SUPPLEMENTARY MATERIAL FOR REACTIVE TRANSPORT AND THE THERMO-HYDRO-MECHANICAL COUPLING IN DEEP SEDIMENTARY BASINS AFECTED BY GLACIATION CYCLES: MODEL DEVELOPMENT, VERIFICATION AND ILLUSTRATIVE EXAMPLE Sergio A. Bea(1), K. Ulrich Mayer(2) and Kerry T.B. MacQuarrie(3) (1) CONICET-IHLLA, República de Italia 780, Azul, BA C.C. 47 (B7300), Argentina, sabea@faa.unicen.edu.ar (2) Department of Earth, Ocean and Atmospheric Sciences, University of British Columbia, 2207 Main Mall, Vancouver, BC, V5T 1Z4, Canada. (3) Department of Civil Engineering, University of New Brunswick, P.O. Box 4400, Fredericton, NB, E3B 5A3, Canada. 1. FORMULATION OF ONE-DIMENSIONAL VERTICAL STRESS IN THE FLUID MASS CONSERVATION EQUATION The fluid mass conservation equation, as implemented in MIN3P-THCm, is: K P g Q t Equation S1 where is the fluid density [M L-3], g is the gravity constant [L T-2], is the porosity [-], P is the fluid pressure [M L-1 T-2], is the dynamic fluid viscosity [M L-1 T-1], K is the permeability tensor [L2], and Q is the volumetric fluid source/sink term [L-3 T-1]. The storage term in Equation S1 (left term) can be expanded as: P t t t t P t t C,T Equation S2 where / P [L2 T-2], and / t C, T [M L-3 T-1] account for density changes induced by liquid pressure, and the temporal density changes due to concentrations and temperature, respectively. MIN3P-THCm does not account for the geometric deformation of the grid as a mechanical load is applied, and it solves the mechanical processes in a simplified way as have several recent studies of ice-sheet induced sediment-hydrogeologic interactions (e.g. Bense & Person 2008, Lemieux et al. 2008, Person et al. 2007, McIntosh et al. 2011). In this approach, we assume that during the ice sheet loading cycle, the load changes are relatively homogeneous per unit of surface area (especially on geologic time scales). In this situation, the changes in vertical stress must be considered; however, lateral stress can be neglected. Following Neuzil (2003), and assuming zero lateral strain, the full fluid mass conservation equation implemented in MIN3PTHCm is: K P Sp zz P g Q t t C, T t Equation S3 where is the one-dimensional loading efficiency coefficient [-], and zz is the vertical stress [M L-1 S-2]. The mathematical derivation of Equation S3 is described below following Neuzil [0] (and references therein) and Bea et al. (2011). Rice and Cleary (1976), among others, have presented the description of porosity changes in terms of compression moduli as: 1 1 v 1 1 P t K K s t K K s K s t Equation S4 where K and K s are the drained and solids bulk modulus, respectively [M L-1 T-2], and v is the mean normal or mean total stress [M L-1 T-2]. Likewise, fluid density changes induced by pressure ( / P ) in Equation S2 can be defined as: P K f Equation S5 where K f is the bulk modulus of the pore fluid [M L-1 T-2]. Substituting Equations S4 and S5 in Equation S2, and rearranging yields: 1 1 1 1 P t K f K s t K Ks 1 1 v t C, T K K s t Equation S6 Here, it is helpful to introduce the three-dimensional pressure-based specific storage coefficient ( Sp3 , [M-1 L T2]), and the three-dimensional loading efficiency coefficient (i.e. Skempton’s coefficient ,[-]) as reported by Neuzil (2003) and Van der Kamp & Gale (1983): 1 1 1 1 Sp3 K Ks K f Ks 1 1 K Ks Equation S7 Equation S8 S p3 Substituting Equations S7 and S8 into S6, and rearranging yields: P S p3 v t t t t C,T Equation S9 The storage term in Equation S9 is simplified in the present work by assuming that the main strain field in a glaciation scenario is purely vertical (as was also assumed by Bense & Person (2008) and Lemieux et al. (2008). Poroelasticity defines strains in the porous matrix in terms of displacements, and the three main components of strain are defined by Neuzil (2003): xx 1 xx kk P TT 2G 1 3K yy 1 yy kk P TT 2G 1 3K zz 1 zz kk P TT 2G 1 3K Equation S10 where xx , yy and zz are the strains in the x, y and z coordinate directions [-], 1 (K / K s ) , G is the shear modulus [M L-1 T-2], is the drained Poisson’s ratio [-], kk is the sum of normal stresses, and T is the linear thermal expansivity of the porous medium [oC1 ]. The assumption of purely vertical stress implies that lateral gradients in fluid pressures tend to be small and that the lateral deformations can be neglected (i.e., εxx = εyy = 0), and Equation S10 can be rearranged to yield (Neuzil 2003): xx yy 2G kk P 2G T T 1 3K Equation S11 The sum of normal stresses ( kk , Neuzil 2003) can be computed using Equation S11, and defining and replacing the quantity 2(1 2) /3(1 ) (Neuzil 2003) and the drained modulus K 2G (1 ) /3(1 2) Rice & Cleary (1976): kk xx yy zz 3P 1 (1 ) zz 4(1 ) G T T (1 ) Equation S12 Multiplying the term with zz in Equation S12 by 1 / 1 , the following equation is obtained: kk 3P 1 1 4(1 ) G T zz T (1 ) 1 (1 ) Equation S13 Note that Equation S13 can be rearranged to yield an explicit expression to mean normal total stress ( v kk / 3 ): v P 1 31 21 2 1 zz 4(1 ) G T T 3(1 ) Equation S14 Neglecting the thermal expansion of porous medium in Equation S14 and substituting it into Equation S9 yields: P zz 1 Sp3 1 t 31 21 2 t t t C, T Equation S15 The new quantity Sp , the one-dimensional specific storage coefficient, can be defined as [0, 0]: Sp Sp3 1 Equation S16 In addition, by defining the one-dimensional loading efficiency as (e.g. Neuzil 2003, Normani 2009): 1 31 21 2 Equation S17 the full fluid mass conservation equation implemented in MIN3P-THCm is obtained (Equation S3) by substituting Equations S16 and S17 into S15. 2. FLUID DENSITY AND VISCOSITY In MIN3P-THCm, density ( ) and viscosity ( ) are considered dependent on both temperature and solute concentrations. Different equations of state were implemented for both and . To describe the dependence of fluid density on solute concentrations, a commonly employed approach relates fluid density linearly to the concentration of Total Dissolved Solids (TDS, [M L3 ]). This option is also implemented in MIN3P-THCm and the density changes ( C ) can be computed as: C TDS TDS Equation S18 where TDS are changes in TDS [M L-3], and / TDS [-] is assumed to be a constant independent of fluid composition. Values ranging from 0.688 to 0.714 are used to simulate seawater-freshwater interactions. Although a linear relationship between density and TDS is commonly assumed when NaCl dominates the salinity (e.g., seawater), the presence of CaCl2-enriched brines in many sedimentary basins requires a more sophisticated model for density calculations, because density will depend on the elemental composition of the fluids. MIN3P-THCm also allows for the computation of C using a non-linear relationship as a function of composition. For instance, C can be computed as the difference between a reference density and a density calculated based on Pitzer's equations according to Monnin (1994) and Bea et al. (2010). Using this approach, density calculations are based on the total volume of the solution that contains 1 kg of water ( V ) (Monnin 1994): V Vid Vex 1000 w mi Vio Vex i Equation S19 where Vid is the ideal volume based on the molar volume of solutes and Vex represents the total excess volume of a multicomponent electrolyte solution. w is the specific volume of pure water (L3 M-1), m i is the molality of the ith aqueous species [mol M-1], and V io is the standard partial volume of the ith solute (L3 mol-1). The term Vex in Equation S5 can be expressed as a virial expansion of the solute molalities: Vex v (I) 2 v v f DH mc ma (Bca ( mc z c )Cca RT c a c Equation S20 In this expression, m c is the molality of cation c (of charge z c ), and ma that of anion a. v is the Debye-Hückel term that is a function of the ionic strength, and B v and C v are the f DH ca ca second virial coefficients for the volume that accounts for the interactions among ions. The density of the solution ( ) is computed according to: 1000 mi Wi i Equation S21 V where m i and Wi are the molality and molecular weight [g mol-1] of the ith solute, respectively. The dependence of density on temperature ( T ), can be computed using a linear relationship: T T T Equation S22 where / T [M L-3 K-1] is a constant, and T is the temperature change with respect to the reference temperature at a reference density. With regard to the fluid viscosity, the temperature and concentration-dependent terms are empirically defined and can be computed as (e.g., see Bea et al. 2011, Diersch & Kolditz 1998): f ,C 1 1.85 4.12 44.53 1 1.85f 4.1f2 44.53f Equation S23 1 0.7063f 0.04832 3 f Equation S24 f ,T 1 0.7063 0.04832 3 where and f are the solute mass fractions in the fluid for the actual and reference viscosities, respectively, and (T 150) / 100 ,with T provided in units of oC. Alternatively, the viscosity-temperature dependence ( f,T ) can be computed based on the expression presented by Voss & Provost (2008): f , T 248 .37 Tf 133 .15 10 248 .37 10 T 133 .15 Equation S25 where Tf is the temperature of the reference water, again with T provided in units of oC. 3. MIN3P-THCm ILLUSTRATIVE EXAMPLE: IMPACT OF A CONTINENTAL GLACIATION ON REGIONAL GROUNDWATER FLOW IN SEDIMENTARY BASINS - MODEL PARAMETERIZATION In the MIN3P-THCm application example, initial porosity was considered to vary with depth based on the expression proposed by Bahr et al. [0]: e cg(s )z (z) e cg(s )z K1 K1 1 z 0 z 0 Equation S26 Equation S27 where z 0 is the porosity at the surface, c is the compaction rate [M-1 L T2], s and are the densities for solid and fluid, respectively [M L-3]. The initial porosity distribution was computed according to the parameters shown in Table S3 using Equation S26 and S27. The hydraulic conductivity was also considered to vary with depth as a function of porosity based on the Karman-Kozeny model: K (z) 3 2 1 ref 1 2 3ref K ref Equation S28 where K ref is the reference hydraulic conductivity tensor [L T-1], and ref [-] is the reference porosity corresponding to K ref (Table S3). Vertical hydraulic conductivity was considered to be one order of magnitude less than the horizontal component. The one-dimensional specific storage coefficient ( Ss ), was computed based on mechanical parameters for the different geological units that are shown in Table S4 according to (e.g., see Neuzil 2003): 1 1 1 1 Ss f g 1 K K s K f K s Equation S29 where K and K s are the drained and solids bulk moduli, respectively [M L-1 T-2], and K f is the bulk modulus of the pore fluid [M L-1 T-2]. The remaining terms in Equation S29 are defined as: 2(1 2 ) 3(1 ) 1 K K Ks E 3(1 2) Equation S30 Equation S31 Equation S32 where is the Poisson’s ratio [-] for drained conditions, and E is Young's modulus representing the stiffness of a rock material [M L-1 T-2] (Table S4). The latter is defined, for small strains, as the ratio of the change of stress with strain. The one-dimensional loading efficiency coefficient ( ), is computed as: (1 ) 3(1 ) 2 (1 2 ) Equation S33 where is the Skempton’s coefficient [-], computed according to: 1 1 K Ks 1 1 1 1 K Ks Kf Ks Equation S34 Following an assumption made by Normani [0], the rock formations were considered incompressible in this context ( Ks ). The bulk modulus of the pore fluid ( K f ) in Equation 9 S29 and S34 was set to that of seawater ( K f 3x10 Pa ), independent of fluid composition and pore water pressure. For the parameterization of the thermal properties, MIN3P-THCm employs a harmonic average for bulk thermal conductivity ( λ ) computed as: λ λ l (1 )λ s Equation S35 where λ l and λ s are the fluid and solid thermal conductivities [E L-1 T-1 oC-1], respectively. The initial solid thermal conductivity distribution was also computed to vary with depth as a function of porosity using parameters shown in Table S5 and Equations S26 and S27. REFERENCES Appelo C, Postma D (1993) Geochemistry, Groundwater and Pollution. Balkema, A.A. Ed., 539 p. Avis J (2009) Hydraulic testing to characterize low permeability sedimentary formations proposed Deep Geologic Repository, Tiverton, Ontario, GeoHalifax2009, 1356-1361. Bahr D, Hutton E, Syvitski J, Pratson L (2001) Exponential approximations to compacted sediment porosity profiles. Computational Geosciences, 27, 691-700. Bea SA, Carrera J, Ayora C, Batlle F (2010) PitzerAlgortithm: Efficient implementation of Pitzer equations in geochemical and reactive transport models. Computational Geosciences, 36, 526-538. Bea SA, Mayer KU, MacQuarrie KTB (2011) Modelling reactive transport in sedimentary rock environments - Phase II MIN3P-THCm code enhancements and illustrative simulations for a glaciation scenario. Technical report: NWMO TR-2011-13. Bense VF, Person MA (2008) Transient hydrodynamics within intercratonic sedimentary basins during glacial cycles. Journal of Geophysical Research, 113, 1-17. Birkholzer J, Zhou Q (2009) Basin-scale hydrogeologic impacts of CO2 storage: Capacity and regulatory implications. International Journal of Greenhouse Control, 3, 745-756. Côté J, Konrad J (2005) Thermal conductivity of base-course materials. Canadian Geotechnical Journal, 42(1), 61-77. Curtis J (2002) Fractured shale-gas systems. AAPG Bulletin, 86(11), 1921-1938. Diersch H, Kolditz O (1998) Coupled groundwater flow and transport: 2. Thermohaline and 3D convection systems. Advances in Water Resources, 21, 401-425. Geiger S, Driesner T, Heinrich C, Matthai S (2006) Multiphase Thermohaline Convection in The Earth’s Crust: II. Benchmarking and application of a finite element-finite volume solution technique with a NaCl-H2O equation of state. Transport in Porous Media, 63, 435-461. Harrison W, Grammer G, Barnes D (2009) Reservoir characteristics of the Bass Islands dolomite in Otsego County, Michigan: Results for a saline reservoir CO2 sequestration demonstration. Environmental Geosciences Journal, 16(3), 139-151. Hobbs M, Frape S, Shouakar-Stash O, Kennell L (2011) Regional hydrogeochemistry, Southern Ontario. Nuclear Waste Management Organization Report. NWMO DGR-TR-2011-12 R000, Toronto, Canada. Lemieux JM, Sudicky EA, Peltier WR, Tarasov L (2008) Simulating the impact of glaciations on continental groundwater flow systems: 1. Relevant processes and model formulation. Journal of Geophysical Research, 113, F03017, 1-18. Lever DA, Jackson C (1985) On the equations for the flow of concentrated salt solution through a porous medium. U.K. DOE Report No DOE/RW/85.100. Mazurek M (2004) Long-term used fuel waste management-Geoscientific review of the sedimentary sequence in southern Ontario.Background Paper 6-12 prepared by RockWater Interactions, Institute of Geological Sciences, University of Bern for the Nuclear Waste Management Organization. Mcintosh J, Garven G, Hanor J (2011) Impacts of Pleistocene glaciation on large-scale groundwater flow and salinity in the Michigan Basin. Geofluids, 11(1), 18-33. Medina C, Ruppa J, Barnes D (2011) Effects of reduction in porosity and permeability with depth on storage capacity and injectivity in deep saline aquifers: A case study from the Mount Simon Sandstone aquifer. International Journal of Greenhouse Control, 5(1), 146456. Monnin C (1994) Density calculation and concentration scale conversions for natural waters. Computational Geosciences, 20, 1435-1445. Neuzil C (2003) Hydromechanical coupling in geologic processes. Hydrogeology Journal, 11, 41-83. Normani S (2009) Paleoevolution of pore fluids in glaciated geologic settings. PhD Thesis, Waterloo, Ontario, Canada. Person M, McIntosh J, Bense V, Remenda V (2007) Pleistocene hydrology of North America: the role of ice sheets in reorganizing groundwater flow systems. Review of Geophysics, 45, 128. Rice JR, Cleary MP (1976) Some basic stress diffusion solutions for fluid-saturated elastic porous media with compressible constituents. Reviews of Geophysics and Space Physics, 14(2), 227-241. Raven K, Novakowski K, Yager R, Heystee R (1992) Supernormal fluid pressures in sedimentary rocks of southern Ontario-western New York State. Canadian Geotechnical Journal, 29, 80-93. Sykes J, Normani S, Yin Y, Park Y (2008) Phase I:Hydrogeologic Modelling. University of Waterloo. Van der Kamp G, Gale JE (1983) Theory of Earth Tide and Barometric Effects in porous formations with compressible grains. Water Resources Research, 19(2), 538-544. Voss C, Provost A (2008) SUTRA-A Model for saturated-unsaturated variable-density groundwater flow with solute or energy transport (Version 2.1). Westjohn D, Weaver T (1998) Hydrogelogic Framework of the Michigan Basin Regional Aquifers System. U.S. Geological Survey Professional Paper 1418. Wolery T, Daveler S (1992) EQ6, A computer program for reaction path modeling of aqueous geochemical system: Theoretical manual, user's guide, and related documentation (version 7.0). UCLR-MA-110662 PT IV, Lawrence Livermore Natl. Lab. Livermore, California. Table S1. Chemical composition of initial and boundary solutions for the verification example: Replacement of gypsum by polyhalite (Bea et al. 2010). Parameter Boundary water pH 7.3 Ca(*) 3.8x10-3 Mg(*) 2.2 Cl(*) 6.48 K(*) 7.43 Na(*) 2.72 Br(*) 10-6 TIC (*, **) 5.1x10-5 SO4(*) 5.6x10-1 Ionic strength(*) 13.83 -1 (*) Concentrations in [mol l ]. (**) Total Inorganic Carbon. Initial water 8.1 1.1x10-2 6.3x10-1 5.94 1.2x10-1 5.178 1.0x10-9 1.7x10-5 2.3x10-1 7.35 Table S2. Model parameters for the verification example involving positive and negative thermohaline convection (Geiger et al. 2006). Parameter Value Unit Positive initial buoyancy Initial parcel temperature Initial parcel brine mass fraction Initial parcel density 300 0.14 831 o C [-] [kg m-3] Negative initial buoyancy o Initial parcel temperature Initial parcel brine mass fraction Initial parcel density 250 0.47 919 C [-] [kg m-3] Reservoir temperature Reservoir initial mass fraction Reservoir liquid density 200 0 875 C [-] [kg m-3] Porosity Permeability Molecular diffusion Formation conductivity Specific heat capacity of rock Specific heat capacity of water Density of rock 0.1 5x10-14 10-8 1.8 1000 4184 2650 [-] [m2] [m2 s-1] [J s-1 m-1oC-1] [J kg-1oC-1] [J kg-1oC-1] [kgm-3] o Table S3. Model parameters for the illustrative sedimentary basin example: Porosity, horizontal and vertical hydraulic conductivities for the different geological materials. Flow parameters Unit Dol3 Sh3 Sand4 Dol2 Sand3 Sh2 Ev Dol1 Sh1 Sand2 Lim Sand1 Lithology [-] (1) 0.02 (1) 0.11 (2) 0.04 (1) 0.02 (2) 0.04 (3) 0.11 (1) 0.08 (1) 0.02 (1) 0.11 (2) 0.04 (1) 0.02 (2) 0.04 Dolostones Shales Sandstones Dolostones Sandstones Shales Evaporites Dolostones Shales Sandstones Limestones Sandstones Weathered Crw *0.01 crystalline rocks Cr Crystalline rocks (1,4)0.01 *Estimated. (1) Normani (2009). (2) Medina et al. (2011). (3) Curtis (2002). (4) Sykes et al. (2008). (5) Avis (2009). (6) Harrison et al. (2009). (7) McIntosh et al. (2011). (8) Westjohn & Weaver (1998). (9) Raven et al. (1998). (10) Harrison et al. (2009). (11) Birkholzer & Zhou (2009). zref [m] 724 559 0 724 0 559 394 724 559 4000 724 4000 KH [m s-1] (6) -7 10 (7) 3.2x10-11 (8) 5.3x10-6 (10) -7 10 (8) 5.3x10-6 (7) 3.2x10-11 (5) -13 10 (5) -12 10 (5) -13 10 *10-5 (9) -12 10 (8) -5 10 Kv [m s-1] *10-8 *3.2x10-12 *5.3x10-6 *10-8 *5.3x10-6 *3.2x10-12 *10-14 *10-13 *10-14 (11) -6 10 *10-13 (11) -6 10 zref [m] 1064 0 0 1064 0 0 394 1064 559 0 724 0 861 (8) *10-11 0 861 (1,4) (1,4) 861 10-10 8x10-12[0] 8x10-13 Table S4.Model parameters for the illustrative sedimentary basin example: mechanical parameters, estimated based on lithology. Mechanical parameters Unit Lithology Dol3 Sh3 Sand4 Dol2 Sand3 Sh2 Ev Dol1 Sh1 Sand2 Lim Sand1 Dolostones Shales Sandstones Dolostones Sandstones Shales Evaporites Dolostones Shales Sandstones Limestones Sandstones Weathered crystalline rocks Crystalline rocks Crw Cr E [GPa] 50 15 25 50 25 15 16.5 50 15 25 45 25 Poisson's ratio [-] 0.15 0.1 0.14 0.15 0.14 0.1 0.35 0.15 0.1 0.14 0.3 0.14 15 0.26 55 0.24 Table S5.Model parameters for the illustrative sedimentary basin example: thermal and chemical parameters. Unit Dol3 Sh3 Sand4 Dol2 Sand3 Sh2 Ev Dol1 Sh1 Sand2 Lim Sand1 Lithology Thermal parameters Volumetric heat Thermal capacity conductivity [J m-3 oC-1] [W m-1 oC-1] [0] [0] 6 1.87x10 3.8 1.87x106 2 6 1.87x10 3 1.87x106 3.8 1.87x106 3 6 1.87x10 2 1.87x106 3.8 1.87x106 3.8 1.87x106 2 6 1.87x10 3 1.87x106 2.5 1.87x106 3 Dolostones Shales Sandstones Dolostones Sandstones Shales Evaporites Dolostones Shales Sandstones Limestones Sandstones Weathered Crw crystalline 1.87x106 rocks Crystalline Cr 1.87x106 rocks (*)Estimated based on lithology. Chemical Parameters Bulk CEC density [meq 100gr-1] [kg m-3] (*) [0] 2900 2.5 2650 5 2800 1 2900 2.5 2800 1 2650 5 2900 2.5 2900 2.5 2650 5 2800 1 2700 2.5 2800 1 2.6 2750 1 2.6 2750 1 Table S6.Geochemical reactions considered in the illustrative sedimentary basin example. Homogeneous reactions Ref. 2HSO 4 H + SO 4 (*) MgOH H H 2 O Mg 2 CaOH H H 2 O (*) (*) OH H H 2 O (*) HCO 3 CO 32 H (*) CO 2 (aq) H 2 O CO32 2H (*) CaCO3(aq) CO32 Ca 2 (*) MgCO 3(aq) CO32 Mg 2 (*) CaSO 4 (aq) SO 24 Ca 2 (*) CaCl Ca 2 Cl (*) MgHCO 3 Mg 2 H CO32 (*) Cation-Exchange reactions MgX 2 2Na 2 NaX Mg 2 CaX 2 2Na 2NaX Ca 2 KX Na NaX K Dissolution/precipitation reactions (**) (**) (**) Enthalpy [kcal mol-1] Halite Na + Cl - -0.918 (*) Calcite CO32- Ca 2 2.585 (*) 8.29 (*) Dolomite 2CO32- Ca 2 Mg 2 3.769 Anhydrite SO 24- Ca 2 (*) Equilibrium constant taken from Wolery & Daveler (1992). (**) Selectivity coefficients taken Appelo & Postma (1993). (*) Table S7.Initial mineral contents considered in the illustrative sedimentary basin example. For the purpose of the illustrative example considered in the present simulation, a highly simplified mineralogy was considered. Clay minerals, which are known to be abundant present in shale were not considered beyond the enhanced effect of ion exchange (see Table S5). Mineral Cr Crw Sand1 Lim1 Sand2 Sh1(*) Dol1 Ev Sh2(1) Sand3 Dol2 Sand4 Sh3(1) Calcite 0 0 1 90 1 9 10 4 9 1 10 1 9 Dolomite 0 0 9 10 9 1 90 36 1 9 90 9 1 Halite 0 0 0 0 0 0 0 50 0 0 0 0 0 Anhydrite 0 0 0 0 0 0 0 10 0 0 0 0 0 Nonreactive 100 100 90 0 90 90 0 0 90 90 0 90 90 minerals (*) Based on rock composition for Queenston, Georgian Bay and Blue Mountain Formations reported by Mazurek [0]. Dol3 10 90 0 0 0 Table S8.Chemical composition of brines and brackish waters considered in the illustrative sedimentary basin example (chemical compositions taken from Hobbs et al. (2011). Formational and boundary water compositions (2) (3) (4) 6 6 6 pH (1) 6 Ca(6) 1.5 0.77 0.2 0.19 6.2x10-4 6.5x10-3 6 5.4x10-4 3.4x10-2 1.2x10-4 2.85 1.8 10-2 4x10-2 4.15 4x10-3 5.7x10-2 3.2x10-4 2.49 9.7x10-3 8.6x10-3 6.24 1.1x10-2 0.07 5.2x10-4 5.77 7.4x10-3 10-2 3.61 5.2x10-5 0.12 1.8x10-3 3 5.6x10-3 8.2x10-4 2.9x10-3 2.6x10-4 6.4x10-5 1.7x10-5 5.2x10-4 4x10-5 Ca-Na-Cl Na-Cl Na-Cl Na-Cl - 7.6 4.97 6.48 3.82 5.2x10-3 0.34 0.24 0.37 0.21 2x10-4 1217 1150 1204 1128 997 -0.78 0 0 -2.62 -7.47 -4.44 -7.64 -2.76 Mg(6) Cl(6) SO4(6) K(6) TIC(6,7) Na(6) Br(6) Chemical signature Ionic strength(6) TDS(8) Density(9) SIHal -0.22 -0.71 SICal 0 0 SIDol 0 0 SIAnhy -0.82 -0.25 (1) Units Cr, Crw, Sand1 and Sand2. (2) Units Lim1,Sh1 and Dol1. (3) Unit Ev. (4) Units Sh2, Sand3, Sand4, Sh3, Dol2 and Dol3. (5) Meteoric water. (6) Units in [mol l-1]. (7) Total Inorganic Carbon. (8) Units in [kg l-1]. (9) Units in [kg m-3]. 0 0 0 0 (5) 6 Figure S1.A), B) and C) Distribution of CaX2 in equivalent fractions at 10,000, 20,000 and 30,000 years, respectively. D) Temporal evolution of equivalent fractions of NaX, CaX2, MgX2 and KX at a selected location in the shallow part of the Sand4 unit. Figure S2. A) Distribution of the porosity changes [%] after 32,500 years. B) Temporal evolution porosity changes [%] at two selected locations.