002 0.1868 0.0017 297.3 2.1
0.03 7.37 0.11 4.51 0.10 0.039 0.003 9.85 0.05 0.0290 0.0002 0.1874 0.0008 297.4 1.0
0.03 6.94 0.10 4.56 0.10 0.054 0.005 9.84 0.04 0.0293 0.0004 0.1829 0.0012 294.7 1.5
0.02 6.41 0.10 4.66 0.10 0.720 0.019 9.83 0.04 0.0285 0.0001 0.1880 0.0008 295.2 0.8
0.02 6.53 0.10 4.56 0.10 2.005 0.039 9.85 0.03 0.0286 0.0001 0.1878 0.0009 295.0 1.8
0.02 6.31 0.09 4.42 0.10 1.341 0.033 9.76 0.03 0.0285 0.0001 0.1881 0.0007 297.6 0.9
0.02 6.56 0.10 4.65 0.10 0.711 0.014 9.83 0.03 0.0286 0.0001 0.1878 0.0007 293.9 0.9
0.02 6.29 0.09 4.60 0.10 0.344 0.010 9.80 0.04 0.0287 0.0001 0.1884 0.0006 295.5 0.9
0.02 6.56 0.10 4.57 0.10 0.386 0.010 9.85 0.02 0.0277 0.0002 0.1884 0.0010 295.8 1.3
0.01 2.53 0.04 2.01 0.04 0.073 0.003 9.92 0.09 0.0279 0.0002 0.1871 0.0011 296.3 0.8
0.06 11.70 0.18 6.11 0.13 1.043 0.038 9.79 0.01 0.0284 0.0002 0.1864 0.0008 295.3 0.7
0.02 5.14 0.08 3.59 0.08 0.564 0.020 9.91 0.02 0.0273 0.0001 0.1874 0.0008 295.7 1.0
0.01 4.59 0.07 2.43 0.05 0.058 0.002 9.87 0.05 0.0273 0.0002 0.1866 0.0007 297.2 0.8
5.26 3.61 1.000 9.80 0.0290 0.1880 295.5

Quaternary deposits are also found in the Appalachian Mountains, in Only one medium- to high-yield aquifer exists in Quaternary
the most heavily incised river valleys. A nearly continuous till sheet glaciofluvial coarse-grained surficial sediments, the “Vielles Forges”
(Gentilly till) covers most of the area between the overlain discontinu- aquifer (Lamothe, 1989), but spatial extent remains undefined
ous marine and lacustrine silt (Lampsilis Lake silts) and clay units of the (Lamothe, 1989; Godbout, 2013). The granular Quaternary aquifer is
Champlain Sea (11.2 to 9.8 kyrs old) and the underlain discontinuous separated from the fractured bedrock aquifer by a basal till aquitard
patches of sand deposited during marine regressions (Vielles Forges unit (Becancour till; Lamothe, 1989). Regional groundwater flow is
and Lotbinière sands; Lamothe, 1989) with an optical stimulated lumi- SE–NW and usually follows the topography, with recharge occurring
nescence (IRLS) measured age of 44–50 ka (Godbout, 2013). mainly in the Appalachian Mountains and discharge in the main
tributaries of the St. Lawrence River (Fig. 1B). Local recharge areas are
2.2. Hydrogeology and hydrogeochemistry also found in the plain, along a band located just northwest of the
Appalachian piedmont, where the Champlain Sea clays are absent
The regional aquifer is located in the fractured bedrock consisting (Larocque et al., 2013).
of the Cambrian–Ordovician schists and shales of the Appalachian Groundwater in the Becancour watershed is characterized by low-
Mountains and St. Lawrence Platform. This aquifer is mainly unconfined salinity with Total Dissolved Solids (TDS) ranging from 0.5 to 2.2 g L−1
in the Appalachian Mountains and in the Appalachian piedmont, (Table 1). Water types are dominated by Ca-HCO3 and Na-HCO3 with
progressively becoming semi-confined, northwesterly. In a 10 km a few samples showing Na-SO4 and Na-Cl chemistry (Meyzonnat
wide zone bordering the St. Lawrence River, the aquifer is confined by et al., in press). Freshwater in the Appalachian recharge acquires
thick marine clay of the Champlain Sea and receives limited recharge Ca–Mg-HCO3 and Ca–Na-SO4 chemistry through interactions with
(Larocque et al., 2013). limestone units of the St. Lawrence Platform. This water evolves toward
The hydraulic conductivities of the studied aquifer are low to a Na-HCO3 type by Ca2+–Na+ ion exchange, where Ca2+ water exchanges
moderate (~10−9–10−6 m s−1) and porosities are variable, between 1 with Na+ mineral (Cloutier et al., 2006, 2010) in semi-confined aquifers. Fi-
and 5% porosity equivalent for the Ordovician fractured regional nally, close to the St. Lawrence River, groundwater evolves into a Na-Cl
aquifer to effective porosities of 10–20% for the Quaternary granular type with salinity likely derived from exchange of freshwater with pore
aquifer (Larocque et al., 2013). Wells in the fractured bedrock aquifer seawater trapped in the Champlain Sea clays or the fractured rock
yield enough water to supply single-family dwellings and, in a few aquifers, in particular in areas confined by thick marine clay with
areas, small- to medium-size municipalities (1000–20,000 inhabitants). limited recharge.
98 G. Vautour et al. / Chemical Geology 413 (2015) 94–106

The stable isotope (δ18O, δD) composition of the groundwater is to a vacuum extraction system and noble gases were quantitatively
shown in Fig. 2. With the exception of F3 and BEC124, water samples extracted for inletting into a MAP-215 mass spectrometer. Extracted
yield a relatively narrow δ18O range value with a mean of − 11.6 ± gases were passed over to a Ti sponge getter to remove reactive gases,
0.6‰ (Fig. 2). This value is similar to the δ18O annual mean of and sequentially allowed to enter the MAP-215 mass spectrometer
− 11.15‰ calculated from forty-one precipitation data measured using a cryo-separator. The cryo-separator temperatures were set at
at two local stations of Lemieux and Saint Ferdinand (Quebec) 35, 65, 200, 215, and 270 K for analysis of He, Ne, Ar, Kr and Xe, respec-
(Meyzonnat, 2012). The sampled groundwater δ2Η and δ18O values tively. The 4He, 20Ne, 22Ne, 36Ar and 40Ar were measured using a Faraday
fall slightly below the LMWL (Local Meteoric Water Line) established detector while 38Ar and all other isotopes were measured using an
on the data collected from these two stations, indicating a slight enrich- electron multiplier in ion counting mode. During neon isotope analysis
ment in 18O due to evaporation. No significant variations are found in a liquid N2 cold trap was applied to minimize peak interferences and
the δ18O values between groundwater from the granular and fractured appropriate mass peaks were monitored to correct for interferences
aquifers. This uniformity in δ18O values might point to modern of 40Ar++ and H18 +
2 O on
20
Ne and CO++
2 on 22Ne. The interference cor-
20 22
precipitation as a common origin for groundwater in both aquifers. rections for Ne and Ne were typically ~1.1% and 0.17%, respectively.
Alternatively, uniform δ18O values might result from well-mixed Prior to each sample analysis, a calibrated amount of air standard was
groundwater sources in these aquifers. F3 and BEC124 deviate from analyzed following a procedure similar to that of the sample measure-
this trend with more depleted δ18O values than the average. F3 is ment and blank correction was applied. Isotopic abundances for each
located in the plain and taps a confined aquifer (Larocque et al., 2013). sample are normalized to the air standard after blank correction.
Groundwater chemistry is of Na-Cl type and yields a TDS value of Elemental abundances of 4He, 22Ne, 36Ar, 84Kr, and 132Xe have typical
2.2 g L−1, which is 10 times the median value for all sampled groundwa- uncertainties of 1.5, 1.3, 1.3, 1.5 and 2.2%, respectively and all uncer-
ter in Becancour (0.284 g L−1; Table 1). Thus, one cannot exclude the tainties are at ±1 σ level. Additional details on the noble gas analytical
possibility that part of the water might be fossil and recharged during procedures can be found in Castro et al. (2009).
a colder period. Well BEC124 is drilled in an unconfined aquifer located At the Atmospheric and Ocean Research Institute (AORI) of the
in the Appalachians and is weakly mineralized (TDS = 0.053 g L−1). It is University of Tokyo, samples were degassed offline using a dedicated
possible that groundwater recharge in this well is dominated by spring line and recovered gas was collected in a lead-rich glass ampoule. The
snowmelt rather than summer rainfall, which would explain the more ampoule was connected to a stainless steel line for purification on a
depleted δ18O value (Fig. 2). Ti-getter and cryogenic separation of He from other noble gases. The
helium aliquot was then analyzed for the 3He/4He isotopic ratio using
3. Sampling and analytical methods a Helix SFT calibrated against the HESJ standard (Helium standard of
Japan; Matsuda et al., 2002). Precision on the 3He/4He ratio of HESJ is
Twenty-eight groundwater samples (two duplicates: BEC010 and typically ± 0.2% (2σ) (Sano et al., 2008). Prior to being cryogenically
BEC126) were collected at depths ranging between 2.2 and 78.3 m adsorbed, the amount of 20Ne was measured on a Balzers QMS Prisma
(Fig. 1; Table 1). Twenty-three samples were collected from municipal 80 connected to the preparation line.
and domestic wells, while five (F2, F3, F4, F5 and F9; Table 1) are from Water samples for tritium analysis were collected using 1 L
instrumented wells drilled by our research group for monitoring Nalgene® bottle filled up and sealed before shipment to the Environ-
purposes. Six wells are tubular wells with screen at their bottom mental Isotope Laboratory at the University (EIL) of Waterloo. Liquid
to tap groundwater from shallower Quaternary granular aquifers. scintillation counting (LSC) is the technique used by EIL for quantifica-
Twenty-two wells tap water from the fractured bedrock aquifer. tion of tritium. Samples analyzed for tritium are enriched 15 times by
These latter are equipped with casing in the section crossing the uncon- electrolysis and subsequently counted. The detection limit for enriched
solidated Quaternary granular aquifers (casing length; Table 1) and samples is 0.8 TU (Table 3). EIL uses a scintillation cocktail having a high
down 1–2 m into the bedrock. In the remaining bedrock section, they
are open boreholes. When information on the casing length was not
available from the owners, the length of the Quaternary deposits was
calculated rather precisely from detailed maps of the Quaternary
deposits (Larocque et al., 2013). With the exception of wells BEC101,
BEC110 and F5, which are drilled in a confined section of the fractured
bedrock aquifer, all the other wells are drilled in unconfined or semi-
confined aquifers.
Groundwater was collected for noble gas measurements at the
instrumented piezometers using a Waterra® Inertial Pump System
which consists of a foot valve fixed at the bottom of a high-density
polyethylene tube with a variable diameter of 5/8 in. to 2 in. and an elec-
tric actuator pump Hydrolift-2®. In municipal wells, water was
collected directly at the wellhead. In domestic wells water was collected
at the closest water faucet, taking precautions to avoid intermediate
reservoirs where the water could undergo degassing. Water was purged
from wells and piezometers until measurement of chemo-physical
parameters (conductivity, pH, Eh, temperature) stabilized. Water was
subsequently allowed to flow through armored PVC tubes connected
to 3/8-inch diameter, refrigeration-type copper tubes. After a few mi-
nutes, two steel clamps were closed and the sealed tube was recovered.
All stable noble gases (He, Ne, Ar, Kr, Xe) were measured in twenty Fig. 2. D vs. δ 18 O plot for the Becancour groundwater, plotted against the Local
groundwater samples (BEC126 as duplicate) at the Noble Gas Laborato- Meteoric Water Line (LMWL), calculated from data recorded at Lemieux and Saint
Ferdinand (Quebec) stations (Meyzonnat, 2012). Stars represent the mean δD vs.
ry at the University of Michigan (Table 1). The remaining eleven
δ 18 O of local rainfall during summer and winter seasons, measured from data of
samples (BEC010 as duplicate) were analyzed only for helium isotopic Lemieux and Saint Ferdinand stations. Black dots are groundwater from the regional
composition and Ne concentrations at the University of Tokyo fractured Ordovician aquifer while white dots represent groundwater from the
(Table 1). At the University of Michigan, water samples were attached granular Quaternary aquifer.
Table 2
3
H–3He, 14C and (U–Th)/He groundwater ages of Becancour.
3 3 3
Well name H activity TU ± He tritiogenic × ± H/3He ± A14C ± δ13C DIC 14
C age A0 = 100 14
C age adjusted 4
Heeq 4
Heea 4
Heterr (U–Th)/4He
10−14 cm3 age yrs activity measured NETPATH yrs F&G NETPATH yrs −8
x 10 3
cm STP g−1 age
STP g−1
H2O measured ‰ VPDB (±0.5%) (±20‰) H2O

pMC % in situ yrs

Quaternary granular aquifer


BEC 105 – – – – – – 104.83 0.27 −13.8 208 −1133 4.70 0.98 3.55 6745
BEC 106 16.09 0.94 3.20 1.47 10.44 0.46 – – – – – 4.75 0.87 1.50 2846
BEC 113 7.62 0.94 9.00 2.39 31.06 1.81 – – – – – 4.75 5.22 1.11 2118
BEC 117 13.59 1.07 5.20 1.43 16.54 0.85 – – – – – 4.76 0.87 −0.27 −514

G. Vautour et al. / Chemical Geology 413 (2015) 94–106


BEC 118 17.17 0.98 0.71 1.22 2.72 0.14 119.64 0.28 −17.0 −1482 0 4.66 1.42 Modern
BEC124 10.72 0.85 0.81 0.97 4.70 0.33 – – – – – 4.71 0.10 182

Ordovician fractured aquifer


BEC 009 8.45 1.30 25.84 3.74 45.98 2.53 – – – – – 4.71 1.23 2.08 677
BEC 010_1 8.73 1.49 8.53 1.86 28.34 2.41 – – – – – 4.70 1.03 46.81 15,208
BEC 010_2 9.40 1.60 7.35 1.79 25.26 2.30 – – – – – 4.70 1.30 65.25 21,197
BEC 021 9.53 0.88 11.94 2.10 31.93 1.36 – – – – – 4.70 0.30 0.98 317
BEC 034 6.81 1.07 8.34 1.68 31.60 2.32 – – – – – 4.73 0.96 311
BEC 101 – – – – – – 21.77 0.14 −7.3 15,426 6696 4.73 4.35 1160.33 376,960
BEC 102 9.37 0.86 18.12 2.73 38.59 1.45 – – – – – 4.73 3.22 1046
BEC 103 – – – – – – 69.60 0.21 −13.7 4258 470 4.77 9.19 49.39 16,046
BEC 107 – – – – – – 89.85 0.26 −15.6 891 −3883 4.65 1.34 33.67 10,937
BEC 110 – – – – – – 54.78 0.19 −16.2 5081 147 4.67 1.19 54.80 17,803
BEC 112 – – – – – – 67.18 0.21 −12.7 3288 −685 4.75 0.80 1.16 378
BEC 119 0.72 0.36 3.22 1.36 52.24 8.42 34.15 0.16 −10.6 11,214 4551 4.71 1.54 56.67 18,411
BEC 126a 4.41 0.63 31.05 5.51 60.03 2.45 43.21 0.18 −17.9 7978 2535 4.79 2.90–1.88 4474–2656 1.45–0.86 Ma
BEC 137 – – – – – – 41.14 0.17 −13.0 7342 2029 4.79 1.01 5.57 1808
BEC 138 – – – – – – 64.77 0.22 −14.9 3590 −676 4.75 2.20 0.77 250
BEC 147 8.58 0.73 5.14 1.46 21.79 1.07 65.95 0.20 −12.8 4933 1027 4.74 1.50 1.28 416
BEC 148 1.73 0.37 2.47 1.41 33.85 3.18 43.17 0.17 −15.6 6953 1653 4.72 2.04 5.60 1819
BEC 149 6.95 0.64 11.78 1.99 36.53 1.43 – – – – – 4.74 0.94 92.27 29,974
F2 PACES 0.85 0.57 2.68 1.28 46.45 10.98 – – – – – 4.73 0.66 15.78 5128
F3 PACES – – – – – – – – – – – 4.78 −4.28 45.14 14,664
F4 PACES 9.09 1.04 7.31 8.16 25.64 1.56 – – – – – 4.29 54.53 3.19 1035
F5 PACES 0.76 0.38 2.67 1.26 48.31 8.30 – – – – – 4.58 0.43 7.75 2517
F9 PACES 4.55 0.85 −0.36 0.92 −6.75 1.54 – – – – – 4.59 −2.36 270.30 87,814
F9b – – 11.20 1.99 (42.51) (3.03) – – – – – 4.56 – – –
a
Helium ages are calculated for samples (1) and (2) respectively. 3H/3He age if from data of sample BEC126_2.
b
Because of the choice of Rterr, it was not possible to calculate a 3H/3He age. Assuming a Rterr of 0.012Ra (Pinti et al., 2011) a minimum age of 42.5 ± 3.0 yrs is obtained.

99
100 G. Vautour et al. / Chemical Geology 413 (2015) 94–106

carrying capacity for water with high efficiency and low background Table 3
(Heemskerk and Johnson, 1998). Noble gas temperatures (NGT) in selected groundwater samples of the Becancour
watershed.
Carbon species (TDIC) were converted into CO2 by direct acidifica-
tion, and the 13C content in the purified CO2 fractions was measured Sample name NGT ± χ2 Excess air ± ΔNe
by mass spectrometry (SIRA) at the IDES Laboratory (University of (°C) (cm3) (%)

Paris Sud, France). 13C content is reported using conventional δ (‰) Quaternary granular aquifer
notation as a deviation from the V-PDB (Vienna-Belemnite from the BEC 105 5.36 0.34 1.90 0.0019 0.0002 14.01
BEC 118 8.29 0.86 11.51 0.0027 0.0005 19.47
Pee Dee formation, North Carolina, USA) (Table 2). Graphite targets
for 14C analyses were prepared from pure CO2 fractions at the IDES Ordovician fractured aquifer
Laboratory, and sources were measured by accelerator mass spectrom- BEC 101 4.32 3.56 205.02 0.0083 0.0026 41.61
BEC 103 2.77 0.54 3.75 0.0175 0.0005 59.66
etry (ARTEMIS accelerator of the UMS LMC14 of INSU-CNRS, Gif-sur-
BEC 107 7.74 0.50 3.89 0.0026 0.0003 18.51
Yvette, France). 14C content is expressed as a percentage of modern BEC 110 7.65 0.66 6.95 0.0023 0.0003 16.69
carbon (pMC) (Table 2). BEC 112 3.15 1.11 23.17 0.0015 0.0006 11.43
In order to calculate the production rate of radiogenic 4He in ground- BEC 119 5.86 0.73 9.04 0.0029 0.0004 20.39
water, the parent elements U and Th were measured from rock cores BEC 126_1 −0.40 3.02 189.49 0.0055 0.0023 31.36
BEC 126_2 −0.11 2.95 186.969 0.0036 0.0020 22.89
recovered from wells of the Becancour watershed (Table 4). U, Th and
BEC 137 2.53 0.54 5.62 0.0019 0.0003 14.07
K concentrations were measured by neutron activation at the Slowpoke BEC 138 2.88 0.69 8.52 0.0042 0.0005 26.33
Laboratory at the University of Montreal. BEC 147 3.68 0.50 4.63 0.0029 0.0003 19.70
BEC 148 3.25 0.34 0.85 0.0039 0.0002 24.95
BEC 149 3.25 0.34 2.23 0.0018 0.0002 13.30
4. Results
F2 PACES 1.68 1.13 26.36 0.0013 0.0007 9.50
F4 PACES 9.75 1.09 2.72 0.1041 0.0016 90.48
In Table 1 He, Ne and Ar, Kr and Xe concentrations and He, Ne and Ar F5 PACES 9.70 0.47 3.48 0.0008 0.0002 6.91
isotopic ratios are reported for the wells analyzed at the University of F9 PACES 14.42 3.20 171.90 −0.0045 0.0008 −73.16
Michigan. He and Ne concentrations and He isotopic ratios are reported
for samples analyzed at the University of Tokyo. Well location, depth
and altitude together with water chemistry type, temperature, Total ΔNe (%), given as ([Ne]measured/[Ne]atmospheric equilibrium − 1 ×100;
Dissolved Solids (TDS) and δ18O and δD are also given in Table 1. Table 2) reflects the total amount of excess air (e.g., Aeschbach-Hertig
Helium concentrations vary over several orders of magnitude, et al., 2001).
from atmospheric (ASW or Air Saturated Water at 10 °C;
4.59 × 10 − 8 cm 3 STP g− 1
rock ) values to 4.48 × 10
−5
cm 3 STP g − 1
rock 5. Discussion
(BEC126_1; Table 1), pointing to significant accumulations of
terrigenic 4 He. BEC010 was sampled twice in 2012 and 2013. The 5.1. 3H/3He ages
4
He concentration in 2012 is 5.25 × 10− 7 cm3 STP g− 1
rock, 35% lower
−7
than the amount measured in 2013 (7.13 × 10 cm 3 STP g − 1
rock ). Helium in groundwater is the sum of several components: the
BEC126 was also sampled twice, showing very large differences above-mentioned Heeq and Heea, together with terrigenic helium
in the amounts measured, from 4.48 × 10− 5 cm 3 STP g − 1
rock to (Heterr) produced by U–Th decay in the crust and/or mantle-derived
−5 3 −1
2.65 × 10 cm STP grock (Table 1). These variations could be due helium, and helium from decay of tritium (3Hetri). The proportions of
to the natural variability over short and long sampling time intervals. tritiogenic and terrigenic helium in the sampled groundwater can be
Indeed, natural concentration variability in other groundwater flow visualized in a Weise-type plot (Fig. 3; Weise and Moser, 1987). In
systems where degassing has not taken place has been previously this plot, the excess-air-corrected helium isotopic ratio (3Hetot–3Heea)/
observed (e.g., Warrier et al., 2012). (4Hetot–4Heea) is plotted versus the inverse of the normalized excess-
3
He/4He (R) values, normalized to that of the atmosphere (Ra = air-corrected helium concentration 4Heeq/(4Hetot–4Heea). An addition
1.386 × 10−6; Clarke et al., 1976), range from R/Ra = 0.039 ± 0.003
(BEC126_21) to 3.109 ± 0.065 (BEC009; Table 1). These ratios point
to the presence of tritiogenic 3He and terrigenic He. Ne isotopic ratios
(20Ne/22Ne from 9.75 to 9.92; 21Ne/22Ne from 0.0273 to 0.0293)
and Ar isotopic ratios (38Ar/36Ar from 0.1829 to 0.1907; 40Ar/36Ar
from 293.9 to 297.6) are all also atmospheric within uncertainties
(Table 1).
Table 2 reports calculated 3H/3He ages and 14C uncorrected and
corrected ages using complex inversion technique models from
NETPATH (Plummer et al., 1994). Tritium was measured in 19 water
samples to calculate 3H/3He ages (Table 2). Measured tritium values
range from the detection limit of 0.8 TU (BEC119 and F5) to 19.3 TU
(BEC118) (Table 2). The uncorrected 14C (A14C) activity ranges from
21.8 ± 0.1 pMC for BEC101 to 119.6 ± 0.3 pMC for BEC118, while
δ13C values range from −17.9‰ to −7.3‰ (Table 2).
The measurements of all noble gases dissolved in groundwater were
used to calculate the recharge noble gas temperatures (NGT) using
the unfractionated air (UA) model (Stute and Schlosser, 1993), imple-
mented in the inverse model of Ballentine and Hall (1999). Resulting
temperatures ranges from −0.4 ± 3.0 °C (BEC126_1) to 14.4 ± 3.1 °C Fig. 3. Weise-type plot of helium measured in Becancour groundwater. The lowermost
line represents the mixing between water at the recharge (Air Saturated Water or ASW
(F9) (Table 3). The model allows one to estimate the helium in solubility
with Req of 0.983Ra) and water enriched in radiogenic 4He (Rterr = 0.012Ra). Dashed
equilibrium with the atmosphere (Air Saturated Water He or Heeq) and lines interpolated through samples represent the addition of 3, 11, 21, 34, 51, 81 and
the excess air helium (Heea) which results from air bubbles entering the 104 TU helium, mixed with the terrigenic components having variable Rterr. Gray stars
water table and dissolved in groundwater (Heaton and Vogel, 1981). represent the ASW and terrigenic helium end-members. Symbols are the same as Fig. 2.
G. Vautour et al. / Chemical Geology 413 (2015) 94–106 101

of only 3Hetri will move the points up, along the right-hand side Y-axis
from an initial ASW isotopic composition (Req = Ra × 0.983; Benson
and Krause, 1980). Addition of 3Heterr produced by nucleogenic
reactions with Li (Morrison and Pine, 1955), 4Heterr from α-decay of U
and Th in crustal rocks or both 3He and 4He from addition of a mantle
component will shift points on the left side of the plot to the local Rterr
value. Rterr corresponds to the time-integrated ratio from 3He and 4He
radiogenic production and/or mantle He addition.
Mixing between the different end-members will be represented by a
straight line of equation Y = mX + b (Weise and Moser, 1987):
  !
3
Hetot −3 Heea 3
Hetri 4
Heeq
  ¼ Req −Rterr þ 4 4 þ Rterr : ð1Þ
4
Hetot − Heea 4 Heeq Hetot −4 Heea |ffl{zffl}
|fflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflffl} b
m X
Y

In a Weise-type plot, samples align on several straight lines


representing the mixing between groundwater accumulating 3Hetri
Fig. 4. Initial tritium content (3H + 3Hetri) of groundwater compared with 3H in precipita-
decayed from a constant amount of tritium and terrigenic helium tion as recorded at the GNIP station of Ottawa (IAEA/WMO, 2004). The curve that fits the
(Fig. 3). From the slope “m” and the Y-intercept “b” of each mixing measured 3H at the GNIP station represents the yearly average. Symbols are the same as
line it is possible to calculate the Rterr and the initial 3H amount. Water Fig. 2.
samples plot on mixing lines corresponding to the decay of 3 to 104
TU 3H and with Rterr ranging between 5.99 × 10−8 and 2.95 × 10−7. As-
suming a typical crustal Rterr of the St. Lawrence Platform fractured bed- The majority of samples fall within the calculated yearly-averaged
rock is 1.66 × 10−8 (Pinti et al., 2011), then the calculated Rterr from Ottawa tritium input curve. Samples BEC148, BEC119, F2 and F5
each mixing line suggest the addition of 0.4 to 2.5% of mantle helium fall below the line, suggesting a dilution of the initial 3H input by pre-
(Fig. 3). Although small, the mantle component is nevertheless non- bomb water (Aeschbach-Hertig et al., 1998), which is virtually free of tri-
negligible. Interestingly, a mantle helium of ~2–3% was also identified tium. The dashed curve represents a two-component mixture containing
in deeper brines residing at the base of the Ordovician geological forma- 25% of water recharged at the time and 75% of pre-bomb water, and can
tions of the St. Lawrence Platform near the Becancour municipality explain the low 3Hinit recorded in BEC148. The dashed and double dotted
(Pinti et al., 2011), in the same watershed (Fig. 1). curve represents a two-component mixture containing 2% of water
To calculate the 3H/3He ages, 3Hetri was estimated following the recharged during the observed period and 98% of pre-bomb water, and
equation (Schlosser et al., 1989): can explain the low 3Hinit calculated for BEC119, F2 and F5.

 
3
Hetrit ¼ 4 Hetot  ðRtot −Rterr Þ−4 Heeq  Req −Rterr ð2Þ 5.2. 14C ages
!
4
He 
− 20  20 Netot −20 Neeq Þ  ðRea −Rterr Þ Uncorrected 14C activities (A14C) show a rough inverse trend with
Ne the HCO− 13
ea 3 content (Fig. 5a), as well as with the δ C value of the TDIC
(Fig. 5b). These correlations suggest an evolution of groundwater with
where Rtot is the measured 3He/4He ratio; Req = α*Ratm with α age, with post-bomb freshwater (represented by sample BEC118)
representing a fractionation factor of 0.983 (Benson and Krause, dissolving biogenic soil CO2 (A14C = 120 pMC, Levin et al., 1992;
1980). Ratm and (4He/20Ne)ea are generally assumed to be atmospheric δ13C = − 20‰, Taupin, 1990; HCO− 3 ~ 0), which evolves with time
(Ratm = 1.386 × 10−6; [4He/20Ne]exc = 0.3185; Ozima and Podosek, and accumulates dead carbon from carbonate dissolution (A14C = 0
1983). 4Hetot and 20Netot is the measured amount of 4He and 20Ne in pMC; δ13C = 0‰; HCO− 3 ≫ 0; Le Gal Lasalle et al., 2001; sample
the groundwater sample (Table 1). 4Heeq and 4Heea are calculated BEC101). Uncorrected A14C vs. δ13C suggests the occurrence of a third
from the recharge temperatures (NGT; Table 3). 20Neeq is calculated as source of carbon in the watershed (Fig. 5b). This reservoir is also
20
Neeq = 20Netot − 4Heeq/(4He/20Ne)ea = air. For samples BEC09, 010, enriched in dead carbon (A14C ~ 0 pMC; Fig. 4) but it displays a δ13C of
021, 034, 102, 106, 113, 117 and 124, only the 20Ne amount has been less than −20‰, indicating an organic origin.
measured. In this case, the 4Heeq (and 20Neeq) is calculated using This third dead carbon reservoir could be related to thermogenic
solubility data from Smith and Kennedy (1983) on the basis of the methane. Groundwater from the regional fractured aquifer of the St.
present day temperature measured in the well (Table 1). Because the Lawrence Lowlands contains methane with measured concentrations
He and Ne solubility variability is relatively small at the expected in the Becancour wells ranging from 0.0008 to 32.7 mg L−1 (BEC101)
recharge temperatures of 0–15 °C (Table 3), this approximation can be with a median value for the watershed of 0.133 mg L−1 (Moritz et al.,
considered acceptable. Uncertainty in the calculated 3Hetri is the propa- 2015). Measured δ13C ranges from − 103.6 to − 24.8‰ indicating the
gated error taking into account all uncertainties on each variable of mixing between biogenic methane produced by methanogenesis and
Eq. (2). thermogenic methane produced in the Ordovician age Utica and
Calculated 3H/3He apparent ages are reported in Table 2. The youn- Lorraine shales. These reservoirs contain methane with a δ13C varying
gest 3H/3He apparent age of 2.7 ± 0.1 yrs is from the shallowest water between −30‰ and −40‰ (Moritz et al., 2015). The sample that devi-
sample (BEC118) from the granular Quaternary aquifer (Fig. 1). The ates most from a simple binary mixing between a dead carbon reservoir
oldest 3H/3He age is 60.0 ± 2.5 yrs and was measured in sample (Ordovician limestone) and biogenic CO2 from soil respiration is
BEC126_2 from the fractured bedrock aquifer, in the Appalachian BEC126 (Fig. 5b) which has a methane concentration (2.4 mg L− 1),
recharge area (Fig. 1). To evaluate the validity of 3H/3He groundwater well above the expected natural background level of 0.133 mg L−1.
ages, the initial amount of 3Hinit = 3H + 3Hetri prior to decay was BEC126 also displays a δ13C(CH4) value of −59.4‰, suggesting a mixed
compared with the 3H in precipitation at the Ottawa GNIP station source of methane (Moritz et al., 2015).
(IAEA/WMO, 2004; Fig. 4). The curve that fits the measured 3H at the During methanogenesis, CO2 reduction is accompanied by strong
GNIP station represents the yearly average. carbon fractionation of the DIC. For an initial organic substratum with
102 G. Vautour et al. / Chemical Geology 413 (2015) 94–106

Here we use NETPATH-WIN® software (El-Kadi et al., 2011) to calcu-


late uncorrected and adjusted 14C ages (Table 2). NETPATH uses water
chemistry and δ13C ratios to correct 14C activity by accounting for
mass transfers of carbon that occur along a flow path between an initial
water location, where the initial 14C activity (A0) is fixed, and a final
water composition, where the 14C activity has been measured (Aobs).
NETPATH calculates the term AND, which is the 14C activity at the final
well, adjusted for all inorganic and organic chemical reactions, but not
radioactive decay (El-Kadi et al., 2011). The adjusted 14C age (14CADt)
will be the age increment between initial and final wells calculated as
follows (El-Kadi et al., 2011):
 
14 5730 AND
CAD t ¼ ln : ð3Þ
ln2 Aobs

Ideally, young but tritium-free pre-bomb water should be used as


the initial water where the 14C activity A0 is fixed. Based on the trends
exhibited in Fig. 5a–b, BEC118 is assumed to represent the initial
water even though it contains a small amount of post-bomb tritium
and post-bomb 14C (Tables 2, 3). Indeed, no entirely tritium free pre-
bomb samples are available. Further, the initial and final water chemical
compositions are expected to represent the water chemical/isotopic
evolution along a flow path from the recharge to discharge area. Our
end-members (BEC118, BEC101, BEC126; Fig. 5) represent the shallow
and deep compositions in the down-gradient portion of the basin.
Uncorrected 14C ages obtained with NETPATH range from 15,426 yrs
for BEC101 to 208 yrs for BEC105. Uncertainties on the 14C uncorrected
ages are estimated to be ±0.5% of the value and related to the analytical
uncertainties of the measured A14C (Table 2). Adjusted 14C between
BEC118 and individual wells, using the F & G model implemented in
NETPATH, range from 6696 yrs for BEC101 to 147 yrs for BEC110. Neg-
ative 14C ages for BEC105, 107, 112 and 137 suggest groundwater infil-
trated since the mid-1950s (Table 2). Uncertainties in the adjusted ages
depend essentially on the assumed δ13C of soil CO2, as calculated with
the thermodynamics model 3 (dissolved CO2 is in isotopic equilibrium
Fig. 5. Inverse trends between the measured 14C activity (A14C) and HCO−
3 concentrations with the soil gas) and model 4 (recharge water evolved under open sys-
(in mg L−1) of groundwater (a); and the measured δ13C of the TDC (b). Dead carbon tem conditions) proposed in NETPATH, which is the most realistic
14
reservoir isotopic composition and C activities are taken from Taupin (1990) (soil
CO2); Le Gal Lasalle et al. (2001) (carbonates); and Moritz et al. (2015) (thermogenic
(Plummer and Glynn, 2013). The average uncertainty on the adjusted
methane δ13C for Utica and Lorraine shale). ages has been evaluated to be ±20% of the value (Table 2).
Uncorrected 14C values are in the range of those calculated by
Beaudry (2013) in the same Ordovician regional fractured aquifer of
the St. Lawrence Lowlands in the eastern portion of the Monteregie
δ13C = − 25‰, methane and DIC products would have δ13C values of region (14C from 13,800 to 910 yrs; Beaudry, 2013), located approxi-
− 60‰ and + 10‰ respectively (e.g., Whiticar, 1999). This is not mately 100 km SW of the Becancour watershed. Uncorrected older 14C
observed here (Fig. 5b). However, Moritz et al. (2015) have shown ages of 15 to 13 kyrs (this study; Beaudry, 2013) are roughly coincident
that most of the methane in St. Lawrence Lowlands has been affected with the end of the glaciation and the ice retreat of the 3 km thick
by oxidation. During methane oxidation, strong fractionation of meth- Laurentide Ice sheet in the region. This was followed by a glacio-
ane is not associated with that of DIC (ε 13CCO2–CH4 = 5‰; Whiticar, isostatic marine transgression, known as the Champlain Sea, which
1999). This process could explain why this old carbon reservoir invaded the area at 12.8–12.3 kyrs (Parent et al., 1985). Between 10.6
(A14C ~ 0 pMC) has an unfractionated δ13C of less than −20‰ (Fig. 5b). and 6.7 kyrs, the main phase of isostatic rebound lowered the St.
The initial 14C activities (A14
o C) of Becancour groundwater have been Lawrence River base level from + 60 masl to approximately 16 masl,
modified by the numerous water–rock interactions affecting the studied followed by some fluctuations until 1000 yrs ago (Lamarche et al.,
aquifers, including calcite and dolomite dissolution and Ca/Na ion 2007). At 6.7 kyrs, the hydrographic network of the St. Lawrence Valley
exchange (Meyzonnat et al., in press), as well as isotope exchange reached a configuration close to the present one. It is expected that dur-
between soil CO2, calcite and DIC during recharge, and methanogenesis ing this accelerated isostatic rebound period, new emerging recharge
(Geyh and Künzl, 1981). areas and increased potentiometric heads favored a large meltwater in-
There are several well-known methods for calculating A0 and thus vasion into the Quaternary shallower aquifers, and the confined aquifers
14
C ages (see e.g., Fontes, 1992; Plummer and Glynn, 2013 for a review). of the St. Lawrence Basin. Most of the groundwater samples from the
The method of Fontes and Garnier (1979) takes into account the disso- fractured bedrock aquifer of Becancour show 14C uncorrected ages
lution of carbonates and Ca/Na exchange but it is limited to inorganic within this period of regional re-organization of the hydrological system
carbon dissolution (DIC) reactions (Plummer and Glynn, 2013). For (Table 2). Calculated noble gas recharge paleotemperatures (NGT;
more complex aquifers in which carbonate systems undergo complex Table 3) do not show a clear trend with the uncorrected 14C or the
geochemical reactions, such as oxidation of organic carbon and adjusted 14C ages. The majority of the groundwater samples from the
methanogenesis, adjusted radiocarbon ages are best estimated by fractured bedrock aquifer show NGTs from − 0.40 ± 3.02 to 5.86 ±
inverse geochemical modeling techniques such as those implemented 0.73 (Table 3), 4 to 9 °C below the present-day recharge temperature
in the software NETPATH (Plummer et al., 1994; Aravena et al., 1995). of 8–10 °C, suggesting that this groundwater recharged in a cooler
G. Vautour et al. / Chemical Geology 413 (2015) 94–106 103

period, possibly at temperature approaching 1 °C. NGT measured in and younger) into the fractured reservoir. This is not surprising for
shallower aquifers in the Michigan Basin have already revealed a BEC126, which is located in the main recharge area (Fig. 1) with the
ground temperature of 1 °C toward the end of the Last Glacial Maximum fractured aquifer directly exposed to precipitation, with little or no
suggesting that groundwater recharge occurred under the Laurentide Quaternary cover. Direct infiltration of modern precipitation would
Ice Sheet cover (Ma et al., 2004). also explain why BEC126, which contains the highest amount of radio-
Calculated 14C groundwater ages for samples BEC119, BEC148 and genic 4He measured in Becancour (4.48–2.65 × 10−5 cm3 STP g−1;
BEC126 are much older than the calculated 3H/3He ages for the same Table 1) which requires a long accumulation (residence) time in the
wells (Table 2). This apparent contradiction in age could result from aquifer, also shows an appreciable amount of tritium (4.2 TU; Table 2).
mixing of old groundwater with modern water, as suggested by the de-
viation of the 3Hinit composition from the yearly-averaged Ottawa-GNIP 5.3. Radiogenic 4He in the aquifers: the old water component
tritium input curve for samples BEC148 and BEC119 (Fig. 4). This mixing
could take place along the well casing or within the fractured aquifer. The relation between the 14C uncorrected ages and the 3He/4He
Mixing of groundwater from the granular aquifer with fossil water ratios (Fig. 6) supports the hypothesis of an increasing contribution of
from the fractured aquifer within the well is unlikely, because the well radiogenic 4He to groundwater during its journey in the fractured
casing prevents groundwater from the granular deposits. Thus we can regional bedrock aquifer. The two end-members, BEC101 and BEC101,
assume that this mixing is produced within the fractured regional are the two samples contain the highest radiogenic 4He content, from
aquifer as in situ groundwater mixes with recharge water. 1.17 to 4.48 × 10−5 cm3 STP g−1 (Table 1). This means that groundwa-
The relation between uncorrected 14C ages and the 3He/4He ratios ter in these locations resided in the aquifer long enough to accumulate
(Fig. 6) highlights the mixing between young and old groundwater. significant amounts of radiogenic 4He produced from decay of U and
One trend corresponds to the mixing between freshwater from the Th contained in the aquifer rocks (Andrews and Kay, 1982) or from a
Quaternary shallower aquifer (BEC118) showing a nearly atmospheric sustained crustal flux from underneath the aquifer (e.g., Torgersen and
3
He/4He ratio of 1.077 ± 0.037 Ra and modern 14C ages (Fig. 6) and Clarke, 1985).
fossil water from the fractured bedrock aquifer (BEC126/2) enriched Assuming that the measured radiogenic 4He is produced from U
in radiogenic 4He (3He/4He = 0.054 ± 0.005) and with an uncorrected and Th decay contained in the aquifer rocks (in situ production), the
14
C age of ~ 7980 yrs. A second trend corresponds to the mixing of a U–Th/4He groundwater residence times can be calculated as follows
more evolved tritiated groundwater (3He/4He ≥ 1.5Ra; Fig. 6) (Torgersen and Clarke, 1985):
recharging the fractured bedrock aquifer in the Appalachian recharge
h i
area (BEC138; see Fig. 1A for location) which mixes with a fossil 4
Heterr
groundwater (14C uncorrected ages of 15 kyrs) contained in the frac- t¼   ð4Þ
tured bedrock aquifer and enriched in 4Heterr (BEC101; 3He/4He = 1−ϕ
P4 He  Λ 4 He  ρ
0.074 ± 0.003 Ra; Fig. 6). ϕ
Similarly to Fig. 5, it is worth noting that these trends do not repre-
sent the chemical/isotopic evolution of groundwater along a flow-path where [4Heterr] is the measured radiogenic 4He concentration in
from the recharge to the discharge, but rather reflect local flow direc- groundwater (cm3 STP g− 1
water; Table 2); P4He is the radiogenic He
4

tions. This could be partially caused by the fact that recharge of the production rate in the fractured bedrock (cm3 STP g− 1
rock ); Λ4He is the
deeper regional fractured aquifer is not uniform but is produced He retention factor (4Hereleased/4Heproduced) taken as 1 (Torgersen,
through direct infiltration in the main Appalachian recharge area and 1980; Sano et al., 2008); (1 − ϕ/ϕ) is the void ratio where ϕ is the frac-
in the St. Lawrence plain through unconfined or semi-confined portions tional effective porosity; and ρ is the aquifer matrix density (assumed to
of the Quaternary granular aquifers (Larocque et al., 2013). Thus, the be 2.72 g cm−3 for a dominant carbonate matrix and 2.65 g cm−3 for a
trend between BEC118 (located downgradient and in the granular aqui- sand matrix). Effective porosities of the Ordovician silty-carbonate-
fer) and BEC126 (located upgradient in the fractured aquifer) should be shale facies of the regional fractured aquifer from 1 to 5% have been
interpreted as a direct infiltration of modern precipitation (mid-1950s estimated from pumping well tests (Larocque et al., 2013) and
measured on core samples (Tran Ngoc et al., 2014). Effective porosities
of the Quaternary granular aquifers range between 10 and 20%
(Larocque et al., 2013). 4He production rate (P4He) were estimated on
the basis of U and Th amounts measured by neutron activation and
ICP-MS analyses on eight aquifer cores sampled from domestic wells
(BEC101, BEC105 and BEC107) and from monitoring wells (F1, F2, F5,
F7 and F10) in the study area (Table 4). U contents range from 0.85
to 2.98 ppm and those of Th range from 1.3 to 12.9 ppm. Calculated
P4He assumes secular equilibrium among the U and Th descendants
(see Ballentine and Burnard, 2002), and range between 1.39 and
6.84 cm3 STP g−1 rock yr
−1
with a geometric mean and average deviation
of 3.5 ± 1.4 × 10−13 cm3 STP g−1 rock yr
−1
. This P4He production rate is
−13
close to that of 3.27 × 10 cm STP g−
3 1
rock yr
−1
calculated by Pinti
et al. (2011) for limestone facies of the St. Lawrence Lowlands.
Average in situ U–Th/4He water ages were calculated using a mean
porosity value of 3% for the fractured bedrock aquifer (and a density of
2.72 g cm− 3) and of 15% for the granular aquifer (and a density of
2.65 g cm−3). Taking into account uncertainties in the variables used
in Eq. (4), particularly the P4He (40%) and effective porosities (50%), it
Fig. 6. Helium isotopic ratios 3He/4He plotted against the uncorrected 14C ages of ground- is assumed here that the calculated U–Th/4He groundwater ages have
water. The two trends indicate mixing between modern precipitation and groundwater
a total uncertainty of at least 50%, as generally assumed in these age
infiltrated in the mid-50s with a “fossil” water infiltrated the system since the end of
deglaciation. The increase in ages and in radiogenic 4He (as indicated by the decrease in models (e.g., Torgersen and Clarke, 1985; Marty et al., 1993).
the 3He/4He ratio) corresponds to a chemical evolution of groundwater from freshwater Calculated in situ U–Th/4He groundwater ages (Table 2) range from
dissolving carbonates to groundwater affected by cation exchange and NaCl dissolution. 311 yrs for BEC034 to 1.45 Ma for BEC126 (Table 2). U–Th/4He ages have
104 G. Vautour et al. / Chemical Geology 413 (2015) 94–106

Table 4 line (Fig. 7). The first group of samples requires very low helium
U and Th contents measured in core rock from selected wells. production rates P4He from 2.5 to 9 × 10−14 cm3 STP g− 1
rock yr
−1
with
−14 3 −1 −1
Rock type Weight U Th P4He x 10−13 cm3 an average of 5.8 × 10 cm STP grock yr (Fig. 7b). If the corrected
(g) (ppm) (ppm) STP g−1 yr−1 F&G 14C are valid (Table 2), then the P4He required to explain the mea-
BEC 101 2.345 1.19 5.36 2.96 sured helium amounts are from 1.4 to 3.85 × 10−13 cm3 STP g−1 rock yr
−1
,
BEC 105 2.488 2.98 8.91 6.11 again within the range of expected production rates in the fractured
BEC 107 2.203 2.63 12.90 6.84 bedrock (Table 4).
F1 2.058 1.11 6.48 3.19
The second group of water samples that plot above the 1:1 reference
F2 2.635 1.16 5.24 2.89
F5 2.082 1.36 5.89 3.31 line (Fig. 7a) points to an additional helium source in the Becancour
F7 2.290 2.00 6.67 4.30 aquifers capable of explaining the large 4Heterr amounts measured and
F10 2.362 0.85 1.30 1.39 the U–Th/4He ages higher than the equivalent 14C ages. This additional
Average 1.515 5.670 3.50 ± 1.41 helium source can be ascribed to a helium flux entering the base of
the aquifer, following the model of Torgersen and Clarke (1985). In
this case, Eq. (3) should be expressed as (e.g., Kulongoski and Hilton,
been plotted against uncorrected 14C ages (Fig. 7a). The dashed line is 2011):
the 1:1 reference line (Kulongoski et al., 2003). If the U–Th/4He and
14
C ages are concordant, all points should plot on this line. Only 4

Heterr
BEC119 plots close to the 1:1 line suggesting that radiogenic production t¼   ð5Þ
100−ϕ J0
of 4He in the aquifer rocks could be sufficient to explain the measured P4He  Λ 4 He   ρrock þ
4 ϕ zx ϕρwater
Heterr amounts (Fig. 7a). To accumulate the measured radiogenic 4He
amount of 5.67 × 10−7 cm3 STP g−1 H2O in 11.2 ka (uncorrected
14
C age;
Table 2) it requires a P4He of 5.7 × 10−13 cm3 STP g−1 rock yr −1
, within where ρwater is assumed to be equal to 1 g cm−3 and zx is the distance
the expected range of helium production rates reported in Table 4. from the base of the aquifer, where the external flux enters, to the
The remaining samples show two distinct trends (Fig. 7). A first well screen. Because the wells have no casing along the fractured
group of samples (BEC112, BEC137, BEC138, BEC147 and BEC148) bedrock aquifer section, zx is here assumed as the half-distance between
have calculated U–Th/4He water ages that are younger than the calcu- the bottom-hole and the top of the fractured bedrock aquifer.
lated 14C ages and thus plot below the 1:1 reference line. A second Data from Table 1 indicate that the fractured bedrock aquifer has an
group of water samples (BEC101, BEC103, BEC105, BEC107, BEC110, average thickness of 28 m, thus a mean zx of 14 m is here assumed.
BEC119, BEC126) show calculated U–Th/4He water ages higher than J0 is the 4Heterr flux entering the aquifer across the bottom boundary
the corresponding 14C ages and thus they plot above the 1:1 reference (in cm3 STP cm−2 yr−1).

Fig. 7. Plots of corrected U–Th/4He groundwater residence times (porosity ϕ = 3%) vs. 14C uncorrected ages for in situ production (helium production rate of 3.50 × 10−13 cm3 STP g−1
rock yr
−1
) (a);
in situ production and helium production rate of 5.8 × 10−14 cm3 STP g−1
rock yr
−1
(b); and for two crustal fluxes J0 of 1.0 × 10−8 cm3 STP cm−2 yr−1 (c) and of 18 × 10−7 cm3 STP cm−2 yr−1 (d).
G. Vautour et al. / Chemical Geology 413 (2015) 94–106 105

The best fit between U–Th/4He and 14C ages for samples BEC103, Acknowledgments
BEC107, BEC110 and BEC101 is obtained by adding a helium flux J0 of
1.0 × 10− 8 cm3 STP cm−2 yr−1 (Fig. 7c) To explain the high 4Heterr The authors thank the Quebec Ministry of Environment (Ministère
amounts of BEC101 and BEC126 (Table 2), a helium flux J0 of du Développement durable, de l'Environnement, des Parcs et de la
1.8 × 10−7 cm3 STP cm−2 yr−1 is required (Fig. 7d). These again are Lutte contre les Changements climatiques), the Quebec Research Fund
minimum estimates because they are based on uncorrected 14C ages. (“Fonds de recherche du Quebec - Nature et Technologies”), the
If the corrected 14C ages are assumed to be valid, then the minimum Becancour River Watershed Organization (“organisme de bassin versant
and maximum flux entering the bedrock aquifer would range from to GROBEC”) and the municipalities that contributed with funds to this re-
9.0 × 10−8 to 5.9 × 10−7 cm3 STP cm−2 yr−1. search. The authors thank the UMR GEOPS, the INSU and the Artemis of
Calculated minimum 4Heterr fluxes J0 are 18 to 330 times lower LMC14 for 14C analyses. DLP thanks Tokyo University and Mombusho
than the average helium continental crustal flux of 3.3 × for supporting his sabbatical stay at AORI.
10−6 cm3 STP cm− 2 yr−1 (O'Nions and Oxburgh, 1983) and even
lower (32–583 times) than the helium flux calculated for the Canadian
Precambrian Shield (5.83 × 10−6 cm3 STP cm−2 yr−1; Torgersen, 2010). References
Assuming a P4He of 3.5 × 10−13 cm3 STP g−1 rock yr
−1
and a rock density of
−3 4 Aeschbach-Hertig, W., Schlosser, P., Stute, M., Simpson, H.J., Ludin, A., Clark, J.F., 1998. A
2.72 g cm , the minimum Heterr fluxes J0 can be sustained by the total 3
H/3He study of groundwater flow in a fractured bedrock aquifer. Ground Water
4
He radiogenic production in a sedimentary layer of 100 to 1900 m, 36, 661–670. http://dx.doi.org/10.1111/j.1745-6584.1998.tb02841.x.
much less of the total thickness of the Cambro-Ordovician St. Lawrence Aeschbach-Hertig, W., Beyerle, U., Holocher, J., Peeters, F., Kipfer, R., 2001. Excess air in
groundwater as a potential indicator of past environmental changes. International
Platform sedimentary sequence which is evaluated to 2900 m (Lavoie Conference on the Study of Environmental Change Using Isotope Techniques IAEA-
et al., 2013). Maximum 4Heterr fluxes J0 are equivalent to the radiogenic CN-80/29. IAEA, Wien, pp. 34–36.
production of a layer of 950–6200 m, which implies an additional pro- Aggarwal, P.K., Matsumoto, T., Sturchio, N.C., Chang, H.K., Gastmans, D., Araguas-Araguas,
L.J., Jiang, W., Lu, Z.-T., Mueller, P., Yokochi, R., Purtschert, R., Torgersen, T., 2015.
duction from a 2 km-thick section in the Grenvillian crystalline base- Continental degassing of 4He by surficial discharge of deep groundwater. Nat. Geosci.
ment rocks (P4He of 6.4 × 10−13 cm3 STP g−1 rock yr
−1
; Pinti et al., 2011). 8, 35–39. http://dx.doi.org/10.1038/ngeo2302.
These helium fluxes suggest that local sources rather than crustal Andrews, J.N., 1985. The isotopic composition of radiogenic helium and its use to study
groundwater movement in confined aquifers. Chem. Geol. 49, 339–351. http://dx.
fluxes can explain the few anomalous helium amounts measured in
doi.org/10.1016/0009-2541(85)90166-4.
Becancour groundwater. A possible source of radiogenic helium could Andrews, J.N., Kay, R.L.F., 1982. Natural production of tritium in permeable rocks. Nature
be the organic-rich (0.5–2% TOC; Lavoie et al., 2013), 400 m-thick 298, 361–363. http://dx.doi.org/10.1038/298361a0.
Lorraine Shale that partially releases methane in the aquifers of the Aravena, R., Wassenaar, L.I., Plummer, L.N., 1995. Estimating 14C groundwater ages in a
methanogenic aquifer. Water Resour. Res. 31, 2307–2317. http://dx.doi.org/10.
region (Moritz et al., 2015) or the 100 m thick Utica Shale. Gas analyses 1029/95WR01271.
in shale gas wells showed that helium occurs together with shale gas Ballentine, C.J., Burnard, P.G., 2002. Production, release and transport of noble gases in the
methane at concentrations between 0.01 and 0.05 mol%. Local fractur- continental crust. Rev. Mineral. Geochem. 47, 481–538. http://dx.doi.org/10.2138/
rmg.2002.47.12.
ing could enhance the release of helium into groundwater (together Ballentine, C.J., Hall, C.M., 1999. Determining paleotemperature and other variables by
with methane), creating anomalous concentrations of helium in using an error-weighted, nonlinear inversion of noble gas concentrations in water.
groundwater. Methane and helium surveys in groundwater from the Geochim. Cosmochim. Acta 63, 2315–2336. http://dx.doi.org/10.1016/S0016-
7037(99)00131-3.
Becancour and surrounding area's fractured aquifers have shown that Beaudry, C., 2013. Hydrogéochimie de l'aquifère rocheux régionale en Montérégie Est,
both methane (Moritz et al., 2015) and helium (Pinti et al., 2013) con- Québec (MS Thesis) INRS - Centre Eau Terre Environnement, Québec.
centrations in groundwater increase linearly toward the major faults Benson, B.B., Krause, D., 1980. Isotopic fractionation of helium during solution: a probe for
the liquid state. J. Solut. Chem. 9, 895–909.
present in the region. Interestingly, the two samples with the highest Bethke, C.M., Johnson, T.M., 2008. Groundwater age and groundwater age dating. Ann.
4
Heterr concentrations, i.e BEC126 and BEC101 are close to two major Rev. Earth Planet. Sci. 36, 121–152. http://dx.doi.org/10.1146/annurev.earth.36.
tectonic features, the Foulon Fault and the Logan Fault (Fig. 1) and 031207.124210.
Carrera, J., Neuman, S.P., 1986. Estimation of aquifer parameters under transient and
groundwater from these samples contains methane concentrations
steady-state conditions: 1. Maximum likelihood method incorporating prior informa-
from 2.4 to 32 mg L−1 (Moritz et al., 2015). Pinti et al. (2013) found tion. Water Resour. Res. 22, 199–210. http://dx.doi.org/10.1029/WR022i002p00199.
that anomalous areas of radiogenic helium dissolved in groundwater Castro, M.C., Stute, M., Schlosser, P., 2000. Comparison of 4He ages and 14C ages in simple
in the St. Lawrence Lowlands aquifers are also rich in methane, suggest- aquifer systems: implications for groundwater flow and chronologies. Appl.
Geochem. 15, 1137–1167. http://dx.doi.org/10.1016/S0883-2927(99)00113-4.
ing a common source. Castro, M.C., Goblet, P., 2003. Calibration of regional groundwater flow models—working
toward a better understanding of site-specific systems. Water Resour. Res. 39. http://
6. Conclusions dx.doi.org/10.1029/2002WR001653.
Castro, M.C., Goblet, P., 2005. Calculation of groundwater ages — a comparative analysis.
Ground Water 43 (3), 368–380.
Results from this study showed a complicated flow system in the Castro, M.C., Ma, L., Hall, C.M., 2009. A primordial, solar He–Ne signature in crustal fluids
Becancour watershed with three isotopically distinct water masses of a stable continental region. Earth Planet. Sci. Lett. 279, 174–184. http://dx.doi.org/
10.1016/j.epsl.2008.12.042.
mixing together in the regional bedrock aquifer. The mixing between Clarke, W.B., Jenkins, W.J., Top, Z., 1976. Determination of tritium by mass-spectrometric
modern water that infiltrated since the mid-50s and fossil water, possi- measurement of He-3. Int. J. Appl. Radiat. Isot. 27, 515–522.
bly recharged at the end of the deglaciation period in the region, has Cloutier, V., Lefebvre, R., Savard, M.M., Bourque, É., Therrien, R., 2006. Hydrogeochemistry
and groundwater origin of the Basses–Laurentides sedimentary rock aquifer system,
been revealed by 3H/3H and14C ages and helium isotopes. The calculated St. Lawrence Lowlands, Quebec, Canada. Hydrogeol. J. 14, 573–590.
ages are compatible with the chemical evolution of the waters, yet the Cloutier, V., Lefebvre, R., Savard, M.M., Therrien, R., 2010. Desalination of a sedimentary
observed isotopic mixings do not correspond to an evolution along the rock aquifer system invaded by Pleistocene Champlain Sea water and processes
controlling groundwater geochemistry. Environ. Earth Sci. 59, 977–994.
flow path. This is probably caused 1) by a bias related to the sampling
Cook, P.G., Solomon, D.K., 1997. Recent advances in dating young groundwater:
density and 2) by local recharge areas which complicate, isotopically chlorofluorocarbons, 3H/3He and 85Kr. J. Hydrol. 191, 245–265. http://dx.doi.org/10.
and chemically, the deconvolution of groundwater flow paths. The 1016/S0022-1694(96)03051-X.
regional fractured bedrock aquifer of the St. Lawrence Lowlands is the El-Kadi, A.I., Plummer, L.N., Aggarwal, P., 2011. NETPATH-WIN: an interactive user version
of the mass-balance model, NETPATH. Ground Water 49, 593–599. http://dx.doi.org/
main local water resource for drinking, agricultural and industrial 10.1111/j.1745-6584.2010.00779.x.
purposes, including the extraction of fluids for hydraulic fracturing in Fontes, J.Ch., 1992. Chemical and isotopic constraints on 14C dating of groundwater. In:
shale gas exploitation. The occurrence of old groundwater in this aquifer Taylor, R.E., Long, A., Kra, R.S. (Eds.), Radiocarbon After Four Decades. Springer-
Verlag, New-York, pp. 242–261.
system clearly limits the renewable resource and increases the risk of Fontes, J.Ch., Garnier, J.M., 1979. Determination of the initial 14C activity of the total
overexploitation in the case of increased uses or in the case of pollution dissolved carbon: a review of the existing models and a new approach. Water Resour.
from different sources. Res. 15, 399–413. http://dx.doi.org/10.1029/WR015i002p00399.
106 G. Vautour et al. / Chemical Geology 413 (2015) 94–106

Geyh, M.A., Künzl, R., 1981. Methane in groundwater and its effect on 14C groundwater Moritz, A., Hélie, J.-F., Pinti, D.L., Larocque, M., Barnetche, D., Retailleau, S., Lefebvre, R.,
dating. J. Hydrol. 52, 355–358. http://dx.doi.org/10.1016/0022-1694(81)90181-5. Gélinas, Y., 2015. Methane baseline concentrations and sources in shallow aquifers
Globensky, Y., 1987. Géologie des Basses-Terres du Saint-Laurent, Quebec. Ministère des from the shale gas-prone region of the St. Lawrence Lowlands (Quebec, Canada).
Richesses Naturelles du Quebecp. 63 (v. MM 85-02). Sci. Total Environ. 49, 4765–4771. http://dx.doi.org/10.1021/acs.est.5b00443.
Globensky, Y., 1993. Lexique stratigraphique canadien. Volume V-B: région des Morrison, P., Pine, J., 1955. Radiogenic origin of the helium isotopes in rocks. Ann. N. Y.
Appalaches, des Basses-Terres du Saint-Laurent et des Iles de la Madeleine. Ministère Acad. Sci. 62, 69–92. http://dx.doi.org/10.1111/j.1749-6632.1955.tb35366.x.
de l'Énergie et des Ressources et Direction Générale de l'Exploration géologique et O'Nions, R.K., Oxburgh, E.R., 1983. Heat and helium in the Earth. Nature 306, 429–431.
minérale, p. 327 (DV 91e23). http://dx.doi.org/10.1038/306429a0.
Godbout, P.M., 2013. Géologie du Quaternaire et hydrostratigraphie sur les versants de la Ozima, M., Podosek, F.A., 1983. Noble Gas Geochemistry. Cambridge University Press,
zone Becancour, Quebec (MSc Thesis) Université du Québec à Montréal, Quebec, Cambridge.
Canada. Parent, M., Dubois, J.M.M., Bail, P., Larocque, A., Larocque, G., 1985. Paléogéographie du
Goode, D.J., 1996. Direct simulation of groundwater age. Water Resour. Res. 32, 289–296. Quebec méridional entre 12 500 et 8000 ans BP. Recherches amérindiennes au Quebec
http://dx.doi.org/10.1029/95WR03401. 15, 17–37.
Harvey, C.F., Gorelick, S.M., 1995. Mapping hydraulic conductivity: sequential condition- Phillips, F.M., Castro, M.C., 2003. Groundwater dating and residence–time measurements.
ing with measurements of solute arrival time, hydraulic head, and local conductivity. Treatise Geochem. 5, 451–497. http://dx.doi.org/10.1016/B0-08-043751-6/05136-7.
Water Resour. Res. 31, 1615–1626. Pinti, D.L., Marty, B., 1998. The origin of helium in deep sedimentary aquifers and the
Heaton, T.H.E., Vogel, J.C., 1981. “Excess air” in groundwater. J. Hydrol. 50, 201–216. problem of dating very old groundwaters. In: Parnell, J. (Ed.), Dating and Duration
http://dx.doi.org/10.1016/0022-1694(81)90070-6. of Fluid Flow and Fluid–Rock Interaction. Spec. Publ. Geol. Soc. London 144,
Heemskerk, A.R., Johnson, J., 1998. Tritium Analysis: Technical Procedure 1.0. University pp. 53–68. http://dx.doi.org/10.1144/GSL.SP.1998.144.01.05.
of Waterloo, Waterloo, Ontario, Canada. Pinti, D.L., Béland-Otis, C., Tremblay, A., Castro, M.C., Hall, C.M., Marcil, J.-S., Lavoie, J.-Y.,
IAEA/WMO, 2004. Global network of isotopes in precipitation. The GNIP Database Lapointe, R., 2011. Fossil brines preserved in the St-Lawrence Lowlands, Quebec,
(Available at: http://isohis.iaea.org). Canada as revealed by their chemistry and noble gas isotopes. Geochim. Cosmochim.
Kulongoski, J.T., Hilton, D.R., 2011. Applications of groundwater helium. In: Baskaran, M. Acta 75, 4228–4243. http://dx.doi.org/10.1016/j.gca.2011.05.006.
(Ed.), Handbook of Environmental Isotope Geochemistry, Advances in Isotope Pinti, D.L., Gélinas, Y., Larocque, M., Barnetche, D., Retailleau, S., Moritz, A., Hélie, J.-F.,
Geochemistry. Springer-Verlag, Berlin Heidelberg, pp. 285–303. Lefebvre, R., 2013. Concentrations, sources et mécanismes de migration préférentielle
Kulongoski, J.T., Hilton, D.R., Izbicki, J.A., 2003. Helium isotope studies in the Mojave des gaz d'origine naturelle (méthane, hélium, radon) dans les eaux souterraines des
Desert, California: implications for groundwater chronology and regional seismicity. Basses- Terres du Saint-Laurent. Strategic Environmental Evaluation Committee on
Chem. Geol. 202, 95–113. http://dx.doi.org/10.1016/j.chemgeo.2003.07.002. Shale gas. Report E3-9, p. 104 (Available at: http://ees-gazdeschiste.gouv.qc.ca/
Kulongoski, J.T., Hilton, D.R., Izbicki, J.A., 2005. Source and movement of helium in the wordpress/wp-content/uploads/2013/11/Rapport-etude-E3-9_Geotop-UQAM.pdf).
eastern Morongo groundwater Basin: the influence of regional tectonics on crustal Plummer, L.N., Prestemon, E.C., Parkhurst, D.L., 1994. An interactive code (NETPATH) for
and mantle helium fluxes. Geochim. Cosmochim. Acta 69, 3857–3872. http://dx.doi. modeling NET geochemical reactions along a flow PATH, version 2.0. USGS, Water-
org/10.1016/j.gca.2005.03.001. Resources Investigations Report 94-4169.
Lamarche, L., Bondue, V., Lemelin, J.-M., Lamothe, M., Roy, M., 2007. Deciphering the Plummer, L.N., Glynn, P.D., 2013. Radiocarbon dating in groundwater systems. Isotope
Holocene evolution of the St. Lawrence River drainage system using luminescence Methods for Dating Old Groundwater. International Atomic Energy Agency, Vienna,
and radiocarbon dating. Quat. Geochronol. 2, 155–161. http://dx.doi.org/10.1016/j. pp. 33–89.
quageo.2006.04.002. Sano, Y., Tokutake, T., Takahata, N., 2008. Accurate measurement of atmospheric helium
Lamothe, M., 1989. A new framework for the Pleistocene stratigraphy of the central St. isotopes. Anal. Sci. 24, 521–525. http://dx.doi.org/10.2116/analsci.24.521.
Lawrence Lowland, southern Quebec. Géog. Phys. Quatern. 43, 119–129. http://dx. Schlosser, P., Stute, M., Sonntag, C., Munnich, K.O., 1989. Tritiogenic 3He in shallow
doi.org/10.7202/032764ar. groundwaters. Earth Planet. Sci. Lett. 94, 245–256. http://dx.doi.org/10.1016/0012-
Larocque, M., Gagné, S., Tremblay, L., Meyzonnat, G., 2013. Projet de connaissance des 821X(89)90144-1.
eaux souterraines du bassin versant de la rivière Becancour et de la MRC de Smith, S.P., Kennedy, B.M., 1983. The solubility of noble gases in water and NaCl brine.
Becancour. PACES Final Report. Quebec Ministry of Environment, p. 187 (Available Geochim. Cosmochim. Acta 47, 503–515. http://dx.doi.org/10.1016/0016-
at: http://www.grobec.org/hydrogeo/pdf/Rapport_synthese_PACES_Becancour_ 7037(83)90273-9.
2013.pdf). Solomon, D.K., Schiff, S.L., Poreda, R.J., Clarke, W.B., 1993. A validation of the 3H/3He
Lavoie, D., Rivard, C., Lefebvre, R., Séjourné, S., Thériault, R., Duchesne, M.J., Ahad, J.M.E., method for determining groundwater recharge. Water Resour. Res. 29, 2951–2962.
Wang, B., Benoit, N., Lamontagne, C., 2013. The Utica Shale and gas play in southern http://dx.doi.org/10.1029/93WR00968.
Quebec: geological and hydrogeological syntheses and methodological approaches Stute, M., Schlosser, P., 1993. Principles and applications of the noble gas
to groundwater risk evaluation. Int. J. Coal Geol. 126, 77–91. http://dx.doi.org/10. paleothermometer. In: Smart, P.K., Lohmann, K.C., McKenzie, J., Savin, S. (Eds.),
1016/j.coal.2013.10.011. Climate Change in Continental Isotopic RecordsGeophys. Monogr. 78. AGU,
Le Gal Lasalle, C., Marlin, C., Leduc, C., Taupin, J.D., Massault, M., Favreau, G., 2001. Renewal Washington. D.C., pp. 89–100.
rate estimation of groundwater based on radioactive tracers (3H, 14C) in an uncon- Takaoka, N., Mizutani, Y., 1987. Tritiogenic 3He in groundwater in Takaoka. Earth Planet.
fined aquifer in a semi-arid area, Iullemeden Basin, Niger. J. Hydrol. 254, 145–156. Sci. Lett. 85, 74–78. http://dx.doi.org/10.1016/0012-821X(87)90022-7.
http://dx.doi.org/10.1016/S0022-1694(01)00491-7. Taupin, J.D., 1990. Evaluation isotopique de l'évaporation en zone non saturée sous climat
Levin, I., Bösinger, R., Bonani, G., Francey, R., Kromer, B., Münnich, K.O., Suter, M., Trivett, sahélien et evolution géochimique des solutions des sols (vallée moyen du Niger)
N.A., Wölfli, W., 1992. Radiocarbon in atmospheric carbon dioxide distribution and (PhD thesis) Université de Paris Sud, Orsay, France.
trends. In: Taylor, R.E., Long, A., Kra, R.S. (Eds.), Radiocarbon After Two Decades. Torgersen, T., 1980. Controls on pore-fluid concentration of 4He and 222Rn and the
Springer, New York, pp. 503–518. calculation of 4He/222Rn ages. J. Geochem. Explor. 13, 57–75. http://dx.doi.org/10.
Ma, L., Castro, M.C., Hall, C.M., 2004. A late Pleistocene noble gas paleotemperature 1016/0375-6742(80)90021-7.
record in southern Michigan. Geophys. Res. Lett. 31. http://dx.doi.org/10.1029/ Torgersen, T., 2010. Continental degassing flux of 4He and its variability. Geochem.
2004GL021766. Geophys. Geosyst. 11, Q06002. http://dx.doi.org/10.1029/2009GC002930.
Ma, L., Castro, M.C., Hall, C.M., Lohmann, W.M., 2005. Cross-formational flow and salinity Torgersen, T., Clarke, W.B., 1985. Helium accumulation in groundwater, I: an evaluation of
sources inferred from a combined study of helium concentrations, isotopic ratios and sources and the continental flux of crustal 4He in the Great Artesian Basin, Australia.
major elements in the Marshall aquifer, southern Michigan. Geochem. Geophys. Geochim. Cosmochim. Acta 49, 1211–1218. http://dx.doi.org/10.1016/0016-
Geosyst. 6 (10), Q10004. http://dx.doi.org/10.1029/2005GC001010. 7037(85)90011-0.
Ma, L., Castro, M.C., Hall, C.M., 2009. Crustal noble gases in deep brines as natural tracers Tran Ngoc, T.D., Lefebvre, R., Konstantinovskaya, E., Malo, M., 2014. Characterization of
of vertical transport processes in the Michigan Basin. Geochem. Geophys. Geosyst. 10 deep saline aquifers in the Becancour area, St. Lawrence Lowlands, Quebec, Canada:
(6), Q06001. http://dx.doi.org/10.1029/2009GC002475. implications for CO2 geological storage. Environ. Earth Sci. http://dx.doi.org/10.
Marty, B., Torgersen, T., Meynier, V., O'Nions, R.K., de Marsily, G., 1993. Helium isotope 1007/s12665-013-2941-7.
fluxes and groundwater ages in the Dogger Aquifer, Paris Basin. Water Resour. Res. Warrier, R.B., Castro, M.C., Hall, C.M., 2012. Recharge and source-water insights from the
29, 1025–1035. http://dx.doi.org/10.1029/93WR00007. Galapagos Islands using noble gases and stable isotopes. Water Resour. Res. 48,
Matsuda, J., Matsumoto, T., Sumino, H., Nagao, K., Yamamoto, J., Miura, Y., Kaneoka, I., W03508. http://dx.doi.org/10.1029/2011WR010954.
Takahata, N., Sano, Y., 2002. The 3He/4He ratio of the new internal He Standard of Weise, S., Moser, H., 1987. Groundwater dating with helium isotopes. Techniques in
Japan (HESJ). Geochem. J. 36, 191–195. Water Resource Development. IAEA, Wien, pp. 105–126.
Meyzonnat, G., 2012. Estimation de la vulnérabilité de l'aquifère au roc de la zone Whiticar, M.J., 1999. Carbon and hydrogen isotope systematics of bacterial formation and
Becancour (Centre-du-Québec) (MSc thesis), Université du Quebec à Montréal, oxidation of methane. Chem. Geol. 161, 291–314. http://dx.doi.org/10.1016/S0009-
Canada (139 pp.). 2541(99)00092-3.
Meyzonnat, G., Larocque, M., Barbecot, F., Gagné, S., Pinti, D.L., 2015. The potential of
major ion chemistry to assess groundwater vulnerability of a regional aquifer in
southern Quebec (Canada). Environ. Earth Sci. (in press).

S-ar putea să vă placă și