• Aucun résultat trouvé

Regional groundwater flow dynamics and residence times in Chaudière-Appalaches, Québec, Canada : insights from numerical simulations

N/A
N/A
Protected

Academic year: 2021

Partager "Regional groundwater flow dynamics and residence times in Chaudière-Appalaches, Québec, Canada : insights from numerical simulations"

Copied!
114
0
0

Texte intégral

(1)

Regional Groundwater Flow Dynamics and Residence

Times in Chaudière-Appalaches, Québec, Canada:

Insights from Numerical Simulations

Mémoire

Debora Janos

Maitrise interuniversitaire en sciences de la Terre

Maître ès sciences (M.Sc.)

Québec, Canada

(2)

Regional Groundwater Flow Dynamics and Residence

Times in Chaudière-Appalaches, Québec, Canada:

Insights from Numerical Simulations

Mémoire

Debora Janos

Sous la direction de :

John Molson, directeur de recherche

(3)

iii Résumé

Dans le cadre du projet PACES III pour la région de Chaudière-Appalaches, situé au sud de la ville de Québec, au Canada, l'étude présente une analyse approfondie de l’influence des dynamiques d’écoulement sur la qualité des eaux souterraines dans un contexte régional. L’écoulement régional, le transport d’âge et l'impact d'une faille sur la qualité de l'eau souterraine sont étudiés par l’entremise de modèles numériques bidimensionels. La combinaison des connaissances hydrogéologiques physiques et chimiques, y compris une analyse des concentrations de 14C dans les eaux souterraines échantillonnées, a conduit à l’ébauche d'un modèle conceptuel de l’écoulement régional. Ce dernier est mis à l’essais pas l’entremise d’un modèle d'écoulement numérique suivant une ligne d’écoulement régionale dans le plan 2D vertical à l’aide du logiciel FLONET. Le modèle est d'abord calibré à l’aide d’une méthode semi-automatisé qui utilise le logiciel PEST en comparant les charges simulés à la piézométrie régionale, et est validé par la comparaison des flux simulés à la recharge. Bien que le modèle affiche l’existence d’un écoulement régional profond, la région à l’étude apparaît être dominée par des systèmes d'écoulements locaux à des échelles maximales d'environ 5 km, avec un écoulement significatif dans le roc fracturé peu profond. L’écoulement actif se limitant à une profondeur maximale de 40 m à 60 m du roc fracturé, confirme que la géochimie des eaux souterraines échantillonnées à partir de puits résidentiels est susceptible d'être affectée par les eaux faisant parti de l’écoulement intermédiaire et régional. Le transport advectif-dispersif de l'âge est ensuite simulé avec le simulateur de transport TR2 et comparé aux temps de déplacement advectifs le long des lignes d’écoulements et à l'âge 14C des eaux échantillonnées. Enfin, l’influence de la faille de la Rivière Jacques Cartier sur le contexte hydrogéologique régional est étudiée à travers divers scénarios hypothétiques de perméabilité de faille.

(4)

iv Abstract

As part of the PACES III project in the Chaudière-Appalaches region, south of Quebec City, Quebec, Canada, the study herein presents insights into the extent to which regional groundwater quality is shaped by flow dynamics. In this context, 2D numerical modelling is used to simulate regional flow, transport of groundwater age and the possible influence of a fault on groundwater quality. Combining physical and chemical hydrogeological knowledge, including an analysis of 14C concentrations in sampled groundwater, leads to the development of a regional conceptual flow model. The conceptual model is tested by representing the system with a two-dimensional numerical flow model oriented in the vertical plane roughly south-north towards the St. Lawrence River using the FLONET code. The model is first calibrated to regional piezometry through a semi-automated workflow using PEST and is then validated with average recharge values. Although some evidence for deeper regional flow exists, the area appears to be dominated by local flow systems on maximum length scales of about 5 km, with significant flow through the shallow fractured sedimentary rock aquifer. This regional scale flow model is also supported by the local hydrogeochemical signatures. Active flow appears contained within the top 40 m to 60 m of the fractured bedrock, which confirms that the geochemical signatures of groundwater sampled from residential wells are likely affected by the slow moving waters of the intermediate and regional flow systems. Advective-dispersive transport of groundwater age is then simulated with the TR2 transport model and compared with advective travel times and sampled 14C water ages. Finally, the possible role of the Jacques-Cartier River fault on regional flow dynamics is investigated by testing various fault permeability configurations.

(5)

v Table of contents Résumé ... iii Abstract ... iv Table of contents ... v List of tables ... vi

List of figures ... vii

Remerciements ... xi

1. Introduction ... 1

1.1 Objectives of the Study ... 3

1.2 Methodology ... 4

2. Study area ... 5

2.1 Geologic and geographic context ... 5

2.2 Regional Piezometry ... 8

2.3 Geochemical Setting ... 11

2.3.1 Major Ion Chemistry ... 11

2.3.2 Evidence of Geochemical Processes ... 13

2.3.3 Radiocarbon ages ... 23

2.4 Regional Scale Conceptual Model ... 34

3. Numerical Modelling ... 36

3.1 Modelling Strategy ... 37

3.2 Domain description ... 38

3.3 Steady-State Regional Flow Model ... 38

3.3.1 Assumptions and limitations ... 38

3.3.2 Model Parameters ... 39

3.3.3 Calibration ... 46

3.3.3 Flow Model Results ... 50

3.3.4 Sensitivity Analysis ... 55

3.4 Groundwater Age Simulations ... 56

3.4.1 Advective-Dispersive Groundwater Age ... 57

3.4.2 Comparison between Advective and Advective-Dispersive Groundwater Ages. 63 3.5 Parametric study on faults ... 67

3.5.1 Modeled fault configurations ... 68

3.5.2 Results: groundwater flow and age ... 70

4. Discussion ... 73

5. Conclusion ... 76

6. References ... 78

Annex A: Geochemical data and 14C age results ... 89

Annex B: Sensitivity Analysis ... 92

(6)

vi List of tables

Table 1 Conventional adjustment models accounting for geochemical reactions affecting the initial activities used for the determination of groundwater sample ages, modified from IAEA (2013) (Han and Plummer 2016; Clark and Fritz 1997) ... 26 Table 2 Distribution of radiocarbon groundwater ages estimated using Han and Plummer's (2016) method ... 29 Table 3 Physical description of the surficial deposits, including reported and calibrated hydraulic conductivities. ... 41 Table 4 Description of the modeled geologic units, including the reported and modeled hydraulic conductivities and porosities (includes properties reported by Gassiat et al. 2013). ... 44 Table 5 Average recharge to the bedrock calibrated from the HELP model and from the current 2-D flow model (SLL = St. Lawrence Lowlands, AHL=Appalachian Highlands). 48 Table 6 Description of sensitivity analysis scenarios. Each case refers to changes with respect to the base case (calibrated) model. ... 55 Table 7 Time steps and stability criteria for the age transport model ... 58 Table 8 Simulated fault permeability scenarios ... 70

Table A 1Radiocarbon ages of samples collected in the Chaudière-Appalaches region, calculated using various correction methods ... 90 Table A 2 Radiocarbon ages of samples collected in the Chaudière-Appalaches calculated using various correction methods ... 91

(7)

vii List of figures

Figure 1 Location and topography of the Chaudière-Appalaches study area, and location of

2D regional cross-Section model, modified from Lefebvre et al. (2015)... 2

Figure 2 Regional geology of the Chaudière-Appalaches region, modified from Lefebvre et al. (2015) ... 5

Figure 3 Geologic north-south cross-Section (see Figure 2 for location), geology based on Séjourné et al. (2013)... 7

Figure 4 Water table elevation with respect to average sea level (m), modified from Lefebvre et al. (2015) ... 9

Figure 5 North-south regional cross-Section of the study area showing conceptual model of regional flow ... 10

Figure 6 Piper diagram of natural water types found in the study area ... 12

Figure 7 Evidence of carbonate dissolution in the C-A study area: a) Calcite and dolomite equilibrium with respect to pH and partial pressure of CO2; b) Ca2+ and Mg2+ ratios of dissolving carbonates, c) Saturation of calcite with respect to partial pressure of CO2, and d) Saturation of calcite with respect to dolomite. ... 16

Figure 8 Expected ratios between the concentrations of DIC and Ca2+ ions for minerals dissolving in various proportions: a) DIC with respect to Ca2+ ions b) DIC with respect to measured Ca2+ ions and exchanged Ca2+ ions, assuming all Na+ ions are from Ca2+/Na+ ion exchange ... 18

Figure 9 Na/Cl ratios indicating mixing and ion exchange ... 19

Figure 10 Comparison of Na+ ions exchanged using Cl- as a conservative tracer with the excess Na+ obtained from the dissolution of calcite: a) Na+ ions exchanged calculated from calcite dissolution with respect to Na+ ions exchanged calculated from surplus of seawater Na-Cl ratio, b) Na+ ions surplus after ion exchange with respect to Cl- concentration in samples ... 20

Figure 11 Br- and Cl- concentrations ... 21

Figure 12 Cl- and Li+ concentrations ... 21

(8)

viii

Figure 14 Carbon isotope concentrations of samples collected in the Chaudière-Appalaches region plotted on Han and Plummer's (2016) graphical interpretation of the validity of correction methods ... 28 Figure 15 Distribution of radiocarbon groundwater ages estimated using Han and

Plummer's (2016) method. Whiskers represent maximum and minimum values, and boxes represent percent quartiles. ... 29 Figure 16 Trends between δ13C, 14C and DIC concentrations (Han et al. 2012). Arrows identify geochemical process affecting the samples. Colors identify samples that have undergone similar reactions. ... 32 Figure 17 Trends between δ13C, 14C and DIC concentration (Han et al. 2012). Arrows identify geochemical process affecting the samples. Colors identify samples that have undergone similar reactions. ... 33 Figure 18 Conceptual regional groundwater flow and geochemical evolution model for the region of Chaudière-Appalaches ... 35 Figure 19 Distribution of hydraulic conductivities in Chaudière-Appalaches region. Grey dots represent measured data from the well drillers’ log (Lefebvre et al. 2015). Black squares show the median measured value per 5 meter slices and the horizontal whiskers show the 25th and 75th percentiles. Solid lines show the calibrated distribution of hydraulic conductivities in the fractured bedrock and dotted lines show different conductivity

scenarios tested in the sensitivity analysis. ... 43 Figure 20 Geological structure, hydraulic conductivity and boundary conditions for the numerical flow model ... 45 Figure 21 Model calibration: Observed heads (from interpolated regional piezometry) against simulated heads ... 47 Figure 22 Calibrated recharge fluxes, simulated hydraulic heads, observed water table elevations (from interpolated piezometry) and topographic elevation along the model cross-Section. Negative fluxes are exiting the model domain and positive fluxes are entering .... 49 Figure 23 Calibrated steady-state flow model showing hydraulic heads and streamline distribution ... 51 Figure 24 Calibrated flow model Zoom 1 (Appalachian Highlands): hydraulic heads and streamline distribution (location shown in Fig. 23). ... 51

(9)

ix

Figure 26 Calibrated flow model Zoom 2 (St. Lawrence Lowlands close to St. Lawrence River): hydraulic heads and streamline distribution (location shown in Fig. 23). ... 52 Figure 26 Calibrated flow model Zoom 3 (St. Lawrence Lowlands close to Appalachian front): hydraulic heads and streamline distribution (location shown in Fig. 23). ... 52 Figure 27 Calibrated flow model Zoom 4 (Appalachian Highlands): hydraulic heads and streamline distribution (location shown in Fig. 23). ... 53 Figure 28 Average vertical Darcy fluxes as a function of depth from the surface of bedrock ... 54 Figure 29 Simulated mean groundwater ages with superimposed streamlines from the calibrated flow model Zoom 1 (Appalachian Highlands): mean age and streamlines

(location shown in Fig. 30). ... 60 Figure 30 Simulated mean groundwater ages with superimposed streamlines from the calibrated flow model. ... 60 Figure 31 Simulated mean groundwater ages with superimposed streamlines from the calibrated flow model Zoom 3 (St. Lawrence Lowlands close to Appalachian front): mean age and streamlines (location shown in Fig. 30). ... 61 Figure 32 Simulated mean groundwater ages with superimposed streamlines from the calibrated flow model Zoom 2 (St. Lawrence Lowlands close to St. Lawrence River): mean age and streamlines (location shown in Fig. 30). ... 61 Figure 33 Simulated mean groundwater ages with superimposed streamlines from the calibrated flow model Zoom 4 (Appalachian Highlands): mean age and streamlines

(location shown in Fig. 30). ... 62 Figure 34 Simulated mean groundwater ages with superimposed streamlines from the calibrated flow model Zoom 5 (middle of St. Lawrence Lowlands): mean groundwater age and streamlines ... 63 Figure 35 Calibrated flow model: mean age distribution and advective particle tracks. Times at top discharge points refer to final advective residence times between recharge and discharge points along selected streamlines. ... 65 Figure 36 Calibrated flow model Zoom 1 (Appalachian Highlands): mean age distribution and advective particle tracks. Times at top discharge points refer to final advective

(10)

x

Figure 37 Calibrated flow model Zoom 5 (middle of St. Lawrence Lowlands): mean age distribution and advective particle tracks. Times at top discharge points refer to final advective residence times between recharge and discharge points along selected

streamlines. ... 66 Figure 38 Calibrated flow model Zoom 2 (St. Lawrence Lowlands close to St. Lawrence River): mean age distribution and particle tracks (location shown in Fig. 30). Times at top discharge points refer to final advective residence times between recharge and discharge points along selected streamlines. ... 66 Figure 39 Jacques-Cartier River fault zone conceptual model. Vertical hydraulic

conductivities shown correspond to scenario P of the parametric study on faults. ... 69

Figure B 1 Average vertical Darcy fluxes as a function of depth from the surface of bedrock for sensitivity analysis scenarios described in Table 6 ... 93 Figure B 2 Steady-state flow model showing hydraulic heads and streamline distribution for sensitivity analysis scenarios described in Table 6 ... 94 Figure B 3 Flow model Zoom 2 (St. Lawrence Lowlands close to St. Lawrence River): hydraulic heads and streamline distribution (location shown in Fig. 23) for sensitivity analysis scenarios described in Table 6 ... 95 Figure B 4 Steady-state flow and transport model showing mean age for sensitivity analysis scenarios described in Table 6 ... 96 Figure B 5 Flow and transport model Zoom 2 (St. Lawrence Lowlands close to St.

Lawrence River): mean age (location shown in Fig. 23) for sensitivity analysis scenarios described in Table 6 ... 97

Figure C 1Calibrated steady-state flow model showing hydraulic heads and streamline distribution for the base case model and the location of the modeled fault (Zoom 1) ... 99 Figure C 2 Zoom1: Steady-state flow model showing hydraulic conductivities in the

horizontal direction and streamlines for fault scenarios described in Table 8 ... 100 Figure C 3 Zoom 1: Steady-state flow and age transport model showing mean groundwater and streamlines for fault scenarios described in Table 8, vertical exaggeration 1.25x ... 101 Figure C 4 Zoom 2: Steady-state flow and age transport model showing mean groundwater and streamlines for fault scenarios described in Table 8, vertical exaggeration 4x ... 102

(11)

xi Remerciements

Dans un premier temps, je tiens à remercier mon directeur de recherche, John Molson, pour m’avoir donné sa confiance et m’avoir guidé dans la réalisation de ce projet. Toujours disponible et patient, son encadrement et ses précieux conseils ont enrichis mes travaux et m’ont permis de développer une passion pour et un regard critique sur la modélisation numérique. Grâce à son appuie et ses encouragements, j’également eu la chance de partager mes travaux avec mes collègues hydrogéologues à l’international et je lui en suis extrêmement reconnaissante.

J’aimerais également remercier mon co-directeur de recherche, René Lefebvre, pour tous ses nombreux conseils qui ont étés phares dans l’orientation de mon projet, sa disponibilité pour m’aider à démystifier la géochimie et son enthousiasme contagieux envers mon projet. Je voudrais aussi souligner l’apport l’équipe PACES de l’INRS, notamment Harold Vigneault, Jean-Marc Ballard, Marc-André Carrier, Marc Laurencelle, Xavier Malet, Annie Therrien, Châtelaine Beaudry et Guillaume Légaré Couture; les données de terrains qu’ils ont compilés, leur contributions techniques et leur précieux conseils sont les fondements de mes travaux. Je remercie aussi le Ministère du Développement durable, de l’Environnement et de la Lutte contre les changements climatiques pour sa contribution financière et sans qui ce projet n’aurait pas pu avoir lieu.

Également j’aimerais remercier Christine Rivard, Geneviève Bordeleau et Pierre Ladevèze de la Commission Géologique du Canada pour avoir partagé leur travaux de recherche avec moi et pour avoir grandement contribué à l’avancement de mes travaux.

Finalement, je remercie tous les étudiants que j’ai côtoyés durant ces années de recherches et sans qui mon expérience à l’université Laval n’aurait pas été la même.

(12)

1 1. Introduction

Understanding relationships between local, intermediate and regional groundwater flow systems is critically important for management of water resources as well as to reduce risk of contamination from deep energy resource development (ex. shale gas or geothermal energy). In this context, in 2009 the Ministry of Environment in Quebec, Canada (Ministère du Développement durable, de l’Environnement et de la Lutte contre les changements climatiques, MDDELCC) launched a systematic groundwater resources assessment program (Programme d’acquisition des connaissances sur les eaux souterraines, PACES) in areas of southern Quebec where one-fifth of the population of 8 million (ISQ 2017) relies on groundwater as their main water source (Rousseau et al. 2004; MDDELCC 2017). By 2015, the PACES program had produced exhaustive regional portraits on the state of groundwater resources in 9 multi-watershed regions. The present study was part of one of these PACES projects and focused on regional groundwater flow dynamics and natural hydro-geochemistry and geochemical processes in the Chaudière-Appalaches region, south of Quebec City (Figure 1) (referred to here as PACES-CA).

The Chaudière-Appalaches study area spans over 14,625 km2 and has a population of approximately 276,000, mainly concentrated near major waterways including the St. Lawrence River, and the Chaudière River and its tributaries (Lefebvre et al. 2015). Although surface water is abundant, almost half of the water usage (42%) in the CA region is extracted from the subsurface, a third of which is used for agriculture (Lefebvre et al. 2015). Agriculture dominates over about 26.7% of the territory and plays a central role in the region’s economy, thus representing a major stakeholder in sustainable groundwater management.

(13)

2

As groundwater is fundamental to the local economy, groundwater resources within the CA region, both in terms of quantity and quality, have been the focus of a number of studies over the past three decades, culminating in the current PACES project. These projects have helped to characterize the resource, focusing on the most densely populated Chaudière River watershed (McCormack 1982; Rousseau et al. 2004; Benoît et al. 2009; Benoît et al. 2011; Brun Koné 2013; Benoît et al. 2014; Carrier et al. 2014; Lefebvre et al. 2015). Based primarily on data collected from shallow residential wells (with average depths of approximately 60m), these studies have provided important information on groundwater quality and aquifer properties, concluding that groundwater flow mainly occurs within

Figure 1 Location and topography of the Chaudière-Appalaches study area, and location of 2D regional cross-Section model, modified from Lefebvre et al. (2015)

(14)

3

permeable valley sediments and in the top fractured part of the bedrock (< 30m) (Brun Koné 2013; Benoît et al. 2014; Crow and Ladevèze 2015; Ladevèze et al., 2016; Ladevèze et al. 2017; Ladevèze 2017). Furthermore, Benoît et al. (2014) and Lefebvre et al. (2015) also highlighted geochemical signatures left behind by the invasion of the Champlain Sea, which flooded the northern areas of the Chaudière-Appalaches region (Figure 1) between about 12,000 to 10,000 years ago (Parent and Occhietti 1988), thereby showing that geochemical signatures of this event can still be seen today.

In parallel, the potential exploitation of unconventional gas resources from the Utica Shale of the St. Lawrence Platform geological province, which includes parts of the Chaudière-Appalaches region, led to a series of additional geological and hydrogeological studies led by the Geological Survey of Canada and INRS (Institut national de la recherche scientifique). These studies aimed to assess the risks of groundwater contamination associated with shale gas exploitation through upward fluid migration and to establish the baseline hydrogeological conditions prior to exploitation. Focusing their case study on the region of Saint-Édouard-de-Lotbinière (Figure 1), Bordeleau et al. (2015) and Rivard et al. (2014) found, despite local flow regimes being more apparent, indications of relatively deep regional flow emerging near the Jacques-Cartier River fault line (Figure 2). Their findings support the conceptualization of regional groundwater flow as a Tothian flow system (Tóth 1999; Tóth 1963; Tóth 2009), in which local flow regimes are embedded within larger intermediate or regional flow regimes. As shale gas exploration areas border the St. Lawrence River (Lavoie et al. 2014; Moritz et al. 2015; Rivard et al. 2014), which is a clear outlet for regional groundwater flow, broadening the understanding of how such systems might interact with local flow systems becomes even more important.

1.1 Objectives of the Study

The aim of this study is to reinforce understanding of the regional flow system in the Chaudière-Appalaches region. More specifically, the study first aims to investigate the characteristic time scales of groundwater flow processes through a review of the regional geochemistry and from estimates of groundwater age based on isotopic evidence. A conceptual model of the regional hydrogeological system is then tested using a two-dimensional numerical flow and age transport model. The numerical model aims to

(15)

4

quantify the magnitude of regional flow, estimate the maximum depth of active groundwater flow, investigate the distribution of groundwater residence times, perform a sensitivity analysis of the influence of normal faults on regional flow and examine the contribution of formation brines and the Champlain Sea to regional hydrogeochemical signatures.

1.2 Methodology

The study objectives are met by first carrying out a thorough review of the regional geological context and Quaternary history along with a synthesis of the state of knowledge of the regional geochemistry and an analysis of groundwater ages. Combining the physical and chemical hydrogeological knowledge will then lead to the development of a regional conceptual flow model.

The conceptual model is then tested by simulating the regional flow system within a two-dimensional vertical plane, about 60 km long, oriented roughly southeast-northwest from a regional topographic flow divide towards the St. Lawrence River. All flow simulations are run using the FLONET/TR2 numerical code (Molson & Frind, 2017), assuming steady-state, saturated conditions. The model is first calibrated using the PEST inverse calibration model (version 13.6; Doherty 2015) and a sensitivity analysis is performed on hydraulic conductivities. The role of faults is then investigated by testing various fault configurations based on the findings of a hydromechanical characterisation of the Saint-Édouard-de-Lotbinière area conducted by Ladevèze (2017) and Bordeleaux et al. (2017). Finally, the calibrated flow model and the fault scenarios are applied in the TR2 transport model to simulate the distribution of groundwater ages.

(16)

5 2. Study area

2.1 Geologic and geographic context

Regional landscape and topography is characterized by a combination of various geological settings, dominated by the geological provinces of the Appalachians and the St. Lawrence Platform (Figure 2). Although both geologic provinces are primarily sedimentary in origin, the Appalachians are highly deformed and slightly metamorphosed whereas the St. Lawrence Platform has undergone almost no deformation.

Figure 2 Regional geology of the Chaudière-Appalaches region, modified from Lefebvre et al. (2015)

(17)

6

Situated to the south of the study, the Appalachians are the remains of a weathered mountain range and are characterised by a hilly terrain. At the southernmost limit of the study area, the Gaspé Belt, a sedimentary basin which formed following the Taconian orogeny, sits on an oceanic sedimentary and volcanic basin called the Dunnage zone (St-Julien & Hubert 1975). The latter is followed to the north by the Humber zone which formed within a passive continental margin and is composed of carbonates, sandstones and shales (Lavoie 2002). Furthermore, the Appalachian region is marked by granitic intrusions forming small mountains, including the Petits Monts Mégantic and the Monts Notre Dame mountains (Figure 1) reaching elevations up to 850 and 1000 meters, respectively (Slivitzky & St-Julien 1987). The intrusions are surrounded by a belt of small hills, which are bordered to the south by the Appalachian plateau and to the north by a foothill area that marks the junction with the sedimentary plains of the St. Lawrence Lowlands (Slivitzky & St-Julien 1987). With a history of high tectonic activity, the Appalachian region is marked by the presence of lineaments oriented NE-SW which manifest primarily in the form of major faults (Lefebvre et al. 2015; Caron 2012).

In the northernmost part of the study area, the St. Lawrence Platform consists of a sedimentary succession sitting unconformably on the base of the Canadian Shield (Globensky 1987). Due to tectonic compression, normal faults mark the gradual lowering of the sedimentary sequence, which plunges below the Appalachian Basin (Castonguay et al. 2010) as shown in Figure 3.

The Cambrian-Ordovician sedimentary sequence of the St. Lawrence Platform is a typical marine transgression-regression sequence resulting from the opening and closing of the Iapetus Ocean. From bottom up, the sequence begins with the Potsdam Group, a poorly cemented sandstone formation becoming well cemented in its upper portion (Cloutier et al. 2006). Above this formation sits the commonly named carbonate platform which is separated into four groups. First, the Beekmantown is composed of dolomitic sandstone layers becoming pure dolostone and, near the top, dolomitic limestone. The Beekmantown is followed by the Chazy goup, a clayey and sandy limestone, the Black River Group, a limestone interbedded with sandstone, and finally the Trenton group, a clayey limestone interbedded with shales (Séjourné et al. 2013; Cloutier et al. 2006). The increase in the ratio

(18)

7

of shale interbeds to limestone layers marks a gradual transition towards the Utica Shale. At depths ranging from 800 to 3000 meters, the Utica Shale, which can be considered as marl due to its high carbonate content, is particularly enriched in natural gas and is of economic interest in the region (Lavoie et al. 2014; Bordeleau et al. 2015; Thériault 2012). The overlying Lorraine group, composed of silty shales with mostly non-calcareous sediments interbedded with fine sandstone and clayey siltstones (Thériault 2012) is the top formation of the St. Lawrence Platform sedimentary succession over most of the study area. In this study, St. Lawrence Platform refers to the geologic sequence composed of the Carbonate Platform, the Utica Shale and the Lorraine Formation while the St. Lawrence Lowlands refers to the relatively flat physiographic region which was formerly covered by the Champlain Sea (Figure 3).

Many past studies have identified the regional aquifer as being located in the uppermost portion of the bedrock where fracture density is highest (Lefebvre et al. 2015; Benoît et al. 2014; Ladevèze et al., 2016). Major faults could also represent areas where the level of fracturing could be relatively high, although the role of these faults on the regional flow system is not clear.

As the study area is generally covered with till deposits of various thicknesses and assemblies, the nature of the till units defines the degree of confinement of the bedrock aquifer. In the Appalachians, surficial sediments are thin or non-existent, therefore unconfined conditions prevail for most of the fractured bedrock aquifer, with the exception

Figure 3 Geologic north-south cross-Section (see Figure 2 for location), geology based on Séjourné et al. (2013)

(19)

8

of valleys where the aquifer is generally semi-confined (Lefebvre et al. 2015). Moreover, glaciation-deglaciation processes, the presence of the Champlain Sea approximately 10,000 years ago, and the formation of post-glacial lakes and shorelines, led to heterogeneous deposits with varying proportions of sand and silt. These sediments together with alluvial deposits represent significant infiltration pathways when directly in contact with the bedrock aquifer. On the other hand, scattered silt and clay sediments, deposited at the bottom of stagnant water bodies following glacial retreat, prevent recharge and are responsible for the accumulation of organic material (Lefebvre et al. 2015). These low-permeability deposits are present mostly within the area formerly covered by the Champlain Sea, specifically in the northwestern area of the St. Lawrence Lowlands (Figure 1). This region is further characterized by significant peatland coverage. Many wetlands in the area were formed over sediment deposits having strong aquifer potential and thus play a role in buffering the effect of seasonal recharge on groundwater levels (Bourgault et al. 2017).

2.2 Regional Piezometry

In terms of regional groundwater flow a semi-regional groundwater divide line can be drawn across the Appalachian Highlands (Figure 4). Because of its prominent elevation as well as its limited or negligible surficial sediment cover, the Appalachian region can be considered a recharge area while the St. Lawrence River, marking the northern boundary and the lowest point of the study area, represents the regional flow outlet (Figures 4 and 5). Streams and rivers within the numerous sub-basins also act as local discharge areas, particularly since the highly fragmented terrain of the Appalachians forms a dense hydraulic network (Figure 4). The extent and contribution of regional flow to the basin-scale water budget is not yet well defined. However, the conceptual groundwater flow system of the Chaudière-Appalaches region appears to fit Tóth’s (1963) concept of nested groundwater flow, in which local flow regimes are embedded within larger intermediate or regional flow regimes (Figure 5).

(20)

9

Figure 4 Water table elevation with respect to average sea level (m), modified from Lefebvre et al. (2015)

(21)

10

Figure 5 North-south regional cross-Section of the study area showing conceptual model of regional flow

Total precipitation over the study area is estimated to be 1149 mm/year, with an average of 166 mm/year of recharge reaching the fractured rock aquifer (Lefebvre et al. 2015). Spatial distribution of recharge is, however, highly variable. In the St. Lawrence Lowlands, areas covered with thick layers of fine sediments can receive less than 15 mm/year whereas rock outcrop areas can receive more than 300 mm/year. In the Appalachian Highlands, where the surficial sediments are fairly thin and permeable, aquifer recharge tends to be higher, varying from 100 mm/year to more than 300 mm/year. Additional studies on specific watersheds within the study area have reported overall recharge rates to the bedrock aquifer of between 27 mm/year and 38 mm/year (Benoît et al. 2014 and L’Heureux 2005, respectively). Discrepancies between recharge values could be attributed to the spatial variability of the bedrock recharge.

(22)

11 2.3 Geochemical Setting

2.3.1 Major Ion Chemistry

The chemical composition of a groundwater sample holds a remarkable amount of information on the path taken by groundwater (Appelo & Postma 2004; Clark and Fritz 1997; Hounslow 1995). As recharge water enters the subsurface and flows downgradient, its interactions with the minerals, gases and, to some extent, any living organisms along its path, will change the water’s chemical composition (Tóth 1999; Cloutier et al. 2008). Benoît et al. (2014) and Lefebvre et al. (2015) carried out a thorough analysis of the regional groundwater geochemistry for the Chaudière River watershed and the Chaudière-Appalaches region, respectively. The first 155 groundwater samples analysed by Benoît et al. (2014) were included in the subsequent geochemical analysis performed by Lefebvre et al. (2015) thus only the major ion water groups defined in the second study are discussed in this Section. Lefebvre et al. (2015) collected an additional 200 samples and classified a total of 355 water samples, collected mostly from residential wells (approximately 60 m deep) across the study area, into seven geochemical composition groups based on the relative proportions of their major ions. From the seven groups, three were considered as resulting from anthropogenic contamination and four were retained as natural water types, one of which is present only in a small part of the study area and thus was not included in the regional geochemical analysis (Figure 6).

The three main natural water types identified in the study area are classified as follows:

Group 1 (G1): Water of type Ca-HCO3, generally associated with young recharge waters. Group 2 (G2): Water of type Na-HCO3, typically evolved from Ca-HCO3 waters which have undergone ion exchange in the subsurface in which Ca2+ ions were exchanged for Na+ ions.

Group 3 (G3): Water of type Na-HCO3/Na-Cl, considered as highly evolved. High Na+ and low Ca2+ concentrations result from ion exchange and an elevated Cl- concentration is likely to be associated with the dissolution of salts. The origins of the chloride ions are discussed in later Sections.

(23)

12

For reference and comparison, four additional water types related to the study site were added to the Piper diagram: 1) the composition of a typical seawater sample (Hem 1985), 2) the composition of Sample S77 considered as Champlain Sea seawater (Cloutier et al. 2012), 3) a sample from the deepest observation well over the study area (Bordeleau et al. 2015), as well as 4) the composition of brines collected in the Cambrian-Ordovician sedimentary formation lying below the Utica Shale in the Bécancour area located to the west of the study area (Pinti et al. 2011).

The spatial distribution of the major ion water groups is consistent with the conceptual flow model discussed earlier (Figure 5). Water from group G1, associated with recharge water, can be found across the study area but is primarily found in the Appalachian Highlands, a region presumably characterized by local flow systems (supported by the geochemistry) which would imply rapid transit times. On the other hand, the dominant water type across Figure 6 Piper diagram of natural water types found in the study area

(24)

13

the St. Lawrence Lowlands is water from the more evolved group G3, implying that these samples would have had time to undergo geochemical changes. This suggests that the sampled groundwater in the St. Lawrence Lowlands has likely traveled over a longer distance and is relatively older. Water in group G2, which has undergone some level of ion exchange, is also mostly found in the St. Lawrence Lowlands. Travel times for this water group are expected to be longer than those of group 1 and shorter than those of G3.

2.3.2 Evidence of Geochemical Processes

Major ion analysis of waters across the study area identified three dominant geochemical processes in the region: carbonate dissolution, ion exchange and mixing. Moreover, the proposed conceptual groundwater flow model suggests a potential input of deep fossil brines to near-surface waters, supporting the hypothesis of mixing of fresh waters and brines.

2.3.2.1 Dissolution of carbonates

Mineral dissolution is controlled by the availability of the dissolving minerals (reactants) and the solubility of the mineral (reaction rate) (Freeze and Cherry, 1979). Since the most soluble minerals, such as halite, can easily be leached away, areas with rapid renewal of water, such as recharge areas, are generally left with the slower dissolving minerals such as carbonates (Freeze and Cherry, 1979). For the present study, carbonate minerals are of particular interest since they not only have a low solubility but they are also present in the study area as 1) the main building blocks of the sedimentary St. Lawrence Platform, 2) components of the till overburden resulting from the movement of glaciers from the St. Lawrence Lowlands towards the Appalachians and 3) to a certain extent, present as cementing mineral in the rocks and fractures of the Appalachian rock formations.

As noted by Freeze and Cherry (1979), only a small or insignificant amount of carbonate minerals can be sufficient to create saturated conditions and influence groundwater chemistry. Thus, across our study area, carbonate dissolution is inevitably taking place early along the flow path; during and soon after recharge. This is indicated by the presence

(25)

14 of G1 water, with major ions Ca2+ and HCO

3, signifying that dissolution of carbonate minerals has occurred through either the surficial sediments or shallow bedrock.

The equilibrium dissolution-precipitation reactions of the two most common carbonate minerals, calcite and dolomite, are given as follows:

Calcite: CaCO3  Ca2+ + CO32- (1)

Dolomite: CaMg(CO3)2  Ca2+ + Mg2+ + 2CO32- (2)

When water enters the subsurface it first migrates through the unsaturated soil zone, where groundwater equilibrates with carbon dioxide (given off through root respiration and the decay of organic matter) to pCO2 levels that are higher than that of the atmosphere (Freeze and Cherry, 1979; Appelo and Postma 2005). Appelo and Postma (2005) point to pCO2 levels in the active soil zone ranging from the atmospheric value of 10-3.5 atm to as high as 10-1.5 atm. Dissolution of gaseous CO2 into water produces excess H+ ions, which at a pH of less than 8.3, bind to CO32- ions to produce HCO3- which drives the dissolution of calcite and dolomite (Freeze and Cherry, 1979). The equilibrium dissolution-precipitation reactions then become:

Calcite: CO2(g) + H2O + CaCO3  Ca2+ + 2HCO3- (3) Dolomite: 2CO2(g) + 2H2O + CaMg(CO3)2  Ca2+ + Mg2+ + 4HCO3- (4) Water is considered to be in an open system when it is in equilibrium with “soil zone” CO2. In this case, pCO2 stays constant while carbonate minerals dissolve until equilibrium is reached. However, if the system is closed and therefore not receiving CO2 input from the soil zone, pCO2 will decrease until equilibrium is reached.

In Figure 7.A the carbon, calcite and dolomite equilibrium lines represent a pure water sample at an average temperature of 9.6 oC equilibrated using PHREEQC version 3 (Parkhurst and Appelo 1999) with CO2(g), CO2(g) and calcite, and CO2(g) and dolomite, respectively. Arrows in Figure 7 show the dissolution path in open and closed systems. The distribution shows that most samples from G1 have not yet reached calcite and dolomite equilibrium while samples from G2 and G3 appear to be saturated or supersaturated with respect to dolomite.

(26)

15

The final control on the dissolution of carbonate minerals is not their availability but their solubility which, for calcite and dolomite, is inversely dependant on temperature; dolomite becoming progressively more soluble than calcite with decreasing temperature (Szramek et al. 2007). At a temperature of approximately 10 oC, because of the relative solubility between the minerals, calcite might reach equilibrium and become supersaturated while dolomite remains undersaturated (Szramek et al. 2007). Furthermore, Freeze and Cherry (1979) state that in North America, supersaturation can result from the temperature gradient in the soil during spring. During spring recharge, dissolution occurs at low temperatures under open conditions. But, at a few meters of depth, temperatures can be several degrees higher than at the surface, leaving the water supersaturated in calcite and dolomite. The temperature difference between the recharge zone and the aquifer could explain the calcite supersaturation of water samples from G2 and G3 observed in Figures 7.A, 7.C and 7.D. Figures 7.A and 7.C also confirm that by having pCO2 levels lower than atmospheric levels, all samples from G3 were likely collected from closed system conditions in which the carbon dioxide initially present in the system was consumed by the dissolution of carbonates until equilibrium. Conversely, samples from G1 and G2 have pCO2 levels corresponding to that of the soil zone, thus indicating that these samples either reached equilibrium under open conditions or that pCO2 in the soil zone was initially high.

(27)

16

Figure 7 Evidence of carbonate dissolution in the C-A study area: a) Calcite and dolomite equilibrium with respect to pH and partial pressure of CO2; b) Ca2+ and Mg2+ ratios of dissolving carbonates, c) Saturation of calcite with respect to partial pressure of CO2, and d) Saturation of calcite with respect to dolomite.

Freeze and Cherry (1979) state that despite laboratory experiments showing that calcite equilibrium can be reached within hours or days, field rates are often slower. Nevertheless, carbonate dissolution reactions can be considered as being fairly rapid over a regional-scale flow system. Thus, samples that have not reached calcite and dolomite equilibrium, mainly those from G1, can be considered as being recently recharged water. Furthermore, high pCO2 levels can also indicate that the aquifer has a shallow circulation zone keeping the system in open conditions or that the flow distance between recharge and discharge zones is short. The large range of pCO2 and pH across the samples testifies to the variability of recharge conditions and state of evolution among the water samples.

(28)

17

In this context, the low final pCO2 levels of waters from G3 relative to G1 could indicate that G3 waters were recharged under lower initial soil pCO2 levels.

For a system in which the only carbon sources are dissolved carbon dioxide and carbonate minerals, stoichiometry can be used to derive the expected ratios between the concentrations of dissolved inorganic carbon (DIC) and Ca2+ ions for minerals dissolving in various proportions (Figure 8). The data show that water from group G1 has DIC/Ca2+ ratios corresponding to the dissolution of calcite and dolomite, while samples from G2 and G3 are affected by a process removing Ca2+ ions from the system (Figure 9a).

2.3.2.2 Cation exchange

Cloutier et al. (2010), Lefebvre et al. (2015) and Benoît et al. (2014) identified Ca2+/Na+ ion exchange as another important geochemical process in the study area. Ion exchange is described as the process by which one ion is replaced by another ion at the surface of a solid (Appelo and Postma 2005). In Chaudière Appalaches, this reaction can be described through the following equation:

½ Ca2+ + Na+-X  ½ Ca2+-X + Na+ (5)

In this case, when considering the Na+ ions as having been exchanged for Ca2+ (Figure 8B), water samples from groups G2 and G3 appear to be strongly affected by the cation exchange process (Figures 9A and 8B).

As a side note, it is worth noting that when accounting for ion exchange, the ratio of dissolved dolomite to calcite indicates that calcite is the main dissolving mineral and that water from G3 is affected by yet another process adding Na+ ions to the system (Figure 8B).

(29)

18

Figure 8 Expected ratios between the concentrations of DIC and Ca2+ ions for minerals dissolving in various proportions: a) DIC with respect to Ca2+ ions b) DIC with respect to measured Ca2+ ions and exchanged Ca2+ ions, assuming all Na+ ions are from Ca2+/Na+ ion exchange

Lefevre et al. (2015), Appelo and Postma (2005) and Cloutier et al. (2010) attribute waters of type Na-HCO3- (group G2) to the process of aquifer freshening. During the salinization process, Na+ ions, being abundant in seawater, bind to ion exchange sites in the aquifer. As fresh water with abundant Ca2+ ions replaces the salty pore water, the new dominant ion in solution will replace the ions at the exchange sites, in this case releasing Na+ ions. The final step of freshening is when Na+ ions have been completely replaced by Ca2+ and the water returns to Ca2+-HCO3- type. The time required for Na+ to be flushed from the system can provide a relative estimate of the groundwater velocity and thus the age of groundwater. For instance the presence of Na-HCO3 type water indicates that seawater has been completely replaced by fresh water in the aquifer so that it is younger than at the beginning of the freshening event.

As Cloutier et al. (2010) states, three conditions are required for cation exchange to occur as described by Equation (5). First, Ca2+ ion concentrations need to be high enough to displace Na+ ions. This condition is met through the dissolution of carbonates. Then, the aquifer must have a high cation exchange capacity (CEC), which is dependent on the specific surface area and the clay mineral, organic matter and oxide or hydroxide content (Appelo and Postma 2005). According to Cloutier et al. (2010), in the St. Lawrence Lowlands, sites with a high CEC can be found either in rock aquifers or in overlying confined aquifers. In the Chaudière-Appalaches study area, fractured sedimentary rock

(30)

19

formations of the St. Lawrence Platform are dominated by organic-rich black shales (Lavoie et al. 2016) and clay minerals in these formations can act as cation exchange sites (Cloutier et al. 2010). Another possibility is for cation exchange to occur on the clay minerals within till aquitard confining units (Cloutier et al. 2010). Finally, for exchange to occur, exchange sites need to be charged with Na+ ions, i.e. the aquifer needs to have been in contact with water having Na+ as its major cation.

Cloutier et al. (2010), Beaudry (2013) and Lefebvre et al. (2015) postulate that during invasion of the Champlain Sea, saline water also infiltrated into the underlying fractured aquifers. Flushing of these waters with freshwater precipitation would have occurred following retreat of the sea. The source of groundwater salinity is thus attributed to seepage of Champlain seawaters composed of a mixture of fresh and glacial melt water (66%) mixed with seawater (34%) (Cloutier et al. 2010). Evidence of groundwater freshening through the emergence of Na-HCO3- water type was also noted in similar sedimentary aquifers across North America and Europe (Siegel 1991; Pärn et al. 2016; McIntosh and Walter 2006; Ferguson et al. 2007). These aquifers remained continental and glacial melt waters formed fresh lakes. The salinity source in these areas is thus attributed to marine formation waters (Siegel 1991; Pärn et al. 2016; McIntosh and Walter 2006; Ferguson et al. 2007). In the context of the Chaudière-Appalaches region, both the Champlain Sea and

Figure 9 Na/Cl ratios indicating mixing and ion exchange

(31)

20

formation water are plausible sources of the initial aquifer salinity and place the maximum ground water ages of group G2 water to different eras.

2.3.2.3 Mixing

The large range of Cl- ion concentrations in the samples indicates that the rate of water replacement is likely to have occurred at variable intensities over time due to the presence of preferential pathways (rapid flushing) and stagnation zones (slow flushing) (Cloutier et al. 2012). Thus mixing with residual seawater or formation waters accounts for the high Cl -concentrations (Benoît et al. 2014; Lefebvre et al. 2015).

Typical ratios for Na-Cl dissolution and mixing with seawater are represented in Figure 9, where the upward arrow represents cation exchange. Although affected by ion exchange, the G3 water type also results from mixing with saline waters (Figure 9). Using Cl- ions as a conservative tracer for the saline water, the excess Na+ ions likely originate from ion exchange. Similarly, Figure 8 illustrates how the amount of exchanged Na+ cations can be calculated from the expected Ca+/DIC ratio for the dissolution of calcite. When comparing the ions exchanged using the two methods (Figure 10A), it appears that waters from group G3 have an excess of Na+ unaccounted for by the ion exchange with respect to seawater and dissolved halite (Figure 10B).

Figure 10 Comparison of Na+ ions exchanged using Cl- as a conservative tracer with the excess Na+ obtained from the dissolution of calcite: a) Na+ ions exchanged calculated from calcite dissolution with respect to Na+ ions exchanged calculated from surplus of seawater Na-Cl ratio, b) Na+ ions surplus after ion exchange with respect to Cl

(32)

21

The possible natural sources of excess Na+ ions, other than seawater mixing and ion exchange are: water-rock interaction along the flowpath and mixing with formation waters (Hem 1985). The specific geochemistry of formation waters of the fractured rock aquifer is unknown. However, considering the marine origin of the sedimentary formation, it is expected to have a Na+/Cl- ratio close to that of seawater at the time of sedimentation. Alternatively, Na+ ions can also be associated with silicate mineral dissolution (Appelo and Postma 2005; Freeze and Cherry 1979; Hem 1985), the kinetics of which would require residence times of hundreds of thousands of years under the chemical conditions found over the study area (Pärn et al. 2016).

Similarly, the concentrations of conservative ions Li+ and Br- increase with increasing Cl -concentrations and the ratios of Li+ to Cl- and Br- to Cl- (Figures 11 and 12) also exceed those of seawater for group G3 samples, suggesting that G3 water has a relatively long residence time. Li+ to Cl- ratios exceeding the seawater composition are also observed in G1 and G2 samples, suggesting Li+ ions might have been released by ion exchange. However, Br- to Cl- ratios show that while G1 waters tend to be depleted in Br-, some G3 samples are enriched which supports the hypothesis of long residence times.

In addition, Saby et al. (2016) looked at the link between groundwater quality, regional geology and residence times in the St. Lawrence Lowlands. Focusing on the Nicolet watershed, located west of the Chaudière-Appalaches region, they found that long Figure 11 Cl- and Li+ concentrations Figure 12 Br- and Cl- concentrations

(33)

22

interaction times between groundwater and the Ordovician rocks from the St. Lawrence platform and the Appalachian mountains were likely to be an important factor in the release of Ba, F, Fe and Mn ions which were found in anomalous concentrations in residential wells.

Furthermore, they also looked at 3H/3He and 14C isotopes, and found that their samples were composed of a mixture of young water less than 60 years of age with water several thousands of years old that likely infiltrated the Cambro-Ordovician aquifers by subglacial recharge. Supporting the hypothesis of mixing, Vautour et al. (2015) suggest that there are two different water mixtures in the St. Lawrence Lowlands. The first is a mixture of modern infiltration with fossil water from the shallow fractured bedrock aquifer which would have resulted from the infiltration and flushing of the Champlain Sea. The second water mixture is likely a blend between an evolved groundwater which recharged in the Appalachian area and fossil groundwaters from the fractured bedrock aquifer that predate the Champlain Sea invasion episode. In concordance with the conceptual flow model presented in Section 2.2, this suggests that a portion of the groundwater recharging the Appalachians travels along a regional flow path, mixes with fossil water and emerges in the St. Lawrence Lowlands.

The hypothesis of the emergence of a regional flow system in the St. Lawrence Lowlands is also corroborated by Bordeleau et al. (2017) who collected two water samples, in the Saint-Édouard-de-Lotbinière area, near the northern end of the discussed regional flow path, at a depth of 48 m below the surface without disturbing the water column in the well and dated the samples over 1.5 Ma using chlorine-36 isotopes. Based on the hypothesis that these samples could contain water from deeper horizons, Bordeleau et al. (2017) estimated that these samples contained 4% and 7% brine. The study concluded that faults present in the Saint-Édouard-de-Lotbinière study area could act as preferential pathways for the migration of formation waters.

(34)

23 2.3.3 Radiocarbon ages

Some of the groundwater samples collected over the CA study area, 37 by Lefebvre et al. (2015), 10 by Benoît et al. (2014) and 18 by Bordeleau et al. (2017), were also analyzed for isotopes aiming specifically at characterizing residence time (tritium, radiocarbon). In order to gain a better understanding of the characteristic time scales associated with the hydrogeological setting of the study area, the interpretation of the groundwater residence times based on carbon isotopes attributed to theses samples is explored in this Section. The radioactive form of carbon, 14C, is produced in the upper atmosphere by cosmic radiation and enters the hydrosphere under its oxidized form, 14CO2(g). In an open groundwater system, the dynamic carbon cycle will maintain atmospheric 14C concentrations. However, when a system becomes closed to the atmosphere, radioactive decay of 14C can be used to measure the time elapsed since a sample entered closed system conditions (Equation 6). 𝑡 = −8267 × 𝑙𝑛 (𝐴𝑡 𝐶 14 𝐴𝑜14𝐶 ⁄ ) (6) Where,

t= time since the sample entered the closed system A014C= initial activity of the sample

At14C = activity of a sample at time t

Nevertheless, Equation 6 is only valid under two major assumptions and its application becomes problematic when different geochemical processes are affecting the sample. The first assumption when using Equation 6 is that the system is closed to the atmosphere such that no further gain or loss of 14C is possible (Clark 2015). Section 2.3.2.1 presents the challenges of determining whether a sample was collected in a closed or open system. For the purpose of radiocarbon dating in the context of this study, an arbitrary level of pCO2=10-2.5 atm (Appelo and Postma 2005) was chosen as the limit between samples in open and closed systems. In Figures 8.A and 8.C, this represents the level beyond which all

(35)

24

samples are near or over the calcite saturation limit and only the samples below this limit are considered for radiocarbon dating.

The second assumption when using radiocarbon as a dating method is that the initial concentration of the parent (A0) is known (Clark 2015). This can be challenging since relative 14C levels in the atmosphere have fluctuated over time, reaching up to 10% during the Holocene and 20 to 30% in the late Pleistocene (Clark 2015). However, the error related to the fluctuation of atmospheric 14CO2 levels is often negligible with respect to the errors induced by the dilution of 14C through various geochemical reactions occurring in the subsurface (Clark 2015).

In some studies, the 14C concentration of a sample identified to be from a recharge area is assumed to represent the initial activity and is assigned to a group of samples (Appelo and Postma 2005). This assumption, however, can be inaccurate since young waters may have been contaminated by the high 14C activity in the atmosphere during the post-bomb era and may not be representative of the conditions found at the time when the water samples were recharged (IAEA 2013). In the present study, samples from group G1 (recognised as recharge water in Section 2.3.1) and identified to be in an open system with respect to the cut-off level described above, have 14C concentrations between 64 and 104 pMC. Samples with concentrations above 100 pMC are representative of modern atmospheric 14C levels (Levin and Kromer 2004). Conversely, concentrations less than 100 pMC in the recharge area are likely to indicate that geochemical reactions are taking place during infiltration (Montcoudiol et al. 2015; IAEA 2013). Given the large range of 14C concentrations found in recharge zone waters in the CA study area, it is evident that a variety of reactions were affecting their radiocarbon signature. Thus, rather than using a bulk value, the initial activity of radioactive carbon should be adjusted for each sample individually. Several correction methods based on empirical models, chemical balancing, statistics and modelling have been proposed to adjust A0 to account for these processes (Han and Plummer 2016) (Table 1). The choice of correction to apply depends on the processes affecting the water samples and the information available with respect to these processes.

All conventional chemical balancing correction models require knowledge or assumptions with respect to the isotopic composition of soil CO2(g) and the carbonate rock. The

(36)

25

carbonate rocks in the study area are of Paleozoic marine origin and can therefore be assigned 14C=0 pMC and δ13C=0.0 ‰. On the other hand, the value of δ13CO

2(g) in the soil is derived from the decomposition of organic matter, plant root respiration and atmospheric CO2 (Gillon et al. 2012; Amundson et al. 1998) and is highly dependent on the biological and physical conditions prevailing in the recharge area which can vary seasonally (Gillon et al. 2012). In a temperate climate, dominated by C-3 plants, soil δ13CO2(g) is expected to be between -35‰ and -19‰ (Gillon et al. 2009). Samples categorized as being under open conditions can be used to estimate of δ13CO2(g) using Equation 7 (IAEA 2013):

𝛿13𝐶𝑂2(𝑔)=1 𝑛∑ 𝛿 13𝐶𝐷𝐼𝐶−𝑖+ (𝐻2𝐶𝑂3 𝐶𝑇 ∙ 𝜀𝐶𝑂2(𝑔)−𝐻2𝐶𝑂3+ 𝐻𝐶𝑂3− 𝐶𝑇 ∙ 𝜀𝐶𝑂2(𝑔)−𝐻𝐶𝑂3 −+𝐶𝑂3 2− 𝐶𝑇 ∙ 𝜀𝐶𝑂2(𝑔)−𝐶𝑂3 2−) 𝑛 𝑖=1 (7) Where, ε = fractionation factor

i= samples identified as being in an open system

In our study area, the average δ13CO2(g) of the samples categorized as being in an open system is 25.4‰ (Annex A, Table A1). Although using a single value for δ13CO2(g) inevitably introduces uncertainty into the age corrections by not taking into account seasonal variations of δ13C of soil CO

2 (Gillon et al. 2012), in the context of this study, it represents a reasonable approximation.

Furthermore, the concentration of soil 14C also depends on the relative contribution of root respiration with respect to the degradation of organic matter from old or young reservoirs (Gillon et al. 2012). However, these effects are the direct result of nuclear weapons testing and are not important when dating sub-modern groundwater (Gillon et al. 2012). Assuming that the samples were recharged during the pre-bomb era, the soil 14CO2(g) can therefore be assumed to be 100 pMC (Appelo and Postma 2005).

(37)

26

Table 1 Conventional adjustment models accounting for geochemical reactions affecting the initial activities used for the determination of groundwater sample ages, modified from IAEA (2013) (Han and Plummer 2016; Clark and Fritz 1997)

Carbonate dissolution Soil gas dissolution

CO2 - Gas aqueous exchange

Calcite - HCO3- exchange

Gypsum dissolution Ca/Na exchange

Model Description

Fontes and

Garniera x x x x x x

 Mixing of two end members: (i) soil gas CO2

(ii) solid carbonate mineral

 Partial isotopic exchange only between a part of the CO2 gas and calcite

mineral in a closed system.

 Reduces to the Tamers model if there is no isotopic exchange. IAEAb x x x

 Isotopic mass balance in two steps:

(i) equilibrium of the HCO3- with the gas phase (open system);

(ii) dissolution of solid carbonate in closed systems. Evansc x x x

 Accounts for isotopic fractionation during dissolution–precipitation of calcium carbonate.

 Assumes calcite to be in isotopic equilibrium with dissolved HCO3-.

 Results similar to Pearson’s model. Mookd x x x

 Considers partial or complete isotopic equilibration in the soil zone between CO2 gas and all aqueous carbon-bearing species, including part

of the dissolving carbonate mineral in open systems.

 Appropriate in systems in which reactions with carbonate minerals are not dominant.

Pearsone x

Similar to Tamers but accounts for dissolution of carbonates by using isotope mass balance relations for 13C and 14C in open systems.

Eichingerf x x x

 Similar to Pearson model.

 Only includes isotope exchange with the solid phase in closed system. Han and

Plummer (2013)g

x x x x

 Revises the mass balance method of Fontes and Garnier.  Models assume either:

(i) isotopic exchange dominated by gaseous CO2 in the unsaturated

zone which yields similar results to the Mook model,

(ii) isotopic exchange dominated by solid carbonate minerals in the saturated zone, which yields similar results to the Eichinger model, or (iii) if no carbon isotopic exchange occurs, it reduces to Tamer’s model Tamersh x x x

 Assumes that the initial quantity of soil zone CO2 containing 14CO2 is

diluted by dissolution of carbonates in recharge area, with no other sources or sinks of carbon in the system.

 Does not consider carbonate isotopic exchange. Extended

Gonfiantini and Zuppii

x x x x x x

 Statistical approach based on the relationship between 14C and δ13C.

 Applicable if 14C vs. δ13C of sample form exponential curve.

 Can be used to identify outliers from the relationship between 14C and

δ13C data.

a- Fontes and Garnier (1979), b- Salem et al. (1980), c- Evans et al. (1979), d- Mook (1976), e- Ingerson and Pearson (1964), f- Eichinger (1983), g- Han and Plummer (2013), h- Tamers (1975), i- Han et al. (2014).

(38)

27

Different methods used to account for various geochemical reactions in the calculation of 14C ages are presented in Table 1. The ages of the samples collected over the study area were estimated using each correction method (Annex A, Table A2). Discrepancies between ages calculated using different methods can be significant (Figure 13), which emphasizes the uncertainties associates with radiocarbon dating of groundwater and the importance of selecting the appropriate adjustment model.

In order to select the right correction method to be applied for each sample, Han and Plummer (2016) suggest a graphical methodology which allows designating the method that more accurately accounts for chemical and isotopic exchanges affecting the samples (Figure 14). On a 14C vs δ13C plot, the three most accurate and up-to-date correction methods (Pearson, Han and Plummer (2013) and IAEA) are each assigned 14C and δ13C ranges under which they should be applied to appropriately account for carbonate reactions and isotopic exchanges under open and closed conditions (Han and Plummer 2016) (Figure

0 5000 10000 15000 20000 25000 30000 35000 0 5000 10000 15000 20000 25000 Co rr e cte d ag e (yr B .P .)

Uncorrected age based on Equation 6 (yr B.P.)

Pearson Group 1 Pearson Group 2 Pearson Group 3 Pearson Other

Han & Plummer (2013) Group 1 Han & Plummer (2013) Group 2 Han & Plummer (2013) Group 3 Han & Plummer (2013) Other IAEA Group 1

IAEA Group 2 IAEA Group 3 IAEA Other Ext. G.&Z. Group 1 Ext. G.&Z. Group 2 Ext. G.&Z. Group 3 Ext. G.&Z. Other

(39)

28

14). This method was used to determine which correction method is most applicable for each sample. Results are presented in Table A2, Annex A.

The distribution of the sample ages obtained using Han and Plummer’s (2016) methodology is consistent with the geochemical interpretation of the major ion water groups (Table 2, Figure 15). G1 waters are the youngest, having ages between 0 and 4000 years and G3 waters contain the oldest water samples, having ages ranging between 1000 and 12000 years. The maximum age of samples from G2 is 8000 years (younger than the Champlain Sea) which supports the conclusion that water type G2 results from the freshening of the aquifer. The presence of G3 samples with ages corresponding to the era of the Champlain Sea indicates that some water in the study area could have been recharged prior to or during the invasion of the post-glacial sea. However, it remains unconfirmed

Figure 14 Carbon isotope concentrations of samples collected in the

Chaudière-Appalaches region plotted on Han and Plummer's (2016) graphical interpretation of the validity of correction methods

(40)

29

whether the long residence times of G3 waters are attributed to stagnation zones, long travel paths or mixing with older waters such as formation brines.

Table 2 Distribution of radiocarbon groundwater ages estimated using Han and Plummer's (2016) method

Water Type G1 G2 G3 Other

Total number of samples 19 9 23 15 Number of samples assumed to be in open system 11 6 3 10 Number of samples for which Han and Plummer (2016) is not

applicable 1 0 3 1

Maximum age (yr B.P.) 3496 7914 11680 8489 75th Quartile (yr B.P.) 2336 5238 5334 4686

Median Age (yr B.P) 1676 2563 3617 2212 25th Quartile (yr B.P.) 1644 1784 1600 994

Minimum age (yr B.P.) 242 1005 1019 959

0 2000 4000 6000 8000 10000 12000 14000 G1 G2 G3 Other A ge d istr u b tion (yr B .P .) b ased o n H an an d Pl u m m e r (2016) Water type

Figure 15 Distribution of radiocarbon groundwater ages estimated using Han and Plummer's (2016) method. Whiskers represent maximum and minimum values, and boxes represent percent quartiles.

Références

Documents relatifs

Distem allows users to configure latencies for the network interfaces defined on the virtualized nodes.. Moreover the latencies can be set to be different for both incoming and

In order to confirm that AOMs can be considered as beam splitters, we give in the Supplementary Material [URL will be inserted by AIP] a classical derivation of the beam splitter

Hence, we are able to guarantee a control on the distance between the maximum likelihood estimates in the approximate mixed meta-models and the maximum likeli- hood estimates

Chemical vapour deposition and atomic layer deposition of amorphous and nanocrystalline metallic coatings: Towards deposition of multimetallic films Elisabeth Blanquet a , ∗ ,

Chromatographic and spectroscopic studies (UV–Vis spectroscopy and IR spectroscopy) showed that the presence of CPs induced an increase in EPS concentration in the biological

Ces deux tendances, apparaissent comme étant deux mondes distincts l’un de l’autre, non seulement dans la forme et le contenu des textes, mais aussi pour les

Résumé – Acté en 1998, le Programme décennal de restauration hydraulique et écologique du Rhône s’est fixé comme objectif principal d’augmenter les débits réservés dans

énoncée par Michele Hanoosh, laquelle me semble très juste.. transformation com ique d ' un autre texte par distorsion de ses traits caractéristiques [qui passe par la]